Kailas Histories Brill’s Tibetan Studies Library

Edited by

Henk Blezer Alex McKay Charles Ramble

volume 38

The titles published in this series are listed at brill.com/btsl Kailas Histories

Renunciate Traditions and the Construction of Himalayan Sacred Geography

By

Alex McKay

leiden | boston Cover illustration: Tibet’s Tise / India’s Mount Kailas. (Courtesy of Toni Huber)

Library of Congress Cataloging-in-Publication Data

McKay, Alex, author. Kailas histories : renunciate traditions and the construction of Himalayan sacred geography / by Alex McKay. pages cm. – (Brill's Tibetan studies library, ISSN 1568-6183 ; volume 38) Includes bibliographical references and index. ISBN 978-90-04-30458-1 (hardback : acid-free paper) – ISBN 978-90-04-30618-9 (e-book) 1. Hindu pilgrims and pilgrimages–China–Kailas, Mount. 2. Tibet Autonomous Region (China)–History. 3. Tibet region–Religious life and customs. 4. Kailas, Mount (China) I. Title.

BL1239.38.C62K3545 2016 294.5'3509515–dc23 2015034536

This publication has been typeset in the multilingual “Brill” typeface. With over 5,100 characters covering Latin, ipa, Greek, and Cyrillic, this typeface is especially suitable for use in the humanities. For more information, please see www.brill.com/brill-typeface. issn 1568-6183 isbn 978-90-04-30458-1 (hardback) isbn 978-90-04-30618-9 (e-book)

Copyright 2015 by Koninklijke Brill nv, Leiden, The Netherlands. Koninklijke Brill nv incorporates the imprints Brill, Brill Hes & De Graaf, Brill Nijhoff, Brill Rodopi and Hotei Publishing. All rights reserved. No part of this publication may be reproduced, translated, stored in a retrieval system, or transmitted in any form or by any means, electronic, mechanical, photocopying, recording or otherwise, without prior written permission from the publisher. Authorization to photocopy items for internal or personal use is granted by Koninklijke Brill nv provided that the appropriate fees are paid directly to The Copyright Clearance Center, 222 Rosewood Drive, Suite 910, Danvers, ma 01923, usa. Fees are subject to change.

This book is printed on acid-free paper. For Jeri McElroy-McKay For her constant support, and not least for encouraging my obsession with a mountain

Contents

Acknowledgements ix List of Maps xii List of Illustrations xiii Maps xiv

Introduction 1

section 1 Indic Histories

1 Mountains and Renunciates: The Early Pan-Asian Cultural Landscape 25

2 “The Play Garden of the Gods … Beyond the Course of Humans” 45

3 Recreating the Divine Order: The Puranic Kailas 65

4 A Tantric Kailas: The Alchemical Trail 92

5 An Early Buddhist Kelāsa 114

6 Kailas on the Edge of Modernity 132

section 2 The Kailas Mountains of India

7 Above the Naga Lakes: Kaplaś Kailas and Manimahesh Kailas 149

8 The Kinnaur and the Adhi Kailas 174

9 Sri Kailas: The Mountain at the Source 200

10 Sri Kailas: The Epic Prototype? 224

Illustrations 248 viii contents

section 3 Tibetan Histories

11 Tibet’s Tise 273

12 Buddhacising the Mountain 304

13 Zhang-zhung, Bön, and the Mountain 339

14 The Tise (Kailas) Deities 364

section 4 Modern Histories

15 The European Construction of Kailas-Manasarovar 375

16 From Theosophy’s Mahatmas to a Globalised Kailas 410

Conclusions 427

Bibliography 453 Index 512 Acknowledgements

This study began with a yatra to Kailas-Manasarovar in late 1986, and in com- pleting a project that has stretched over nearly three decades I am acutely conscious of the great number of people who have assisted me—far more than can reasonably be acknowledged here. Firstly I must thank the institutions that have enabled my work, in particular the ideal model of a post-doctoral centre, the International Institute for Asian Studies, Leiden, The Netherlands, which has provided unceasing support and encouragement for my work. I owe a great deal of thanks to the outstanding, now retired Director Wim Stokhof, to my long-time friend and colleague Senior Consultant Paul van der Velde, and to the support staff. This study formally began when I was a student in the Religious Studies and History departments at the School of Oriental and African Studies (London University). During that time I benefitted from fellowships awarded by the British Academy, the Leverhulme Trust, the Spalding Trust, and the University of London Central Research Fund. I subsequently enjoyed the flexibility of research fellowships at soas and at the Wellcome Trust Centre for the History of Medicine at University College London, and the support of the Namgyal Institute for Tibetology (Gangtok, Sikkim, India), and the Australian National University School of Asia and the Pacific. I am pleased to acknowledge my debt to each of these institutions and academic bodies. At soas i enjoyed support and advice from many individuals: Humphrey Fisher who introduced me to the study of pilgrimage, Peter Robb who demon- strated the practice of the profession of history, and Timothy Barret, Shirin Akiner, and Werner Menski, as well as Tadeusz Skorupski, Daud Ali and Julie Leslie. Dominic Wujastyk gave me unstinting support and Indic expertise at both soas and ucl, and at the latter my thanks are also due to Director Hal Cooke and Administrator Alan Shiel. Of the many librarians to whom I owe thanks, Tim Thomas at the Oriental and India Office Library, Yvonne Sibbald, the Alpine Club Librarian, and Hakan Wahlquist of the National Museum of Ethnography in Stockholm have been particularly helpful. In the Tibetan Studies world I have been helped by many more scholars than I can properly thank, but I must note my gratitude to—in no particular order— Gene Smith, Samten Karmay, A.W. Macdonald, Françoise Pommaret, Geoffrey Samuel, Mona Schrempf, Amy Heller, Charles Ramble, John Bray, Wim van Spengen, Rafal Beszterda, Isrun Englehardt, Dan Martin, John Bellezza, Karma Phuntsho, John Powers, Scott Berry, David Templeman, Christopher Beckwith, x acknowledgements

Mark Turin, Anna Balikci-Denjongpa, Roberto Vitali, Tsuguhito Takeuchi, Christian Luczanits, Ivette Vargas, Sean Gaffney, Diana Cousens, and Robert Mayer. Three others must be singled out for special mention, for they have been a constant source of encouragement, support, and expertise throughout this lengthy endeavour; to Toni Huber, Henk Blezer, and Katia Buffetrille, my warmest thanks. For Indological expertise, I must firstly thank the late Jan Heesterman, un- failingly encouraging of the work of others and the most stimulating of dis- cussants in matters Indic. I am also greatly indebted to Reinhold Grunendahl, Ronald Davidson, Richard Gombrich, Lance Cousins, David Gordon White, Kumkum Roy, Hans Bakker, Peter Bisschop, Frederick M. Smith, Michael Willis, Greg Bailey, Mathew Clark, Kathleen Taylor, R.T.J. Bakker, Arik Moran, and James Mallinson, while I have also benefitted from the different perspectives of Felix Dawe and the late Oliver Winrow. My thanks also to Peter Flugel for his sage guidance in matters Jain and to Carol Dunham for numerous insights into the . In India two incisive historians of the western Himalayas, Laxman Thakur and Mahesh Sharma kindly and generously shared their knowledge of that region, and I was also fortunate to be guided in Manimahesh lore by Molly Kushal and by retired Medical Officer K.P. Sharma of Chamba. My thanks are always due to the Government of India for access to border regions. I am also indebted to Senior Acquisitions Editor Albert Hoffstädt, to Patricia Radder and editorial and production staff at Brill, and to their anonymous readers whose detailed comments were of great value. For technical wizardry I am very thankful to Linden Rudge at Sunne Printing, Taree (nsw) Australia. To those who have accompanied me on my travels to the various Kailas mountains, Danny Shaw and Jacinta (Kailas-Manasarovar), Terry Lehane and Colin Wylie (Kinnaur Kailas), and Richard Beale (Kaplas and Manimahesh) my warmest salutations! Thanks for the companionship and the memories. My thanks also to Mr. Karan, Uttarakashi expert on the region, and to my local guides and porters notably Uttam Singh Kutiyal of Kuti, resourceful Bhadarwah guide Chander Shekhar, my regular Dharamsala driver Sudarshan who calmly shared our experience of a guerrilla war-zone, and finally porters Buddhi Ram and Karam Chand of Kissoram village, Chamba district. To Gangnani pujari the late Siva Prasad and his family I owe a great deal for their assistance, hospitality, and teachings as well as their good name which opened many doors for me in the upper Gangetic regions. Much of the analysis of renunciate activities derives from fieldwork, formal and informal, principally among the Giri order of Śaivite renunciates. I am greatly indebted to these consecrated warriors for their kindness, hospitality, and willingness to share acknowledgements xi their insights and their duni. In particular my thanks to long-time confidant Om Giri Maharaj and to Om Naga Giri, as well as to Gangotri savant Swami Sundaranand. I can only regret that there is much here with which they would not agree. Indeed in sincerely thanking all of the above persons, I acknowledge that they may not necessarily agree with the details or the conclusions of this work, for which I take full responsibility.

Unpublished Crown Copyright documents in the Oriental and India Office Col- lection appear by permission of Her Majesty’s Stationery Office. Some material here was previously published in (1) ‘Asceticism, Power, and Pilgrimage: Kailas- Manasarovar in Classical and Colonial Indian Sources.’ In Alex McKay (ed.) Pil- grimage in Tibet. Richmond: Curzon Press (1998); 165–183, and (2) ‘The British Imperial Influence on the Kailas-Manasarovar Pilgrimage.’ In Toni Huber (ed.) Sacred Spaces and Powerful Places in Tibetan Culture. Dharamsala: Library of Tibetan Works and Archives (1999); 305–321. All photographs not otherwise credited are from the author’s collection.

Alex McKay Belbora 2015 List of Maps

1 The Kailas Mountains of the western Himalayas xiv 2 Western Tibet xv 3 Kailas / Tise circumambulation route xvi 4 Manasarovar and Rakas Tal xvii 5 Kaplas Kailas and Manimahesh Kailas xvii 6 Kinnaur Kailas xviii 7 Adhi Kailas xviii 8 The sources of the Ganges xix 9 The sources of the Ganges (detail) xx List of Illustrations

1 Mount Kailas south face 248 2 Kaplas Kailas 249 3 Manimahesh Kailas 249 4 The Manimahesh massif 250 5 Kinnaur (Kinner) Kailas range 250 6 Kinnaur Kailas peak 251 7 Adhi (or ‘Baba’ / ‘Chhota’) Kailas 251 8 Sri Kailas and surrounds 252 9 Sri Kailas 253 10 Muz Tagh Ata 254 11 Pole shrine (darcho) 255 12 Naga deity Vasukinag and his faithful companion Jitmanji 256 13 Sanskritisation 257 14 Śiva & Parvati at Kailas 258 15 Kailas pilgrimage tangka 259 16 Jigten Gönpo 260 17 The Gangotri glacier at Gaumukh 261 18 The Raktavan (bamak) glacier face at Gaumukh 262 19 Satopanth Tal 263 20 Charles Sherring’s Western Tibet and the British Borderlands 264 21 Swami Jnanananda 265 22 Swami Pranavananda 266 23 Swami Tapovan 267 24 Swami Sundaranand 268 25 Lama Anagorika Govinda 269 Maps

map 1 The Kailas Mountains of the western Himalayas maps xv

map 2 Western Tibet (Ngari) xvi maps

map 3 Kailas / Tise circumambulation route maps xvii

map 4 Manasarovar and Rakas Tal google earth. landsat. 30° 41’49.55 n, 81° 22’56.49 e. july 11, 2013. accessed january 20, 2015.

map 5 Kaplas Kailas and Manimahesh Kailas google earth. landsat. 32° 28’36.80 n, 76° 49’42.19 e. october 4, 2013. accessed january 16, 2015. xviii maps

map 6 Kinnaur Kailas google earth. landsat. 31° 26’49.06 n, 78° 12’24.80 e. october 4, 2013. accessed january 24, 2015.

map 7 Adhi Kailas google earth. landsat. 30° 08’00.39 n, 80° 46’16.74 e. october 4, 2013. accessed january 22, 2015. maps xix

map 8 The sources of the Ganges xx maps The sources of the Ganges (detail) map 9 Introduction

Mount Kailas is the world’s most widely revered sacred mountain. It is a distinc- tive 21,850 feet (6,660 metres)1 high snow-clad peak situated in the centre of a ring of surrounding ranges in the southwestern corner of the Tibetan plateau. On the plains to the south of the mountain are two vast fresh-water lakes, Manasarovar and Rakas Tal, and in close proximity are the sources of four of India’s great rivers, the Brahmaputra, the Indus, the Karnali, and the Sutlej. Hin- dus consider Kailas to be the abode of their great deity Śiva and for Tibetan Buddhists it is home to the powerful Tantric deity Demchok. As one pilgrim describes;

For centuries, Kailas has been to Hindus and Mahayana Buddhists what Mecca is to Moslems, what Mount Olympus was to the ancient Greeks and what Jerusalem was to the medieval Christians … from the age of the Aryans to the age of the atom, Kailas, the holiest of all mountains, and Manasarowar [sic], the holiest of all lakes … [have attracted] … the devout and determined … [who] … have flocked by the millions to revere this holiest of places.2

But the site’s spiritual appeal now transcends religious or national bound- aries with many modernists, particularly Westerners, finding a deeper meaning there. For them Kailas is;

the cosmic pillar that upholds the vault of heaven … an ultimate place of power … radiating its inexhaustible light and energy … a mountain that amplifies prayers … the core of our global structure of life, the sacred heart of joy that makes life worthwhile, or even plausible.3

Pilgrims thus travel to Kailas from all over the world, embracing the idea of its transcendental nature and partaking of a spiritual experience they understand as having endured for millenniums. For them it is an age-old sacred centre described in texts already ancient when the Christian era began, the most

1 Throughout: heights, locations, and mountaineering records cited are those given by the Alpine Club in London. 2 Raghubir Singh (1982: 94). 3 Thurman and Wise (2000: 3–11).

© koninklijke brill nv, leiden, 2015 | doi: 10.1163/9789004306189_002 2 introduction sacred place in the sacred land of Tibet, the centre of arcane wisdom and mystical revelation. It is Kailas, the home of the Gods, a heaven on earth. Indeed, given that few outsiders are worshippers of Śiva, let alone Demchok, it is largely that belief in a greater universal sacrality embracing the extraordi- nary environment that has made Kailas a global phenomena. The mountain has come to embody the spiritual yearnings of people of many beliefs, and none. Indeed so strong, and so appealing is this understanding of Kailas as a universal sacred site for spiritual seekers of all races and creeds that its status has escaped critical analysis; its history has gone unquestioned.

A Critical Kailas

In reality, the belief that Kailas is an age-old pilgrimage site is a myth. As recently as the early 1900s, it was not a site of any great appeal. It was certainly not the most sacred place in Hindu understanding. While Kailas had long been used as a metaphor for strength and purity in Indic literature, the literary Kailas was an ideal rather than a real place. The actual mountain was visited only by the occasional renunciate and not by ordinary Indian pilgrims. Indeed early Hindu and Buddhist texts clearly stated that Kailas was a heavenly place that could only be visited by advanced spiritual practitioners. Others should not, indeed could not, go there. That prohibition was fundamental to the pre-modern Indic understanding of Kailas. Nor was Kailas a place especially exalted by Tibetans, for whom (Gangs)Tise (Ti se), as they most commonly termed it, was just one of a number of moun- tains understood by Buddhists as the residence of a major Tantric deity. Other mountains enjoyed similar status in the Tibetan world and no universally accepted hierarchy ranked them. Indeed the Kailas region was only one centre among multiple networks of both distinct and overlapping sacred geographies that spread across the Tibetan plateau (as well as the Indian subcontinent), and which rose or fell in prominence according to time and circumstance. Most of the known landscape was imbued with religious meanings and associ- ations and Kailas was thus just one sacred mountain in a landscape laden with such mountains, and Manasarovar just one sacred lake among many. Kailas was certainly not Tibet’s premier pilgrimage site. Urban centres such as Lhasa attracted far more pilgrims,4 as did the mountain pilgrimage of Tsari, and the most revered pilgrimage sites for Tibetans were actually those places in India associated with the life of the historical Buddha.

4 On textual sanction for Lhasa as the pre-eminent pilgrimage site, see Smith (2001: n. 736 324). introduction 3

As for Kailas in Western understanding, at the dawn of the 20th century it was unknown to all but a handful of scholars, geographers, and colonial offi- cials. It had no mystical connotations whatsoever in the Western imagination. Such repute as it did enjoy was due to its location in the vicinity of the head- waters of some of India’s great rivers, the exact sources of which had generated some scientific debate. While modern commentators claim ancient textual sanction for the sacral- ity of Kailas-Manasarovar, those early texts did not actually refer to the moun- tain we know today as Kailas. They described a heavenly landscape that bore very little resemblance to the earthly complex. Indeed closer examination of the texts suggests they refer to another mountain located near the source of the Ganges or to other mountains, worldly or archetypal. Similarly, relevant Tibetan literature indicates that the mountain only became a sacred centre less than a millennium ago. Furthermore, that status elevation was controver- sial, with the efforts of one sect of Tibetan Buddhism to associate the mountain with Indian Buddhist cosmological and prophetic schemas strongly criticised by one of the greatest figures in Tibetan intellectual history, the Sakya Pandita (1182–1251). Even the historical focus of sanctity in the region is no simple matter. The earliest Indian Buddhist references to the heavenly Kailas region actually con- cern not a mountain but a lake—commonly assumed today to be Manasarovar. This was consistent with wider Indic beliefs. While there was a central moun- tain in Indic cosmology, bodies of water rather than mountains were more spiritually revered in their earliest traditions. The sources of rivers such as the Ganges carried a particular sanctity in Indic conceptions, and even today Manasarovar rather than Kailas is the ritual focus of Hindu pilgrimage there.5 Tibetan sources, limited as they are, also imply that their earliest sacralities concerned the lake, not the mountain. At the dawn of the 20th century, therefore, Kailas largely existed in the Indic world as a heavenly mountain, one which served as an ideal in Indic culture. It was a symbol of strength and purity, much favoured by poets, and was sometimes associated with the cosmological World-centre Meru, indeed there was a broad conceptual overlap between the two. This ideal Kailas had a long, rich, and dynamic history that reflected, and provides a window into, the cultures of its knowing. The actual physical Kailas, however, existed in a

5 Most of the early modern references to Indian pilgrims indicate that they focussed their rit- ual practice on Manasarovar and did not circumambulate Kailas. See, for example, Swami Tapovan[am] (2001: 160). A prominent renunciate frequently referred to here, he only cir- cumambulated the mountain on his second visit. 4 introduction different imaginary zone, significant only to a small sect of Śaivite renunciates. Although the mountain had come to be regarded as the home of major deities in both Hindu and Buddhist traditions, those deities were also understood to reside or be manifest in other places and other realms. Indeed, in the case of the great deity Śiva there were other Kailas mountains where he was believed to pass his time, for there was not one, but a number of mountains of that name in the Indian Himalayas. The question thus arises as to how this mountain, despite its obscurity at the end of the 19th century,came to possess such elevated status a century later. The answer to that question will be found in what follows.

The principal justification for this work is the absence of any proper historical analysis of the Kailas phenomena. While there are nuanced anthropological studies of Tibetan sacred sites and of their understandings of sacred land- scape,6 historical context is often lacking, while accounts of Indian Himalayan sacred sites are characteristically uncritical.7 Most of the existing literature on Kailas-Manasarovar consists of ahistorical eulogies of place, recycling cliques, mythologies, and imaginary constructions in an on-going process of mutual reinforcement.8 Here we seek to apply historical analysis to the site, which results in very different conclusions to those of the existing literature. This

6 See, for example; Buffetrille (1998); Huber (1999). 7 Kumar (1990: 206) observed that there had been “no significant work” on Indian Himalayan sacred centres, a statement still largely valid more than two decades later; see however Grunendahl (1993). 8 Aside from studies of Tibetan guide-books, the only full-length academic work devoted to the site of which I am aware is the late Nathan Cutler’s doctoral thesis (1996). This was a valuable enquiry in the field of religious studies, but was not intended as a critical historical analysis. Takeuchi (2011: 17–30) includes a précis of early accounts of Kailas by Tibetological pioneer F.W. Thomas (who accepts the Epic, but not Puranic, accounts of Kailas as genuine). The reader seeking reliable if uncritical guides to Kailas-Manasarovar may turn to the popular studies by Pranavananda (1984) and Snelling (1990). Still valuable is Hedin’s exhaustive survey of European knowledge bodies, see Hedin (1917–1922). For the European discovery of the hydrological aspects of the region, see the very readable work by Allen (1982). These works are all primarily orientated to an Indian Kailas; see Buffetrille (2000: 15–105) for a valuable account centred on a 1990 Tibetan guide-book to the site. For a popular account focussing on Tibetan perspectives, see Johnson and Moran (1989). The numerous modern accounts of the pilgrimage by Indians who have, since 1981, trav- elled there under official Sino-Indian auspices form a genre in themselves, the study of which provides numerous insights into issues of class, gender, and literary conventions, on which see Roy (2011). introduction 5 approach incidentally sheds light on the history of several related sacred com- plexes, particularly Gangotri, for Kailas does not exist in isolation but as part of various dynamic and sometimes overlapping networks of sacred geogra- phy. This study also encompasses the neglected topic of the Kailas mountains of the western Himalayas—Kaplas Kailas, Manimahesh Kailas, Kinnaur (Kin- ner) Kailas, Sri Kailas, and the Adhi Kailas (also known as ‘Baba’ or ‘Chhota’ Kailas)—sacred sites that are discussed here collectively for the first time. Only Manimahesh and Kinnaur have attracted even passing academic attention and the general assumption is that these mountains serve as substitutes for the ‘real Kailas’. But each site has its own distinct history and sacred context. Analysing them as centres in themselves, rather than as peripheral to some greater centre, illustrates processes and models both individual and characteristic of all such sites. In addition, we will suggest that the most obscure of them—Sri Kailas, located between the Ganges’ sources—is the strongest contender for an earthly model of the Kailas in the Mahabharata and Ramayana, the foundational texts of the sanctity of a Kailas mountain in Indic understanding. For reasons of historical methodology we limit our search for origins to ascer- tainable cultural phenomena; meaning that our enquiries do not extend back beyond the earliest Vedic sources; i.e., circa 1300bce.9 In addition, for reasons of length we restrict this work to an Asian framework, despite numerous par- allels with sacred sites in other continents. In briefly framing Kailas within the wider Asian phenomena of sacred mountains associated with some form of esoteric practitioners who shaped and articulated the understanding of the sites, we demonstrate that Kailas was just one of many such regional sacred centres. Similarly, we discuss the early Indic world in its wider context in order to illuminate the cultural imperatives that were later to shape the understand- ing of Kailas. These tribal conquests, population movements, and intellectual developments may seem remote from the history of the pilgrimage, yet they are the ideological, economic, and political canvas upon which ideas of Kailas developed. An almost infinitely varied complex of ideas, practices, systems, and events, shaped by the interests of an ever-changing galaxy of agencies and

9 Constructions such as that of Kailas as the axis mundi deriving from Babylonian cosmology provide an explanatory model for certain cross-cultural commonalities in regard to under- standings of sacred mountains. They rest, however, on a now largely discredited chain of historical enquiry deriving from the late 19th and early 20th century “Pan-Babylonianism” popularised by universalists such as Mircea Eliade. For a convincing critique of this lineage, see Korom (1992). 6 introduction power sources produced an on-going series of disparate representations of the sacred site culminating in the modern broad synthesis of the imagined and the real. A fact that is central to the analysis here is that there is no one hegemonic history of Kailas. Until the modern period there were only discrete (albeit organic), bodies of knowledge concerning the mountain-lake complex. These were held by and transmitted within distinct lineages and traditions. Different cults, sects, and regional cultures each preserved their own understandings of the site, which were not concerned to be compatible with the knowledge bod- ies held by others. Nor, particularly in the Indic traditions, were the various representations required to be compatible with geographical realities, which were not generally relevant to the forms of knowledge that had been compiled. Their incompatibility only became an issue in the era of encounter with scien- tific modernity. Within the Sanskritic tradition, there was a process in which the imaginary Kailas illustrated in art, architecture, literature, poetics, and other fields of cul- tural expression which depicted ideals, not realities, was relocated from the heavenly to a worldly realm. The reality of the mountain and its precise loca- tion had been of little relevance in the earlier contexts, and open to various interpretations. (In the later Tibetan tradition a similar—and much debated— division existed between geographical and poetic depictions, contradictions that remained unresolved.) Reconciling the imaginary mountain with disparate bodies of knowledge about earthly mountains known to traders, renunciates and other wanderers took many centuries, and the knowledge took many forms and alighted on many mountains. The first guide-books to Kailas-Manasarovar in traditional Indic form appeared only in the late medieval or early modern colonial period, and in questionable circumstances. They had little or no immediate impact; even in the late 1800s the sacred complex was still visited from India only by a handful of renunciates. Its establishment as an ordinary pilgrimage site came only in the 20th century, under British protection and after indigenous processes and understandings had been fractured by the impact of modernity and foreign rule; Islamic and British. That encounter with the West in particular radically transformed the disparate bodies of knowledge concerning Kailas, synthesising them under European authority. Tibetan historical sources begin much later (c. 7th century ce) than the Indic and while they too suggest that the mountain served as a literary metaphor they are concerned with a specific known geographical feature. Yet as with early Indic literature, most of the earliest Tibetan references concern not the mountain but Manasarovar. This lake, most commonly known to the Tibetans introduction 7 as Mapham or Mapang, was home to Madros, a divine serpent (Skt: nāga, here- after Naga; Tib: klu; hereafter Klu10), and the sacrality of its abode was locally acknowledged. Such ‘Naga lakes’ are still common in the western Himalayas, but if there were early cultural links between Kailas-Manasarovar and the regions to its west they were fractured when central Tibetan forces first took control of the Kailas region in the 7th century ce. Local identities then began to be subsumed in the transformation of the site into Buddhist sacred space and the subsequent political unification of the Tibetan plateau. The process of transforming Kailas into a Buddhist site was a multi-dimen- sional one, most obviously manifest on a mythological level. The great poet- saint Milarepa was said to have claimed the mountain for Buddhism, and later came the imaginal and ritual imposition onto the landscape of the mandala (Skrt: maṇḍala) of the Tantric deity Demchok (the Sanskrit Cakrasaṃvara). This was part of a wider process of transference of Indic Buddhist sacred geog- raphy to the Tibetan plateau by hierarchs of various branches of the Kargyu (bka’ brgyud) sect of Tibetan Buddhism, which, under the patronage of regional powers, became dominant in this area in the 13th century. It was a renunciate of one of the Kargyu sects who was credited with opening the site to popular Buddhist pilgrimage around the year 1215. In the 17th century, Kargyu influence was supplanted in both religious and political spheres by the Geluk (dge lugs) sect, which established a dominant role in the Tibetan polity. The Kailas region, after around seven centuries of political autonomy, was annexed firstly by the kingdom of Ladakh, and half a century later by the expanding Tibetan state. Kailas then became something of a backwater, peripheral to Tibetan state interests and apparently lacking in dynamic agencies.11 But just as the region began to feel the influence of the expansionist Euro- pean powers, Britain and Russia, eastern Tibetan followers of the Bön religion,

10 There are differences in the definition and understanding of the Tibetan Klu and the Indic Naga but for our purposes these differences are not relevant and they have much in com- mon. Both are serpent deities, a category that seem to predate the regional introduction of Buddhism and Hinduism, and both occupy the subaquatic and subsoil underworld of a three-layered cosmos. For a study of the Klu with reference to their medical context, see Vargas (2010). On the Nagas of the western Himalayas, see Handa (2004). The earlier standard source is Vogel (1926). On the Naga in Pali sources, see Rawlinson (1986). 11 That the deity understood to reside there was Demchok, the tutelary deity of the Kargyu sect, may have been a factor in this neglect. His worship was not prominent in Gelukpa- dominated Lhasa, nor was this deity to attract the attention of 20th century Buddhist modernists. (The term “Buddhist modernism” is from Bechert [1984].) 8 introduction seeking their ancient and probably mythical homeland, began to focus their interests there in the late 19th and early 20th centuries. They claimed Kailas- Manasarovar as the heartland of Bön and the centre of an ancient Bön kingdom known as Zhang-zhung, a claim that was to resonate increasingly strongly in the modern world.

The Modern Construction

The modern understanding of Mount Kailas as ‘the most sacred place in Ti- bet’—or in Asia, Hinduism, Buddhism or Bön—actually dates only to the 20th century,12 and Tibetan histories and perspectives had very little influence on it. The roots of the construction were in Indic bodies of knowledge that were reconfigured by colonial and other interests. Most significant here was a British colonial official trying to increase revenue in his obscure Himalayan district, and his construction was advanced by a self-promoting Swedish explorer glo- rifying his own achievements. Their initial construction was then enhanced by a modernist Indian renunciate, one of a small group who responded to colo- nial modernity by reformulating Hindu sacred geography through a creative combination of the scientific and the visionary study of its earliest texts. The construction then took full form through the elegant writings of an erudite Ger- man Buddhist mystic, long-resident in the Himalayas. In addition to advancing the appeal of the site in their own ways, these four very different individuals all imposed a unity on what had formerly been discrete representations compiled under a variety of circumstances by authors from many different world-views. That synthesis was what transformed a local into a World-mountain. How and why that synthesis was constructed, and on what foundation, is demonstrated in this work. An important factor in the construction was that the bodies of knowledge drawn upon by the first two synthesisers of the Kailas pilgrimage were limited to those available in texts already translated into European languages. Indeed none of the four individuals most responsible for the modern understanding of Kailas apparently translated any relevant Tibetan or Indic works. They took, however, the authority to interpret the small selection of texts available to them and to blend the translations into a modern form which they then presented as a timeless and unitary tradition of sacrality. Once their construction was

12 For another example of the modern “rediscovery” of a sacred site, see Burghart (1978); and for an example of how a site may be “invested with new meaning”,see Paramasivan (2009). introduction 9 established, related information was processed within the new paradigm, as Kailas-Manasarovar became a fixed centre to the dominant cultures on both sides of the Indo-Tibetan Himalayas. Lost in this construction process were the voices of the local peoples, which were submerged beneath those of powerful foreign cultures—European, Bud- dhist, Bön, and Brahmanical. All of those who spoke for Kailas-Manasarovar were outsiders; Europeans, south Indians, or central or eastern Tibetans, each of whom had their own images of the mountain realm, their own ‘imagined Himalayas.’13 The traditions and understandings of the local population were articulated,14 if at all, only through the perspectives of those dominant cultures. This meant that local forms of relationship with the landscape were overlaid with those of the World-religions (i.e.; those belief systems claiming world- wide applicability for their salvationist doctrines). Among the apparent conse- quences of that process was the transference of sacral centrality from the lake to the mountain, and the advancement of various claims to spiritual authority over the Naga deity of Manasarovar. Where possible in this work we seek to recover those local beliefs. The con- cept of a Western Himalayan Cultural Complex is advanced to describe the earlier and underlying aspects of the regional culture, the most relevant fea- tures of which are its territorial deities and the Nagas and the lakes with which they are associated.15 It was this culture that was subject to the missionary activ-

13 Bharati (1978: 78) observes that; “[t]o all Hindus, except those who live there, the Himala- yas tend to be ascriptive rather than actual mountains.” 14 Channa (2013: 104) notes the perspective of the Jad peoples of the upper Gangetic region, to whom, “the sacredness of the Himalayas lies … in the very fact of its nearness and familiarity.” It is their dwelling there which makes it “a meaningfully charged space.” 15 These features are of course well known; see for example Oriental and India Office Col- lection (hereafter oioc), mss Eur e321, Sir H.W. Emerson Treatise [hereafter, Treatise]; Chapter xxi. The former colonial administrator observes that the, “distinctive feature of religion in the Western Himalayas is the territorial god.” But Nagas (and stones represent- ing them which are located at sites peripheral to the villages), are equally prevalent. The two categories are not necessarily overlapping, but precise definition of such entities and their spheres may not be possible even in specific local representations; for a critique of attempts to precisely distinguish the overlapping categories of yullha (‘territorial deity’) and gzhi bdag (‘Lord of [a] Place’) in the Tibetan cultural world, see Pommaret (1996: 53). While, as will be seen, it has certain parallels to more formulated models, the Western Himalayan Cultural Complex is used here primarily as a heuristic device that allows local perspectives equal weight with those induced by both modern political boundaries and World-religion discourses. In addition to its frontier character and those religious features 10 introduction ities of Hindus, Buddhists and Bönpos, and which was obscured by the forces of Indic, Central Asian, Tibetan, and even Chinese cultures. But that multiplicity of histories, myths, legends and associations is a part of the status of the dyadic complex, for the sacred is attached to a place through the accumulation of a potentially infinite number of such sacralities. As a powerful spiritual symbol, Kailas existed as contested territory for com- peting religious groups and inevitably attracted political attention. We exam- ine the processes of religious and political territorial accumulation around the Kailas-Manasarovar region, and seek to identify the agents involved, their his- tories, means and motives. Contestations between sects, religions and political powers have helped shape the history of the Kailas mountain(s), and each power has formulated representations that have contributed to its modern status, for in the accumulation of the sacred little is ever discarded. Each addi- tional element only adds to the whole. Given that there is no one Kailas history to be discovered, no early Indic or Tibetan Ur-text from which all understandings of the site derive, then to cite any of the various textual references to the place as more authoritative than another is fundamentally misleading. Each text represents a particular tradi- tion. Similarly we cannot, through comparative analysis of different texts con- cerning the region, ‘correct’ the visions and understanding of a text’s author(s); their variety simply illustrates the different traditions and world-views repre- sented there. We cannot claim the authors intended a particular representation beyond their creation (or that any one author was closer to scientific accuracy than any other), for that was not their intention. Thus while the references to Kailas-Manasarovar that are discussed here are only a selection of those available, I do not believe that others offer any substantial deviations from the conclusions reached.

noted above, its pre-modern (or pre-World-religion) unsystemised form is characterised by primarily worldly goals, and by seasonal rituals with local and often low-caste officiants, trance-based revelation, oral rather than textual expression, and clan rather than state authority. Similar features are of course common in other pre-modern cultures includ- ing those in wide areas of the Himalayas and on other frontiers of the Tibetan plateau (on which see Samuel [1993: 157–198]). Distinctions are primarily of local form and representa- tion, with varying impacts as a consequence of the frontier location between Tibetan and Indic cultures. It must be remembered, however, that the peoples of this region include distinct groups such as the Gaddi, Gujjar, Jad, etc, whose cultures contain distinct ele- ments although they are linked into local economic and social worlds. introduction 11

On Modern Hinduism

This work relies on the now widely accepted determination of the modern Hindu religion as being, to a very large extent, a recent construction which drew together numerous previously diverse strands of belief and practice and pre- sented them as a unitary whole.16 This construction was both a product of and a response to the British colonial state and to the Christian, and more specifically Protestant, definition of religion; a definition that assumed textual authority and a priestly class to interpret those texts, and which presumed monothe- ism as a fundamental Truth. The construction of a Hinduism that agreed with that definition was carried out by both the British and their Brahmanical infor- mants, the class to whom the British naturally turned in their search for an understanding of the local belief system. Perhaps the most significant consequences of Brahmanical influence on the reformulation of Indic beliefs were the privileging of the Sanskritic over the local, the textual above the ritual, and an emphasis on the philosophi- cal position that behind the multiplicity of deities was a single divine princi- ple. Embracing the Vedānta school of philosophy and early Sanskritic textual sources, this reformulation presented a sophisticated belief system that defined itself against a Christian ‘Other’. In political and legal realms, where the colo- nial state was reluctant to interfere in religious structures and practices, this construction enabled the Brahmanical caste to claim indigenous spheres of influence outside of colonial control, thus laying the foundations for what ulti- mately became political challenges to colonialism. Hinduism thus came to be accepted as, and in many ways to be, a monothe- ist religion taught by priests informed by the authority of undeniably ancient texts. Beliefs and practices not conforming to this new paradigm could be categorised as local superstitions, misunderstandings, or even perversions of the original enlightened teachings. Some of the consequences of this process directly relevant to our enquiries are that highly sophisticated bodies of prac- tice such as Tantra were denigrated, traditional links between political elites and renunciates were broken, and the long-standing process of recasting local deities as manifestations of the Sanskritic pantheon gathered pace and author- ity. Recent studies have demonstrated a number of elements of this construc- tion of Hinduism whose claims to antiquity cannot be sustained. We are in- formed that the Allahabad Kumbh Melā took its current form only in the early

16 For a qualification of that position, however, see Lorenzen (1999). 12 introduction

19th century;17 that Ayodhya developed into an important Rāmānadi pilgrim- age centre only in the 18th century;18 that the antiquity of Benares as an ancient Hindu centre is a myth,19 and that Sankaracharya, if he did exist, was not the great Śaivite philosopher of legend, but a minor Vaiśnavite figure.20 Each of those findings represents a fundamental reappraisal of the received wisdom of modern Hinduism and a significant deconstruction of its 18–19th century construction. (Needless to say, these academic positions would be strongly con- tested by the vast majority of devout Hindus today.) In demonstrating the mod- ern construction of the Kailas-Manasarovar pilgrimage in the Indic tradition this work is a further contribution to better understanding of the transforma- tion of pre-colonial Indic belief systems into a modern unitary Hinduism. It is not, of course, intended as a critique of the Hindu religion, rather it is a testament to its dynamic creative impulses; all lasting belief systems are reinter- preted and reformulated according to the social and political needs of the era. Prior to the encounter with the West and the consequent religious systemi- sation, the spiritual world of ‘Hindu’ India was far more diverse than that rep- resented even by the breadth of the Sanskritic tradition. It may be most accu- rate to depict that religious landscape as a series of cults centred on particular deities which were in most cases manifest entirely at a local or regional level. In the Himalayas, for example, belief in the Naga and/or local territorial deities who were understood to rule over the human realm was the predominant reli- gious tendency of the majority of the population. While the Sanskrit-speaking clans were just one of many sub-continental traditions, they had, through a sophisticated system of philosophical and ritual flexibility, the capacity to embrace virtually any other belief system. Local cults could be included within the broad, fluid, and inclusive conceptual frameworks of the Brahmanical ideology by such means as the application of an overarching mythology and the gradual introduction of various Sanskritic rituals, symbols, and social structures into a local world. With the support and patronage of the elite classes, that system then became hegemonic at state level. In the Sanskritic traditions the earliest revelations, attributed to the semi- divine sages who authored the Four Vedas, were the property of individual families or clans associated with particular territories. Over time these bodies of knowledge and associated ritual practice developed within ‘schools’ and later

17 On which see Maclean (2008). 18 Bakker (1986); Van der Veer (1988: 37, 143–146); also see Pinch (2006: 42). 19 Dalmia (1997: 50–82). 20 Clarke (2006); which acknowledges the work on Sankara by Paul Hacker. See, for a collec- tion of Hacker’s papers from 1948–1972, Halbfass (1995); also see Brooks (1996: n. 46 218). introduction 13 in deity-centred sects that formed their own socio-economic world. These lin- eages of transmission were the dna of later Hinduism. While their surrounding structures altered over time, and they were reproduced in different environ- ments, the tradition of distinct bodies of knowledge survived as one of the prin- cipal tendencies in Indian civilisation until their modern reformulation. It is in that context that we situate the Indic Kailas—as a phenomena that was under- stood in different ways by different traditions until the diverse representations were, like so many other aspects of the earlier religious world, formulated into a semi-cohesive whole in the modern period.

On Buddhism and Bön

While influences from Chinese and Central Asian Buddhism may be detected, Tibetan Buddhism primarily emerged from the religion as it was transmitted from India. This Indic origin is important to our study, not least in regard to the Tibetan acceptance of Indic Buddhist cosmology (which was conceptually sim- ilar to that of early Hinduism). In that cosmology, the centre of the physical uni- verse was a mountain, Meru (or Sumeru),21 which in the modern ‘universalist’ construction is said to be identified with Kailas, although that understanding is rarely explicit in Indo-Tibetan traditions. Given their shared roots, the many similarities between early ideas of Kailas in the Buddhist and Hindu traditions are to be expected. But, as will be seen, there are distinct fractures between the Kailas of the early Indic Buddhist sources and those of later Tibetan Buddhism. These fractures were pointed out in internal Tibetan critiques and resolved only, as is the case with so many pil- grimage sites, in their ultimately being part of a multiplicity of understandings of sacrality. Modern scholarship tends to emphasise the extent to which Hindu and Bud- dhist practices overlap in the Himalayas and to downplay divisions between the two.22 While anthropological studies of the practices of non-elite communities support this picture, identities based around the cults of specific deities have

21 See, of a convenient summary of the qualities of Meru, Eck (2012: 122–125). 22 See, for example, Staal (1982a) and (1982b). He argues that the concept of religion is inap- plicable in this region. Religion is not, however, merely a Western academic conceptual device but is fundamental to the politics of identity in Asia; something clearly apparent from the time of the earliest regional text, the Ṛgveda. For excellent studies of religious competition for control of a sacred Himalayan landscape in recent times, see Buffetrille (1994) and (2012). Also see Bouillier (1997). 14 introduction been powerful historical agents in this region. Historical analysis of the inter- play of the systems brings out the extent to which the central institutions and agents of the different traditions were in competition for resources, patronage, and followers. Buddhism in particular strongly articulated a claim to a separate identity, and a distinct linguistic and civilisational divide existed between Bud- dhist Tibetan-speaking communities to the north of the Himalayan watershed and the Indic communities to the south. The extent to which these communi- ties were in communication varied. But while renunciates, particularly those of Tantric or alchemical focus, shared certain sacred spaces and the routes between them their understandings of those sites were very different, and the renunciates practiced almost entirely within their own traditions rather than mediating between them.23 It was the continuing existence of local traditions in the western Himala- yas—particularly those centred around territorial deities—that offered both Buddhist and Hindu agents fertile territory for assimilation and growth.24 Rath- er than followers of the two World-religions engaging in inter-faith dialogue or contention, both primarily directed missionary activities towards followers of the local traditions, seeking to absorb their deities into Buddhist or Hindu pan- theons, mythologies, and identities.25 This was also the case with Bön, the other belief system that has claimed authority over western Himalayan religion. The term Bön occurs in many con- texts in the Himalayas and is applicable to a variety of religious practices and practitioners. But it is used here specifically for that elite textual tradition sys- temised around the 10–11th centuries, apparently as a result of identity-forming processes of contestation with the Buddhist ‘Other’. That systemised Bön reli- gion contained elements of earlier and indigenous spiritual beliefs and prac- tices but emulated, or informed, Buddhist monastic and literary strategies as part of its wider formulation.

23 On Indo-Tibetan renunciate interactions and issues of identity, see Huber (2008: 193–231); also see Templeman (1997). 24 For an anthropological study demonstrating the extent to which Buddhist monastic ritual practice is still primarily concerned with local deities, see Mills (1997). 25 Local traditions in Tibet have been analysed under the rubric of “Folk religion” (Tucci), the “Nameless Tradition” (Stein), or in many cases as within Bön. I prefer the term ‘local reli- gion’ in this context, because religious understandings in the western Himalayas focussed on deities associated with and extending power over, particular defined territories or features of the local landscape. While overlaid with understandings derived from World- religions, including Bön, these traditions remain manifest in our region. introduction 15

That systemised, or “clerical”26 Bön which we discuss here is ultimately concerned with transcendence rather than immediate worldly concerns. Thus it cannot, despite certain parallels, be easily equated to the local belief system of the western Himalayas, which was orientated to worldly concerns. But Bön claimed authority over those local spaces and traditions, and when it moved to (re?)establish itself in the Tise/Kailas region in the 19th century that claim to authority over local sacralities was a central part of its discourse.

The Centrality of Renunciation

In his study of warrior renunciates in medieval India, William Pinch stated that;

I would even go so far as to suggest that [renunciates] stood at the center of [Indian] religion.27

Renunciates certainly stand at the centre of a Kailas history. It was an influx of Kargyu renouncers around the early 13th century that paved the way for its transformation in the Tibetan conceptual world and it was almost certainly Śaivite renunciates who brought the site into the Indic world around the same time, if not later. Again, in the early 20th century, it was a group of Hindu renunciates who reformulated Himalayan sacred geographies around Kailas and the Ganges source at Gaumukh/Gangotri, and who were key agents in the construction of the modern understanding of these sites. Renunciates are thus at the centre of our enquiry. Centralising the renunciates represents a realignment of the usual academic approach to Indo-Tibetan religions. In Tibetan studies the predominance of approaches through Buddhist institutional centres is overwhelming. Studies of Tibetan monasteries, church-state relations, and other such aspects of cen- tralised religion are common, but studies of renunciates as a category in Tibet- an society are a lacuna. While their biographies demonstrate their historical significance, analysis of the historical role of groups such as the Tokden (rtog ldan) is lacking.

26 The term is from Samuel (e.g.: 1993; 7) classifying that “monastic and scholarly” form of Buddhism that reinforces state power; it seems also applicable to systemised Bön. 27 Pinch (2006: 56). 16 introduction

This neglect is partly the result of historically limited access to the remote areas where they practice and also an acknowledgement of the extent to which renunciation as a category was drawn into the institutional structures of Tibet- an Buddhism. But it is primarily due to the Western fascination with the insti- tution of the Dalai Lama, envisaged as the head of a centralised Buddhist state ruled through its monastic institutions. Although that concept of a monolithic centralised state has been convincingly critiqued, a bias in favour of approaches through centralised aspects and institutions of that state remains, and renun- ciation plays only an ideal role in that discourse.28 Similarly, despite their prominence on the Indian landscape, the study of renunciates in the Indic world was, until recent decades, greatly neglected. While this was partly a result of their peripatetic lifestyle, it was primarily due to their negative image in modernist thought, an image for which there were certainly historical precedents in Indic society, but which was magnified under the colonial lens. The first British clash with the renunciates came over economic issues. Renunciates’ trading networks and monopolies represented unwanted competition to the East India Company, which in the late 18th century used its military forces to disperse and destroy that competition.29 In the 19th century, renunciates became the symbol of an esoteric and untamed India that British modernity sought to undermine and abolish. Renunciates also represented an alternative source of power to the colonial state and their itinerancy and religiously assumed identities were in opposition to colonial categorising projects. Perhaps most importantly however, the appearance and behaviour of many renunciates was so far removed from the European ideal of the Christian hermit as to be entirely alien to the Victorian imperial mind. Such common prac- tices as nakedness, sexual rituals, the heavy consumption of cannabis, and dis- plays of self-mortification or human bodily relics such as skulls all grievously offended Victorian sensibilities and a number of these manifestations were consequently criminalised.30 While individual exceptions were made, renun- ciates as a class came to be regarded as shiftless, unreliable, socially para- sitic, and immoral if not criminal. As such, their authority in the religious

28 On political relations in the formation of an image of Tibet, see McKay (1997: esp., 195–217). For an influential critique of the centralised state concept, see Samuel (1993). For a colourful example of renunciate practice in Tibet, see Huber (1999: 87–93); also of interest in this context is Sihlé (2009). 29 On which see, for example Ghosh (1930). 30 See, for example, the discussions in oioc f/4/1951–84977, concerning the Vagrancy Act of 1840. introduction 17 sphere was largely denied by scholarship, an attitude that persisted down to the modern era.31 Brahman priests (paṇḍits), with their fixed abodes, textual sources, and more seemly mores have thus been the main informants of schol- arship. The denigration of renunciates as informants and as a class was an under- standing that was transmitted from the colonial state to Western-educated Indians and for several decades independent India followed the colonial ten- dency to marginalise or criminalise these figures. Reflecting those prejudices, only a handful of Indian scholars worked on renunciates or their traditions.32 More recently, however, renunciates have become a political force in India and scholars have begun to redress the balance of our approaches to Indian religions through work on the renunciate movements and fieldwork among them.33 This work seeks to add to those understandings by demonstrating renunciate agency in the expansion of both Buddhist and Brahmanical tra- ditions, while also demonstrating that renunciates are, historically, regional manifestations of a wider Pan-Asian phenomena. There are numerous terms used to denote this particular category of reli- gious practitioner; a short list from India alone would include munis, sādhus, saṃnyāsīs, j/yogīs, gosains, fakīrs, śramaṇas, and siddhas, each of which has been used historically both as a generic and as a specific term. Given that such usage has varied both in context and over time there is no single established term that embraces all of the manifestations of the phenomena. A general term is needed here, however, and in the absence of any one suitable designation renunciate is (except in a few specific cases), used throughout.34 Although it was not necessarily the case (particularly in the modern period) with other sects or elsewhere, the overwhelming majority of Indic renunci- ates with whom we are concerned here were born into the Brahmanical caste. While many sects (particularly in modern India), accept all castes, those dis- cussed here have been among the more puritanical in this regard. Outstanding charismatic individuals from other castes may have been accepted on occa- sion, but the understanding here is of renunciates as one of two forms of Brah- manical religious occupation, the other being those who maintain the house-

31 See for example, Snellgrove (1959: 214). He describes yogins as “grotesque figures”. 32 See Ghurye (1953); Sarkar (1957); Giri (1976); Tripathi (1978). 33 See, for a pioneering modern study placing motion at the centre of analysis, Burghart (1983). 34 Its use must be qualified, renunciates remain very much a part of the social world. What they renounce varies somewhat, but in general it is the householders’ life and associated caste or duties and prohibitions that they abandon. 18 introduction holder life and its associated caste duties. While divided over purity issues, the two manifestations are essentially siblings with shared interests, although they may or may not approve of the alternative occupation and its associated prac- tices.35 In pre-modern India, renunciates, assumed to be Brahmans, enjoyed high social status. While in many ways feared and mistrusted in general, individ- ual renouncers could be highly esteemed and enjoy access to all sections of society. They were integral to the Sanskritic world, and a balanced study of pre-modern Indic society and culture must take their position into account. Similarly, Tibetan renunciates were important religious agents, notably in the 10–14th century period. Immediately prior to that, the charismatic power and authority of esoteric renunciates apparently threatened the evolving institu- tional forms of Buddhism and while that threat was averted by renunciates being at least theoretically absorbed into the monastic orders, their biographies indicate that many Tibetan renouncers spent very little time indeed within monastic walls.

Sources and Content

In the following chapters we discuss Kailas under four principal headings; (a) Indic histories of Kailas-Manasarovar; (b) the Kailas mountains of the Indian Himalayas; (c) Tibetan histories of Tise/Kailas, and, finally; (d) the modern construction of the sacred complex that rendered it a universal spiritual goal. Sources for this study include Indic and Tibetan texts in translation, relevant academic studies, British imperial records, travel literature, oral sources, and the rigidly empirical accounts of surveyors and mountaineers. Particularly in regard to the Indian Himalayan Kailas mountains, local literature and travel guides have been utilised as appropriate. Fieldwork was carried out at Kailas- Manasarovar in 1986 (the original inspiration for this study), and at each of the Indian Kailas mountains during 2001–2009.36

35 Conclusions concerning renunciates discussed here are not necessarily intended as appli- cable to other such orders, or to Vaisnavite renunciates, who seem less inclined to what are, from an etic perspective, more extreme modes of practice. Nor are we concerned here with the veracity of Brahmanical genealogies. 36 In order to avoid a particular World-religion perspective, to concentrate on ‘reading’ the landscape, and to obtain local understandings, fieldwork was deliberately undertaken at times that did not coincide with major festivals, and enquiries were generally made only from local inhabitants (and semi-resident renunciates). introduction 19

I have used the term Indic rather than Indian except when referring to or implying early Buddhism or the modern Indian nation-state.37 (Discussions concerning historical fluctuations of the borders of India are not intended to concern their modern appropriateness or legal status.) I use the term “San- skritisation” to describe the process by which Sanskritic cultural elements were introduced into and promoted within societies, rather than in the sense famously advanced by M.N. Srinivas.38 In the main body of the work Tibetan language is transcribed into English, syllable markers are thus omitted and capitalisation follows English practice; where confusion might exist the word is transcribed in following brackets according to the Wylie system. The Wylie system is also used in certain foot- notes as appropriate. In the case of Sanskrit, Pali, Pahari or other Indian lan- guages, diacritics are used only on the first appearance of a word or where confusion might arise, and are not used for common personal names or well- known/modern toponyms. Pluralisation is not indicated by the detached ‘s’, which will be obvious to specialists, and throughout the basic principle is read- ability rather than accord with linguistic conventions that in any case do not necessarily accord with local pronunciation. I have chosen, however to retain diacritics in certain cases where this is becoming standard; i.e., the Tibetan Bön (Bon), and the Indic Śiva (Shiva). Common toponyms take modern form, e.g.; Ganges is used for the Sanskrit gangā. The now common spelling Kailas (from the Sanskrit kailāsa)39 is favoured rather than Kailash, although the Pali Kelasa and the Tibetan Tise are used in the relevant chapters. Where it is intended to refer to the mountain-lake complex or is necessary to specifically distin- guish it from the Kailas mountains of the Indian Himalayas, the term Kailas- Manasarovar is used. The titles of texts are italicised but not their wider system; thus Atharvaveda, but Atharvavedic; Tantra, but Tantric.

37 The term is intended to represent a cultural and ethnographic India rather than a political unit with demarcated boundaries, not least because Kailas-Manasarovar is within its cultural, but not political boundaries. The term also serves to imply a world of Sanskritic elites in which the impact of Islam was largely peripheral. Islamic rule, like that of the British colonial state, was largely a phenomena of the Indian plains, and rarely impacted directly on the Himalayas in the contexts with which we are concerned. 38 Thus I follow Staal (1963) (and see Witzel [1997]), in using the term to indicate a dynamic heuristic conceptual device relating to processes in Indian society, particularly here the introduction of the norms, rituals, and values of Vedanta Hinduism. See the discussion in Samuel (2008: n. 11 53, 78). While sustainable in certain contexts, the classic use of the term by Srinivas surely understates the missionary impetus of Hinduism and its coercive power, see Srinivas (1989); also see Seeland (1993: 357). 39 Perhaps from the Sanskrit kelāsa = crystal, but Grünendahl (1993) concludes it is non- Sanskritic. 20 introduction

There are omissions (on grounds of length) that are regretted. There is certainly much here that might be considered in the wider context of frontiers and borderlands for Kailas exists on the interface of cultures in border regions which, far from being peripheral and culturally insignificant are arenas where a continuous negotiation of culture and identity take place.40 But even without venturing into such theoretical areas this work is lengthy, and I am content to locate it within a particular (critical academic) approach to history. Where appropriate, my findings are located within or related to existing theoretical models, but it is not my intention to engage in such issues; I seek a historical narrative. For those reasons of space, the history of pilgrimage from Ladakh and Nepal has been omitted.41 Jain perspectives are only briefly considered and those (apparently derivative) of the Sikhs not at all. Neither are Chinese (or other southeast Asian) understandings of the site considered, although there was a Chinese imaginal construction that drew on Buddhist cosmologies and ‘Silk Road’ informational transmissions. The historical development of maps of the region is also passed over, as are the history of recent decades and the understandings of a modern generation of pilgrims whose impressions are found in the new forms of social media. Kailas is a dynamic process as well as a place, and there will be other histories for other times.42 As this is primarily a work of history it does not seek to articulate insider constructions such as those in Tibetan guide-books, other than in the context of their critical analysis. This reduces colourful reading, but there are numerous accounts of Kailas that represent the believers’ world. It is the absence of a critical history that this work seeks to counter. The author is well aware of the attractions of the emic accounts, having undertaken the pilgrimage from that perspective. But there are distinct incompatibilities between myth and history and between study and direct experience that ultimately cannot always be reconciled. This work is not intended as a challenge to any beliefs and certainly cannot detract from the spiritual experiences undergone by pilgrims to any of the sacred sites discussed. Insiders, including travelling companions, local friends, and spiritual informants who will be disappointed by the findings here might understand it in the following way.

40 For a model study of the Himalayas in this context, see Lewis (1994). 41 I am not aware of studies of pilgrimage to Kailas from either place. On pilgrimages within Ladakh see, for example, Dollfus (1996); and re Nepal see Buffetrille (1994) & (2012); and Tautscher (2007). 42 For a recent account, see Urbanska-Szymoszyn (2011). introduction 21

A Tibetan guidebook to the Kailas pilgrimage describes how;

To the sight of tenth stage bodhisattvas … [Kailas] is made of precious substances and has a height of five hundred yojanas, and a heavenly mansion of gods inside it … To the sight of mediocre people it appears as a splendoured, massive mountain and self-created body of a deity, and so forth, and it has a covering of rainbows, etc. To the sight of inferior people it appears as nothing but just ordinary earth and rock.43

This is, then, a history by a mediocre, if not inferior person. But it is to be hoped that it might be seen as a preliminary study that will stimulate further critical enquiry into sacred Himalayan histories, for these are at the heart of the Indo-Tibetan spiritual worlds and there is much still to be revealed about the multiple layers of spirituality affixed to these regions.

43 Huber (2003: 403), citing the Gangs ri’i gnas bshad shel dkar me long.

section 1 Indic Histories

chapter 1 Mountains and Renunciates: The Early Pan-Asian Cultural Landscape

Introduction

While a globalised world may consider Mount Kailas its most sacred mountain, that conception is the result of particular historical trajectories. It does not represent a Pan-Asian understanding of its status. There are innumerable Asian mountains considered sacred in local, regional, and national perspectives. Many, if not most of them possess characteristics and associations that might, under different historical circumstances, have seen them become pre-eminent. Before focussing on Mount Kailas, we shall briefly note some of these other sites in order to properly locate Kailas in the wider context of sacred mountains in Asia. While we do not have space to discuss these other sites in any depth, even a brief survey demonstrates that many aspects of the sacred status of Kailas are far from unique. Indeed many are generic, not least its close historical associations with esoteric ritualists, or renunciates, and we shall also briefly locate these ubiquitous figures in the history of Iran and China, those cultures most directly impacting on the Indo-Tibetan world. For Koreans, the most auspicious centre in their sacred geography is Mount Baekdu (‘white-headed’). Situated on the North Korean border with China, the 6,000 foot (2,744 metre) peak is the highest point on the Korean peninsula. In accord with the number three being the most auspicious number in Korean geomancy, it is the source of three major rivers; the Tumen, the Yalu, and the Songhua (the largest tributary of the Amur). According to the founding myth of the Korean peoples, the son of the Creator-god descended from heaven onto Baekdu, where he granted the wish of a female bear to become human and fathered her child. Their offspring, Dan-gun, united the Korean peoples and founded their first kingdom, ruling over it for 1500 years before retiring and becoming the mountain god. From at least the middle of the first millennium ce, Korean ruling dynasties held the mountain to be sacred and sponsored its ritual worship. Around the 12–13th century (broadly contemporary with formative constructions of a Buddhist Kailas), in an attempt to promote national unity in the face of threats of foreign invasion, the legend of Dan-gun was promoted in Korean Buddhist texts.1

1 https://en.wikipedia.org/wiki/Dangun, accessed 9 September 2015; http://www.san-shin.org/

© koninklijke brill nv, leiden, 2015 | doi: 10.1163/9789004306189_003 26 chapter 1

But the Manchu peoples also looked to Baekdu as their ancestral home, and not until 1712 did Korea and Manchu China formally recognise the mountain watershed as their border. Baekdu’s sacred associations had long been recognised by Chinese. The Jin and later Qing Manchu dynasties held annual rites for Baekdu, and the Kangxi emperor decreed it the birthplace of the imperial family.2More recently, according to his official biography, the late Kim Jong-il, the ‘Dear Leader’ of the Democratic People’s Republic of (North) Korea, was born on Mount Baekdu in 1942, his birth heralded by the appearance of a double rainbow over the mountain and a new star in the heavens.3 In reality, the emperor Kangxi’s claim lapsed, while Kim Jong-il was actually born in Russia where his father was fighting the Japanese. But these claims to origin on Baekdu demonstrate how sacred peaks may be used as legitimising devices, linking rulers to the traditional origin-place of their peoples, the divine mountain of the ancestors, thus implying their manifestation as the essential identity of their peoples. Even in an atheist state, such claims resonate power- fully across time, reinforcing cultural identities and rooting a community in a territory, as well as legitimising the political interests of the elite classes. In northern Thailand certain mountains are similarly “constitutive of what it means to be northern Thai.”4 One such mountain is Doi Ang Salung. In a late medieval chronicle (the Tamnan Ang Salung: ‘The Legend of Water Basin Mountain’), this peak is identified as Meru, the central World-mountain of Indic cosmology.5 The text also describes the Buddha’s (legendary) visit to the

Dan-gun_Myth.html, accessed 9 September 2015; the texts cited are the 12th century Samguk- Sagi and the late 13th century Samguk-Yusa; Ju Brown and John Brown. 2006. China, Japan, Korea, Culture and Customs. Charleston (nc): Booksurge; 91. 2 Some 2000 years earlier, a Chinese geographical text (the Shan Hai Jing) referred to the mountain as Buxian Shan (‘the Mountain with [a] God’). In 1172 the Manchu Jurchen Jin dynasty bestowed Baekdu with the title, ‘the King Who Makes the Nation Prosperous and Answers with Miracles’. In 1193 it was further ennobled with the title ‘the Emperor Who Cleared the Sky with Tremendous Sagehood’; see Elliot (2000: esp. 603–614); also see, https:// en.wikipedia.org/wiki/Paektu_Mountain, accessed 9 September 2015. 3 http://www.telegraph.co.uk/news/worldnews/asia/northkorea/8292848/The-Incredible -Kim-Jong-il-and-his-Amazing-Achievements.html; http://en.wikipedia.org/wiki/-Kim_Jong -il; accessed 9 September 2015. 4 Swearer, et al. (2004: vi). 5 As part of an Indic cultural ‘package’ Meru is widespread; e.g., Java has a 3,676 metre (Su)Meru peak in East Java; Klokke (1995: 84). mountains and renunciates 27 site, which remains a sacred centre shared by shamans, renunciates on the Indic model, and Buddhist monastics.6 Just as similar associations are made in regard to Kailas, so they are in neighbouring Cambodia and Laos. For the Lao it is Mount Phou Si on the Luang Prabang peninsula that is Meru,7 while Cambodians consider Phnom Kulen, the site of the founding of the Angkor kingdom, to be a mountain visited by the Buddha.8 In a process which will be seen to be common to the Indo-Tibetan worlds, the Buddhist claims involve the subjugation of local understandings of these sites and their overlay by formulations transmitted from Indic culture. This process is ancient. At Ba Phnom, for example, a mountain sacred to Cambodian Buddhists, there are inscriptions that indicate the presence of Śaivites as early as 629ce.9 None-the-less the earlier layers remain visible in textual accounts of the Buddha’s subjugation of the territorial deity of the site and at least at Doi Ang Salung, that local deity remains the subject of an annual ritual of worship. When we turn to China, we find numerous references to local and regional mountain sacralities that predate the rise of World-religions. These also ad- vance elements that are Pan-Asian, albeit with local characteristics that seem specific to Chinese culture. The earliest Sinological records are those of the Shang dynasty, which ruled the northeast of what is now China in the 16th–11th centuries bce. Shang inscriptions indicate that popular beliefs associated with human and agricultural fertility accumulated around mountains. The inscrip- tions indicate that sacrificial offerings, including those “usually reserved for the high god(s)”, were made to the yue, a character which appears to represent a sacred peak(s), the powers of which were considered to include control of the elements and thus the harvest.10 In the earliest period yue may refer to a particular peak in the Shanxi region, with that mountain understood as ruling over the local community and its environment. By the Zhou dynasty (1046–221bce), if not earlier under the Shang, a wider political complex of sacred mountains emerged which articulated an order of ruler, territory and culture. Mirroring the ideal political order in which the king was at the centre of his kingdom surrounded by his tributary nobles each at the centre of their own realms, this model envisaged a supreme central mountain surrounded by the four marchmounts of each cardinal direction, each a sacred centre subsidiary to the primary peak.

6 Swearer, et al. (2004: 22–30). 7 Holt (2009: 48). 8 Harris (2006: 62); Coe (2003: 25, 99). 9 Harris (2006: 62). 10 Kleeman (1994: 226–227). 28 chapter 1

The model was to prove almost infinitely replicable, and by around the middle of the first millennium bce a complex sacred geography had developed in which such mountains were understood to be “intimately linked to the state and its well-being.”11 The model allowed for territorial expansion, with the peaks of the four cardinal directions acting as cultural frontiers. By the Han dynasty (206bce–220ce), the Shanxi mountain Song had emerged as the imperial central mountain, with its four marchmounts drawn fully into the expanding boundaries of the Han state. While elite discourse increasingly bureaucratised the mountain gods’ realms on the Chinese political model, the peaks were also home to local and even trans-regional cults which transacted with the mountains on a more immediate basis. Mountains were envisaged as home to a complex pantheon of gods, spirits, ghosts and divinised humans. The famed alchemist Ge Hong (c. 280–340ce) who visited the mountains in search of herbs, stated that, “[a]ll mountains, whether large or small, have gods and spirits. If the mountain is large, the god is great; if the mountain is small, the god is minor.” But these gods were not necessarily favourable to humans, for, in an understanding that will be seen to be fundamental to pre-modern conceptions of Kailas, the alchemist went on to state;

If someone enters the mountain possessed of no magical arts, he will certainly suffer harm … [from diseases, weapons, acts of nature or wild animals and insects] … One cannot enter a mountain lightly.12

A somewhat later text describes those drawn to the mountains in terms of various categories of renunciate;

Among them were those who did away as hermits, Avoiding speech. Among them were those who practiced quietude, Nourishing purity. Among them were those who employed breathing techniques, Seeking to prolong their existence.13

11 Kleeman (1994: 226–230). On Chinese mountain cults, also see Bujard (2009: 804–807). 12 Kleeman (1994: 230–231); also see the translation in Naquin and Yü (1992: 14). Bill Porter (1993: 18) describes how Chinese hermits entering mountains thus walked the, “Walk of Yu, dragging one foot, as a wounded animal might, to elicit the pity of the mountain spirits.” 13 Brashier (2001–2002: 13), quoting a Baishi mountain text by Hong Gua dating to 181ce. mountains and renunciates 29

The mountains were considered sites of other-worldly power outside the boundaries of Han civilisation. They were thus considered too dangerous to be visited without knowledge of the appropriate rituals of propitiation to the other-worldly powers there. That knowledge was limited to certain ritualists and those they served; a point which will be seen to be of considerable signifi- cance to our thesis. In early Iranian cosmology all mountains were manifestations of Mount Harā, the mountain at the centre of creation. In the Zoroastrians’ sacred text, the Avesta, that peak is clearly other-worldly. It is described as, “much convo- luted, shining … where [there] is no night or darkness, no wind cold or hot, no deadly illness, no defilement [ie., impurity].”14 On Mount Hara lived the solar deity Mithra, whose qualities included those frequently associated with mountain deities; rain-bringer, healer, judge, guardian of treaties and of war- riors. Mithra was worshipped on Hara by Haoma; the deified soma plant, the sacred hallucinogenic said to have first grown on that mountain.15 Many of the beliefs surrounding Mount Hara were to later surround Kailas or Meru. It came to be regarded as the destination of the souls of the departed, and it was also associated with a sea, Vourukaša. This was fed by the river Harahvaitī, which issued from the crystal peak of Hara into two great rivers, which flowed around the earthly continent and returned in a “perpetual rhythm”, giving rise to the understanding that all rivers were one. In time these cosmological rivers were identified with earthly rivers,16 although conceptually Hara remained in the spiritual realms. Sacrificial rituals, along with fire and hoama offerings, were the central feature of early Iranian worship. Temples and image-worship do not predate the 4th century bce,17 and these sacrificial rites were commonly carried out on sacralised ground near rivers or on mountains, with hills or even artificial mounds also favoured centres of cultic practice. This association was sanctified by the Avesta,18 which states that the heights of Hara were ideal for worship, and cites mythological heroes who made offerings there.19 It is clear, therefore, even from this brief survey, that in the earliest historical periods there were numerous concepts in the wider Asian cultural complex that situated mountains within the realm of the sacred. Most of these will be

14 Yašt 10.50, cited in Boyce (1975: 135). 15 Yašt 10.10 & 8.33, Boyce (1975: 147–175). 16 Boyce (1975: 134–137, 145–146). 17 Boyce (1975: 167). 18 Yašt.10.88, Boyce (1975: 167). 19 Yašt.5.21; 25; de Jong (1997: 91, 96–99). 30 chapter 1 seen to manifest in the history of Kailas. Similarly, the wider regional category of esoteric ritualists is prominent, indeed integral, to that history.

Renunciates in Persia and China

In both ancient China and Persia there were esoteric ritualists who existed on the fringes of religious orthodoxy. In Persia there were the Magi, whose pres- ence was essential to the sacrificial rituals.20 Originally hereditary clan-based priests of a tribal confederacy inhabiting the rugged lands southwest of the Caspian sea, they were later known through much of the Near East, includ- ing the Achaemenid empire, which expanded until its easternmost frontier reached the Indus around 500bce. The Magi acted as royal priests and while mentioned only once in the surviving Avesta sections,21 apparently played an important role in worship of the Achaemenid deity Ahura Mazdā and the prophet Zarathustra. Most sources for the Magi are by Greeks (notably Herodotus: 480–425bce), and other outsiders, but there must have been considerable variation in their practices over the many centuries and regions of their activities. There are, for example, references to Magi as bureaucrats or administrators.22 But Magi were associated not only with the court and orthodox ritual, philosophy and recital, but with popular and esoteric practices—divination, dream interpre- tation, rain-making and the potentially transgressive purification and funeral rites. Herodotus describes Magi as carrying out human sacrifices to control a violent storm at sea,23 and in accounts such as those of Plutarch (45–125ce), they appear as a highly transgressive sect practicing inverted rituals of the dominant belief systems. He also associated them with a goddess cult involv- ing sacred prostitution and identified them as being accompanied by prosti- tutes.24 The Magi as a general category were clearly able to serve both royal and pop- ular interests through a broad spectrum of religio-magical practices. They could adapt to changing religio-political circumstances, manifest within different tra- ditions, and travel widely in search of patrons. As Victor Mair concluded, the “Magi were technical experts available to all rulers who might contract with

20 Herodotus 1.132. I cite a translation by William Beloe (1830: vol. i; 104–105). 21 Yašt.65.7, which may be a late interpolation; Malandra (1983: 25). 22 Dandamayev and Livshits (1988: 457–458). 23 Herodotus 7.191 (Beloe: vol. 3; 204). 24 Bausani (2000: n. 53 76); Duchesne-Guillemin (1973: 160). mountains and renunciates 31 them for their services.”25 Mair also brings out another aspect of their char- acter. The Indo-European roots of the Old Persian maguš (Magi) include ‘to have power’ and ‘to fight’.26 Magi were understood to embody spiritual power in the sense of mana, and their control of other-worldly forces was naturally in demand. But they may also have fought on the battlefield for as will be seen, both spiritual and fighting abilities were associated with later manifestations of such characters. In both Iran and China, sacrifice was central to religious ritual.27 While in Iran sacrifice to any deity seems to have been available to any patron, in China a sacrificial hierarchy replicated the social and divine order, thus;

The Son of Heaven sacrifices to the famous mountains and great rivers throughout the empire … the feudal lords sacrifice to the famous moun- tains and great rivers within their domain.28

Among those associated with the sacrifice in China were the wu,29 whose ambiguous character is immediately reminiscent of the Magi. In addition to their sacrificial role, they are associated with divination, control of the weather, and healing: the term later equates to ‘doctor’. As with the Magi we know little of the selection and training of the wu, although again it was probably through family lineages,30 and their role and status again seems to vary with time and place. A later source (the Zhouli) from the Warring states period (481–221bce) places wu in the category of ‘invocators’ (zhu), attached when needed to the Ministry of Rites. The wu transact directly with the deities “averting evil and pollution”, yet they are not regular officials, and seem a “necessary evil” of distinctly liminal status.31 Their duties include participation in rain-making, sacrifices to the deities of mountains and rivers to prevent disease, drought or

25 Mair (1990: 39) states that the “Magi were a professional priesthood to whom Zarathuštri- anism was merely one of the forms of religion in which they ministered against payment.” 26 Mair (1990: 39). 27 Kleeman (1994: 185). 28 Kleeman (1994: 192), quoting the Record of Rites (Li-chi), a Han dynasty compilation of rituals dating to the Warring States period. 29 The term wu itself appears to indicate either “a function or the name of a people (or individual) coming from a definite territory or nation”; the latter apparently outside of Shang cultural domains; Boileau (2002: 354–355). Victor Mair argues, primarily on linguistic grounds, that the wu of the Shang court were Magi; see Mair (1990: 27–47). 30 Boileau (2002: 375). 31 Boileau (2002: 363); von Falkenhausen (1995: 279–300, quotation from 293). 32 chapter 1 flooding, and, in the case of female wu, performing sexually enticing dances to attract the spirits to the funerary rites.32 Significantly the wu were not associated with the principal Chinese act of worship, the ancestral rites.33 Medicine, funeral rites, and animal sacrifice— and associated aspects of their lifestyle—all involved the wu in contact with states and substances considered impure, thus precluding a role in ancestral worship. As will be seen in the Indic context, this association with impu- rity is a crucial signifier of a particular type of religious functionary. It is also explanatory of the Chinese perspective that such practitioners were poten- tially threatening and commonly transgressive. Thus the wu are generally pre- sented in negative terms in Chinese texts which locate them outside of nor- mal society, dwelling in the mountains or in the wilderness among groves of mulberry trees (implying a place of unfettered sexual congress), where they practiced ‘black magic’ and abused their powers.34 Reflecting that view, wu is commonly translated as ‘wizard/witch’ in Western scholarship. The relevance of Magi and wu to the history of Kailas will become apparent; in short they are manifestations of a regional phenomena particularly prominent in Indic soci- ety.

This outline is not intended to be a balanced assessment of either Asian sacred mountains or renouncers, or to suggest (although it may imply), causal con- nections that can rarely if ever be demonstrated. Rather it is intended to show that while there were specific local manifestations of these phenomena, sacred mountains and their association with esoteric ritualists were Pan-Asian con- cepts from the earliest recorded period, common throughout the lands stretch- ing in a broad arc from western Iran to eastern China (and beyond). Certainly in the case of China and Iran in the pre-Buddhist millennium beginning around 1400bce, models, processes and archetypes that are central to the Kailas his- tory are already clearly apparent. This regional approach reminds us that the status of the most sacrosanct mountain is generally ascribed to one with which a people associate their own local, regional, or national cultural and territorial heritage. Long before the rise of World-religions, the concept of a central World-mountain existed in Iranian cosmology at least, while in China certain mountains were identified

32 This practice has not died out, see http://news.bbc.co.uk/go/pr/fr/-/2/hi/asia-pacific/ 5280312.stm; 23 August 2006, which reported the detention of five persons in China accused of organising strip-tease performances at funerals. 33 Boileau (2002: 361–362). 34 Boileau (2002: 370–378). mountains and renunciates 33 as central to earthly polities that claimed to represent cosmological forma- tions.35 Thus there were real or imagined mountains of greater than local sta- tus, reflecting their ideological centrality in the cultural constructions of these lands. We have seen that in China, the understanding of mountains as sacred centres was shaped on the same model as the royal court and territories, and even deities were partially bureaucratised in this model. Mountains remained, however, places of hostile forces that existed outside of the Chinese cultural world and were safely accessible only to those ritually equipped to deal with alien powers. There was no sense that they were suitable places for ordinary Chinese to visit. Iranian mountains seem rather less forbidding, perhaps in reflection of the actualities of geography. Or it may be that our limited sources for Iranian conditions are a factor here—with much of the Zoroastrian sacred texts lost. But in neither culture in the earliest period is there evidence for the establishment on mountains of structures of popular worship (such as temples or images). Those mountains identified as places where sacred power existed were understood as both the mountain deity itself and as the residence of numerous lesser deities, who might be accessed by those equipped to transact with those powers. There were individuals so equipped, in both China and Iran we may identify a class of esoteric ritualists of ambiguous social and political status closely associated with magical powers and esoteric or transgressive practices. These ritualists, understood to possess ‘magical’ powers that they could utilise on behalf of both individuals and states, were not associated with temples but were qualified to dwell and practice among the mountains. As possessors of such powers, they were inherently threatening to the established order, yet essential to statecraft. Even in this early period their role in transacting with impurities meant they must have existed in a paradoxical relationship to other ‘pure’, or ‘clerical’ religious functionaries. Prior to the rise of World-religions, acknowledgement of various forms of territorial deities, including those associated with mountains, was common throughout Asia. These deities were understood to hold power over their own realms, which might be limited to a single mountain, lake, or other local geo- graphical feature. Other deities with power across territory were acknowl- edged, and these territorial and deities of place seem fundamental to these

35 On pilgrimage in China, see in particular Naquin & Yü (1992: esp. 1–33); also see Bernbaum (1984). On early Chinese constructions of major Buddhist mountain pilgrimages, see Huber (1999: 29). 34 chapter 1 cultural regions in our earliest sources (and, as will be seen, remain prominent in the Himalayas today). While fundamental to Asian religious worlds, local or territorial deities have not attracted Western practitioners of Asian religions, and have been neglected by Western scholarship in favour of the study of texts focussed on belief rather than practices. In locating these concepts at the centre of our enquiry, however, we may begin to perceive a socio-religious world rather different from that con- structed through the now well-known processes of colonial era-scholarship, ‘Protestant presuppositions’, indigenous elite informants, and ‘Orientalist’ power paradigms. Those constructions privileged perspectives and under- standings advanced by the orthodox religious elites in the urban centres of Asia, views which drew support from texts used by those elites. The view from the mountain, however, was somewhat different.

The Indic World

Turning to the early Indic world, we see, as in contemporary Iranian and Chinese cultures, indications that mountains had sacred associations. So too was there an esoteric priestly class of liminal social status, which is prominent in the earliest Indic sources. These sources, texts known as the Vedas (Skrt: veda = ‘knowledge’), are where we begin our enquiry, for it is the Vedic peoples in the northwest of the subcontinent who provide the earliest references to the Himalayas. The subcontinent had been populated for several hundred thousand years before the Vedic period, and a major urbanised culture had existed along the Indus valley from c. 2300–1700bce. With its script yet to be deciphered, how- ever, our knowledge of its culture and world-view is largely conjectural. But we can identify significant elements of later Indic beliefs as part of the cultural heritage of the Vedic peoples who called themselves Ārya. These ‘Aryans’ were part of waves of Indo-European language-speaking groups who, around 1900– 1700bce, reached the sub-continent by diverse routes (notably via Iran and Baluchistan and via Bactria/Margiana/Swat). Indo-European tribal structures were based on three patrilineal social classes; warriors, agriculturalists, and a sacerdotal class of bardic priests who preserved tribal lore and spiritual knowledge. It was these bards who first recorded the traditions of a group that settled in the Punjab. Their collation of the disparate tribal lores and bodies of knowledge, initially as oral and later as written texts, enabled the sacerdotalists to reserve for themselves the authority to interpret the texts. Thus they became an intellectual elite, a group of special- mountains and renunciates 35 ists whose monopoly over language and ritual ensured the social power and authority of their class.36 The first of their Sanskrit texts, the Ṛgveda, was fixed over several cen- turies from around 1300bce.37 A liturgical work comprising ten collections (maṇḍalas) of hymns in praise of the Indo-European deities and the cosmic order the gods protected and maintained, it also contains some evidence of Vedic history. Among its pantheon of deities, for example, are several known in Iranian texts, evidence of the eastward migration of Indo-Iranian elements of Vedic culture. In addition, some hymns are attributed to composers with non- Sanskritic names, indicating an on-going synthesis of cultures and interaction with other language or ethnic groups. We may understand accounts of the martial exploits of Vedic deities as references to military triumphs over local or other Indo-European tribes.38 Indra, for example, one of the great Vedic gods, also represents a tribal chief, quaffing the hallucinogenic draught pressed from the soma plant,39 smash- ing his enemies and destroying their forts as he leads his forces eastwards. A charismatic figure, a god of war and of weather, a colourful martial hero, fond of drinking and feasting, Indra was a deity with whom the warrior class (kṣatriya) must have identified. Others, such as Soma, the anthropomorphised god of the sacrificial hallucinogenic, were more associated with the priestly class. So too was that power the Rgveda only hinted at, a mystical charis- matic force which gave its name to the Vedic sacredotalists, who became known as brāhmaṇs (hereafter Brahmans), the possessors of Brahman (‘uni- versal soul’). Tradition attributed the original composition of the Vedic hymns to the divinely inspired ṛṣis, the Brahman poet-seers associated with residence in the Himalayas.40 Already legendary figures when the Vedas were compiled, the rsis are portrayed as sages practicing austerities (tapas: literally ‘heat’), ‘leading

36 Carpenter (1992: 58–64). 37 Gonda (1975: 12, 22–23). 38 These opponents included other Indo-European tribes, for they understood each others’ language; see Parpola (1988: 219 & passim). 39 The exact nature and identification of soma has been subject to considerable debate, summarised in Falk (1989: 77–90); for a study in the context of Indo-European migration, see Parpola (1995: 353–379). 40 On the composition process, see Elizarenkova who states (1995: 17–20) that the root of ṛṣi is arṣ-árṣati; “to gush forth, flow swiftly, rush forth”; alternatively from the root rs; “to go, to move”, while the Pali ṛṣis derives from is, “to seek”; Norman (1983; 319). The element of motion is predominant, and characteristic of the rsis and their successors. 36 chapter 1 to heaven.’41 What is important to our history is that the visionary teachings ascribed to particular rsis were claimed as the heritage of families or groups of Brahmans,42 who specialised in those particular teachings and associated prac- tices, and who maintained authority over the teaching lineages.43 Particular lineages were closely associated with the specific territory of the tribes whose Brahmans maintained them,44 enhancing the development of sacred associa- tions with tribal centres. The powerful Kuru tribe who settled Kurukśetra (by the Sarasvati river), for example, claimed their capital to be a World-centre so significant that even the gods sacrificed there!45 Such geographical references indicate that the Rgveda was composed when its people’s territory was centred around the ‘five rivers’ region of the Pun- jab.46 A later Rgvedic hymn (x.121.4) provides the earliest indications of an Indic knowledge of the great mountain chain(s) to the north, mountains termed himavantaḥ (‘snow mountain/s’).47 They were the source of soma and medic- inal herbs,48 and the Vedic peoples acknowledged a minor god, Rudra, who dwelt there.

41 Rgveda x.109.4 and x.154.2. 42 On which see Brough (1953: esp. xiii, 20–21). 43 Gonda (1975: 9); Rgveda x.109.4; Smith (1986: 76), observes that the “… body of knowl- edge absorbed by a student during brahmacārya [learning phase] was instrumental in contributing one’s identity … One was defined by the branch of the Veda, the śākhā to which the teacher was affiliated. Individuals, like texts, were in ancient India associated with one veda and within that veda, one recension or school.” 44 Witzel (1997a: 259). 45 Witzel (1997b: 15–16). 46 The rivers were the Jhelum, Chenab, Ravi, Beas and Sutlej. With the addition of the Sindhu (Indus) and Sarasvati, these became a set of seven sacred rivers, which set was later to include the Ganges and Yamuna rivers or be transformed into regional models embracing various rivers; Eck (2012: 167–168). 47 himavantaḥ = “the snowy (mountains)”, is the Vedic term (i.e. Ṛgveda x.121.4). In later Sanskrit the term used is himālaya = “abode of snow”; Staal (1982: 41). Despite evidence for trade with Central Asia and northward population movements in late Indus Valley culture, there is no archaeological evidence for Harappan settlements further northeast than Rupar, a frontier outpost in the Himalayan foothills on the Sutlej river. Archaeological excavation at Rupar has not produced any evidence for trade with the plateau, see the excavation reports in Indian Archaeology (1954; 6–7), and (1955; 9–11). On Harappan trade, see Dikshit (1984); Joshi, Bala & Ram (1984: 511–530); Sharma (1955–1956: 121–129). For a thought-provoking enquiry into possible Harappan links to the Tibetan plateau, see Samuel (2000: esp. 657–658). 48 On soma being found in what is now the Muzh Tagh Ata region, see Witzel (2004: 594–597); also see Witzel (1980: n. 16 &n. 17 104–105). mountains and renunciates 37

A fearsome and amoral deity of uncertain temper, Rudra dispensed both disease and medicines.49 But while sharing Indra’s command of storms and thunderbolts, the Vedic Rudra remained a shadowy figure, lacking Indra’s lively and more rounded character. Rgveda i.114 describes Rudra as a god with braided hair ( jaṭila) ruling over heroes (a description later applicable to Ṥiva in his mountain-dwelling form). Rudra, however, was not among the deities invited to the great sacrifices at the heart of Vedic culture and is described as the patron of lower castes,50 an element of impurity in his manifestation likely indicative of a local deity being drawn into the Indo-European world. If these references suggest only a vague knowledge of the Himalayas, they do indicate that the Vedic peoples imagined themselves as being in a sacred relationship with the mountains. They were described as tribal guardians, or fathers, who possessed the power to protect the tribes. Thus we read in Rgveda vi.52.4 that: “My guardians be the firmly seated mountains, the Fathers, when I call on God, protect me”. We also read that Indra’s followers are sent to “fixed abodes” in the mountains, indicating a concept of sacred landscape in which himavantah was imagined as a home, or heaven, for those blessed by the gods.51 In addition to these otherworldly mountains, there existed a complementary and apparently even stronger sense of the sacred nature of rivers, the “purify- ing waters” that were understood as divine entities. While lakes do not seem to have been attributed any specific other-worldly qualities there were, in addi- tion to the sanctification of the five rivers of the Punjab, three rivers that were singled out for particular sanctity, the Sarayu, the Sindhu (Indus), and in par- ticular the then mighty Sarasvatī (now much reduced), which is praised as a goddess in two Rgvedic hymns.52 This concept of a divinised landscape implic- itly indigenised its peoples; rivers, mountains, and tribe were imagined as a family—the mountains as “Fathers”, the rivers as “Mothers”.53 The importance of the rivers is also indicated by the fact that the course of the Indus was known to the Ṛgvedic composers. In Rgveda ii.15.6, a hymn in praise of Indra, the deity is exalted for having, “made the Indus through his power flow in the northern direction”. The Indus does indeed turn northwards near its traditional source, which is around 9 miles (15 kilometres) north of Mount Kailas. This verse implies a Vedic encounter with Kailas, for as Frits Staal

49 Atharvaveda 6.90 and 7.42. 50 Yajurveda 16.1; also see for early textual citations of Rudra, Modak (1993: n. 589 424, n. 614 425). 51 Rgveda iv.54.5. On the Vedic concept of sacred landscape, see Ali (1973: 17). 52 Rgveda vii.49.1, and x.64.8–9, quoted in Kane (1953: 555–556). 53 Rgveda x.64 and x.17.10. For a gender perspective, see Pintchman (1994: 22–29). 38 chapter 1 has pointed out, the particular sanctity attached to river sources means that if the Vedic peoples travelled far enough along its course to learn that the Indus flows north, “they could not have failed to follow it to its source”.54 It seems, however, that the Indus river’s northward change was more remarkable to that culture than the sight of Mount Kailas,55 for the mountain is not specifically mentioned. While there is just one, probably late, mention of the Ganges river in the Rgveda (x.75), Vedic culture spread eastward into the agriculturally rich Gang- etic plains in the centuries following its composition. There political divisions based on territory rather than clan started to emerge, and urban centres began to develop around the settled courts of the former clan leadership.56

Early Indic Renunciates

In identifying themselves as compositions deriving from the visions of the rsis, the Vedic texts indicate that individuals within a social model of renunciation were a continuous authoritative presence in the sub-continent from the earli- est period of Indo-European settlement. While renunciates may have existed in the sub-continent in an earlier period, rsis were an integral—and ideal—part of Indo-European culture.57 We saw that a liminal class of esoteric religious specialists also existed in early Persia and China, but historically renunciates have been particularly prominent in the Indic world. Indeed their contempo- rary manifestations remain prominent figures on the Indian landscape today and certain aspects of their distinguishing characteristics seem to represent continuities with earlier forms, as will become apparent. The Vedas indicate that from the earliest period there existed a division within the Indo-European priestly class. Married householder (gṛhamedhin) Brahmans were required to maintain ritual purity. Other Brahmans were thus needed to deal with impure fields that had not been ritually sanctified—such realms as medicine, funerals, and the exploration and incorporation of new ter-

54 Staal (1990: 290). 55 The sacred city of Kāśī/Benāres/Vārāṇasī is situated at the northward diversion of the Ganges. 56 On which see Roy (1993: 1–32). 57 On which see Heesterman (1988: 251–271); also see Heesterman (1993). As Kaelber (1989: 106–109), points out in a valuable discussion of the issue, Heesterman’s model of renun- ciation emerging directly from the sacrificial system is not necessarily incompatible with models emphasising “challenge and assimilation” from outside the system. mountains and renunciates 39 ritory into the Indo-European world. Specialists in those impure areas included those who not only preserved the rsis’ teaching lineages but directly followed their social model of itinerancy, in contrast to householder Brahmans restricted to orthodox society.58 As well as rsis, the Rgveda mentions other forms of renunciate such as munis and yatis,59 while another such group, the vrātya, are the subject of an arcane chapter in the last of the four Vedas, the Atharvaveda. The vratya were eso- teric ritualists who formed itinerant bands to make raiding expeditions beyond established tribal territory.60 Their identity seems to cross castes (or more pre- cisely varna) and embrace ritualist and renunciate, suggesting that they pre- date the full emergence of these divisions.61 Vratyas did not maintain Brahman- ical purity and were, like the Magi, described as being accompanied by prosti- tutes. But they could be (re-?) included in the Brahmanical fold after a simple purification ritual (vrātyastoma).62 Jan Heesterman famously defined this as originally a rite of preparation for, and celebration of, the vratyas’ annual raid- ing campaigns. He also identified these long-haired wanderers as the ancestors of Brahmanical sacrificers (dikṣita), themselves rendered temporarily impure by their association with sacrificial killing.63 Central to the identification of vratya with sacrificer in Heesterman’s model is the element of motion characteristic of vratya (and rsi) practices and sym- bolically present in sacrificial rituals. Indeed the early sacrificer also set out on a chariot seeking the resources necessary for the sacrifice, journeys “thinly dis- guised as a religious begging rite”, but essentially of violent intent. While such journeys were later rendered symbolic, the original diksita was thus, like the vratya, a “consecrated warrior”, whose role involved ritual movement in which “the underlying notion … is that of a full-scale conquering or raiding expedi- tion.”64

58 Travel beyond the boundaries of Brahmanical territory brought a loss of ritual purity; Witzel (1980: n. 26 105); also see Bharati (1965: 62), who states that, “Caste-Hindus lose their caste when they cross the northern mountains just as when they cross the ocean.” 59 In Rgveda x.136, the munis were said to have “drunk the magic cup of Rudra … poison to ordinary Mortals”; Heesterman (1962: 9) observes that ‘poison’ in these texts usually refers to food from an impure, i.e., non-Brahmanical, source. Also see Warner (1989). 60 For a recent discussion of the vratya situating them in their wider social and historical context, see Samuel (2008: 127–128, 183–186, 237–242). 61 Heesterman (1962: 8, 32); also see Macdonell & Keith (1912: 342–344). 62 Gonda (1975: 306); Macdonell & Keith (1912: 342–344); Kane (1953: 385–387); Heesterman (1962: 4–7). 63 Heesterman (1962: 4–11, 16, 18–24, 27–29, 32–34). 64 Heesterman (1988: 254–257); also see Heesterman (1962: 16). Given that the vrātya are 40 chapter 1

Heesterman’s vratyas are located in a “pre-classical” period before the final redaction of the sacrifice,65 but his understanding enables us to see the sanctifi- cation of ritual movement. Like the aśvamedha (horse sacrifice), these journeys acted as a means, or rite, of territorial expansion, a pattern of trekking and set- tling, an “unending cycle of violence” to gather resources which, if successful, ended in a ritual enabling the sacrificer to settle as a wealthy landlord on the conquered territory.66 It was such figures equipped to deal with impurities who encircled and brought outlying and impure territory into the pure and known world of their society, paving the way for its settlement and development. Their various rites legitimised and sacralised territorial acquisition and provided the land-owning basis for later elites. Such expansionism did not, of course, go unopposed,67 and a growing short- age of land available to conquer must have eventually restricted these formal rites of expansion to less accessible areas (such as the Himalayas), and neces- sitated their transformation from actual to symbolic rites of conquest. What is important to our study is that these processes demonstrate the cultural prece- dents for the journeys of renunciates beyond known and settled territory, their purification of unknown territory, and the acceptance of violence as one poten- tial necessity of their journeys.

Sacrifice and the Brahmanical World

As in the classical Chinese and Persian worlds, the central feature of Vedic religious practice was the sacrifice (yajña; from yaj, ‘to worship’). The major sacrifices were awe-inspiring public ceremonies, conceptually embracing the whole universe. Generally sponsored by the merchant class or warrior aristoc- racy, they were conducted by Brahman ritualists with the expectation that the

described as keśinā (‘long-haired’), the suggestion that they were a seasonal phenomena is improbable; long hair cannot have been acquired in a single season. 65 Heesterman’s work in this area remains controversial, but developments in the chronol- ogy and understanding of the Indo-European influx into the sub-continent as involving a long period of settlement (c.f. Parpola [esp., 1992]; Witzel) support his findings and sug- gest that a pre-classical period might be located in developments in the now Swat valley circa 1900–1700bce. 66 Heesterman (1988: 254–258). 67 Thapar (2000: 922) notes that (in a later period), while forest hermitages were “often the vanguard of the colonisation of the area by settlers or agriculturalists … hermitages were often under attack by those who claimed the forest as their territory or hunting ground.” mountains and renunciates 41 gods would provide specific requested benefits in return. But in time, in a fun- damental intellectual development that shaped the direction of Indic thought over the ensuing millenniums, this expectation gradually became a belief that a sacrificial ritual correct in every detail compelled the gods to act as requested.68 The gods’ actions thus came to be seen as being, ultimately, controlled by those who performed the rituals; a conceptual development radically different from that which evolved in the monotheist traditions of the Middle East; Judaism, Christianity, and Islam. Brahmans were considered (or claimed), to be not just intermediaries with or interpreters of divine will, but as able to control the deities. By around 900bce., as sacrifices became increasingly complex, three addi- tional liturgical texts emerged. The Samaveda and Yajurveda developed sac- rificial ritual and do not concern us, but the Atharvaveda is significant. The teachings of each Brahmanical school developed independently as ‘branches’ (śakhā) of Vedic learning,69 and the Atharvaveda was of somewhat different orientation to the teachings preserved in the other Vedas.70 It included six (and in later related texts, ten71) classes of defensive and offensive magical rites and, perhaps most significantly, rites for the redressing of ritual defilements and pol- lution (such as from funeral practices). The concept of purity and pollution was central to the Vedic social system. ‘Impure’ elements such as dead bodies or unsanctified territory were a source of pollution to the ‘pure’ Brahmanical caste and ‘ordinary’ Brahmans thus avoided contact with such polluting elements. The Atharvavedic specialists, however, were Brahmans equipped and able to deal with these polluting factors,72 char- acteristics that align their role with the vratya, as well as the Magi and wu in Persian and Chinese cultures.

68 See for example, Mahabharata 3.305.20 (cited in Bloomfield: 1978; 27), which states that the sage Kunti knows the Atharvavedic mantras that compel the gods to appear. Heester- man (1993: 161) points to the identification of the sacrificer with the Brahman officiant as the point in the process of sacrificial development at which the “final resolution is taken as a foregone conclusion.” Also see Heesterman (1997: 251–252, 66) who concludes that ritual became so crucial that, “The institution of sacrifice did not depend on the gods. It can do without them.” It was, Heesterman (1988: 260) states, “the unflinching rules of the ritual that promise absolute certainty.” 69 Gonda (1975: 29–31); also see Carpenter (1992: 63); also see for Vedic schools, Renou (1957: 50–53). 70 On the Brahmanical nature of the text, see Gonda (1975: n. 2 277, 279, 309). 71 e.g.: the Āngirasakalpa of the Atharvaveda; Gonda (1975: 277). 72 Gonda (1975: 272, 283, 286, 305). 42 chapter 1

The Atharvavedic Brahman’s mastery of magical rites and ability to transact with elements considered impure made them ideally equipped for the role of royal priest. As Vedic culture developed, an ideal model of authority emerged in which Ksatriya kings maintained a Brahman priest (purohita) at their right- hand side to intercede on their behalf with the deities, and to provide ideo- logical support for the ruler through textual sanction and the performance of rituals legitimating and enhancing royal power.73 Although the Atharvaveda contains practices appealing to wider society, its Brahmans’ primary histor- ical significance derives from their claim—generally successful—to occupy that post.74 But the impurity of landscape that had not been drawn into the sacralised territory of the Indo-Europeans also meant that the Atharvavedic Brahmans were qualified to deal with that territory. Thus those Brahmans who travelled to new realms or sanctified territory must—in some transformation of the vratyas’ role—have derived from, or had links to, the Atharvavedic schools. The Brahmanical caste did not develop a single source of authority or cen- tral organisational structure. Their power remained decentralised and the var- ious schools and individuals sought patronage among all sections of society. There were, therefore, numerous different groups of Brahmans, many doubt- less unknown to us now, who specialised in particular practices and teachings. In the Atharvaveda, for example, seven hymns are ascribed to the sage Garut- man, all of which concern counteracting poison.75 It requires no great leap of the imagination to see these hymns as a record of the wisdom of an individual renowned as a specialist in that field. While texts from the late-Vedic era (c. 900–500bce.) develop earlier tenden- cies, they also indicate new concerns. The Brāhmaṇas explored cosmological aspects of the sacrifices, indicative perhaps of a search for causal unity underly- ing apparently diverse phenomena. This tendency is also indicated by the more mystically orientated Āranyakas (designed for those practicing in the forests), and the deeply philosophical speculation of the early Upanisads. In these texts Vedic polytheism was challenged by new understandings, not least of a single underlying causal reality which was, put at its simplest, ultimately the union of the individual and universal soul (ātman and brahman). The Upanisads indicate that novel ideas of an after-life and wider explana- tions of existence were prominent in Indic thought at that time. The old gods of

73 As was to be characteristic of such models, this was portrayed as replicating a divine order; see Heesterman (1962: 22); also see Kosambi (1946: 43). 74 Bloomfield (1978: lviii, lxi–lxx); Gonda (1975: 285); for a discussion of this strategy, see Sanderson (2007: 203–208). 75 Whitney (1971: vol. 1, 1039); but see Karambelkar (1959: 248). mountains and renunciates 43 the Rgveda were fading from the forefront of belief and becoming overlaid by others. Gradually, the major unifying ideas of Indic civilisation were emerging amidst a welter of intellectual speculation, the idea, for example, that the power of the sacrifice could be internalised through yogic practices.76 The Upanisads contain frequent references to means of realising ultimate reality through yoga, which now appeared as a wide-spread religious practice.77 Many of these ideas emerged from the renunciate movement, for a key point in Indic religious history is that Brahmans did not have the authority, the means, or indeed any particular reason to exclude new religious developments, particularly those which were an adjunct to, or arose from, the sacrificial rites. With the exception of the marginalisation of sexualised rites there is little or no evidence of the suppression of any belief systems or practices. If ideas stood the test of time they were accepted within the wider belief system. Given that they were free to devote themselves to religious contemplation, numerous aspects of Brahmanical and later Hindu beliefs must have arisen through the speculation of renunciates. Indeed, Louis Dumont has famously proposed that the “true historical development of Hinduism is in the sanyasic developments on the one hand and their aggregation to worldly religion on the other.”78 The renouncers’ interaction with Brahman sacerdotalists was thus both complementary and competitive, with many diverse groups of renouncers and their ideas and practices developing and producing new Brahmanical rituals, ideas, and texts which were eventually reintegrated into the wider system. The day-to-day relationship between the two groups must always have been a complex one, organic, overlapping, and localised in character. In texts from this era there are also glimpses of developments in regard to Mount Kailas. While we have no Tibetan sources for the period, Atharvaveda 6.24.1 states that the medicinal waters which irrigated the tribal lands, “flow forth from the snowy (mountain); in the Indus somewhere (is their gathering)”; suggesting an awareness that somewhere in the mountains which sheltered the Indus source there were other river sources. Yet there was already a sense that these regions were inaccessible to ordinary men: the Aitareya-Brāhmaṇa (8.23) referred to the northern realm of the Kuru (Uttarākuru) as, “the land of God, which no man can conquer.”79

76 “The ultimate sacrificer became the renouncer who drew the whole world into himself.”; Heesterman (1991: 305). 77 See, for example, Kaṭha Upaniṣad 6.10–11, & 11.6–18; also see Śvetāśvatara Upaniṣad 11.8– 140. 78 Dumont (1970: n. 21 46); also see Bronkhorst (1998). 79 Tucci (1971: 508). 44 chapter 1

Then in the Taittiriya Āranyaka (1.7.20) we find the earliest reference to Mount Mahāmeru (‘Great Meru’). It does not seem a significant mountain here, just one referred to in passing as a place where the rsi Kaśyapa wishes to remain.80 It is not geographically located in any way or associated with Mount Kailas, which is not mentioned in these earlier texts. Yet Mount Meru was to become the mountain at the centre of world in later Indic cosmology, and sometimes, at least in the modern period, even to be identified with Kailas. That identification evolved from a concern to replicate the cosmic order on earth, which will be seen to be a fundamental tendency in the ensuing period, one which would bring Kailas to the forefront of Indic thought.

80 Kaśyapa may also be linked to the Himalayas as the author of the 9th book of Rgveda, which contains the soma hymns. chapter 2 “The Play Garden of the Gods … Beyond the Course of Humans”1

Introduction

While the Vedic peoples did not know a distinct Kailas, they had ritualised mechanisms for territorial expansion and their cultural ‘Heroes’, the rsis, were renunciates reputed to have travelled into the Himalayas. Those mountains had funereal and even sacred associations in Vedic thought, as did the rivers that had their sources there. In the centuries leading to the Christian era those understandings were to undergo considerable development and to be linked to many other aspects of Indic culture. New belief systems emerged in north India with followers of the historical Buddha (c. 480–400bce2) and the Jain prophet Mahavira (c. 540–468bce) self- consciously identifying themselves as belonging to autonomous traditions.3 Both founders were renunciates who emerged from the Brahmanical world and accepted most of the existing culture, albeit those beliefs later devel- oped on somewhat different lines. The Brahmanical system itself was trans- formed by various imperatives (including the rise of the new traditions), and by around 200bce., new formulations of worship and of the Absolute emerged. The sacrificial rites largely gave way to temple and sacred site-based worship of newly prominent deities, most notably in our context Śiva and Visnu (Viṣṇu).4 This new phase, commonly termed ‘Classical Hinduism’,5 was most famously expressed in the Epic sagas, the Mahābharata and the Rāmāyaṇa, the texts where a sacred Kailas emerges.

1 Mahabharata (hereafter MhB.), 3.(33).156.20–25; (the edition cited here throughout is J.A.B. van Buitenen [ed./trans.], 3 vols. Chicago, University Press, 1973–1978; [hereafter, van Buite- nen]); van Buitenen (vol. 2; 524). 2 Gombrich (1992); also see Cousins (1996). 3 Not all such claims led to successful traditions; see for example, Basham (1951). 4 The earliest mention of ‘Śiva’ is as one of 30 Rgvedic clans/tribes; Witzel (1995b: 313), suggest- ing his association with a particular Vedic lineage. 5 Divisions such as ‘Vedic’, ‘Brahmanical’, and ‘Classical’ Hinduism are of course misleading in many senses, and should be understood solely as heuristic devices.

© koninklijke brill nv, leiden, 2015 | doi: 10.1163/9789004306189_004 46 chapter 2

Many new sects developed from Vedic traditions held by regionally-based gotras.6 The continued existence of practices associated with the Atharvavedic schools, for example, is testified to in early Buddhist sources,7 as well as those of classical Indian medicine (āyurveda).8 While renunciate forms multiplied most renouncers, then as now, were from the Brahman caste,9 and their teachings and practices generally remained complementary to the activities of temple Brahmans. The close links between the two types merged in the ideal of the rsi, those great sages dwelling in the Himalayas, Brahman renouncers who were “devoted scholars of the Vedas.”10 But the development of temples dedicated to specific central deities gave rise to a new distinction; renouncers did not require images of deities or temples for their religious practices. Deities were visualised by the individual renouncer and the division between pure and impure space was overcome by the renouncer’s mastery of impure forces. Indeed certain renunciate traditions deliberately sought out impure locations such as cemeteries in order to con- front, overcome, and manipulate the powerful forces of impurity to their own ends. The ideal of motion inherent in renunciation also dictated independence from temple practice, although the observable modern tendency for renunci- ates to eventually settle at a sacred site probably always existed (not least for practical reasons of old age and infirmity preventing itinerancy). The significance of these developments will become more apparent in this chapter as we trace the origin and development of pilgrimage as an Indic phenomena and discuss possible sources for the earliest knowledge of Kailas in Indic society. While the assumption that the Mahabharata and Ramayana refer to the complex we know today as Kailas-Manasarovar underpins much of the traditional status of the sacred site, analysis of these texts indicates that not only was Kailas not a popular pilgrimage place, it was not an earthly mountain either.

6 “Vedic texts have a history of transmission and belong to various schools of tradition. The textual divisions … [are] among the marks that distinguish individual schools from one another.”; Griffiths (2003: 5). 7 The Buddha called them Atharvanikas, “because of their preoccupation with sorcery”; Gupta (2002: 237). 8 On which see Zysk (1991). 9 Renunciates were given textually sanctioned status through the varṇāśramadharma sys- tem (ie. ‘caste-lifestage-appropriate behaviour’). On this textual concept, see in particular the works of Patrick Olivelle; e.g. Olivelle (1974). 10 MhB. 3.(33).145.30–35; van Buitenen (vol. 2: 497). Interestingly, these sages practice tapas yet also perform sacrifices, echoing the earlier vratya model. “beyond the course of humans” 47

Pilgrimage

There is no evidence of pilgrimage in the Vedic period as a distinct or sys- temised phenomena. Its origins seem connected to sites where there had been sacrifices, including funeral rites. Those sites, with their ritually demarcated sacred space, were usually close to rivers or water sources,11 and the commonest Sanskrit term for pilgrimage is tīrtha-yatra, now used in the sense of a “spiritual journey to a sacred place”, but literally meaning a “journey to a river crossing, or ford”.12 The classic pilgrimage rite of bathing is referred to in the Yajurveda, where a priest of the soma sacrifice is described as bathing at a tirtha,13 and sites known to have been the location of a successful sacrifice—understood as a place where the gods had been present—likely retained spiritual status and subsequently became gathering places for religious seekers The first textual reference to Indic pilgrimage may be in the Aitareya Brāh- mana, a late Vedic text which states that;

Evil is he who stayeth among men, Indra is the comrade of the wanderer, Do thou wander.

The text goes on to praise

The flower-like heels of the wanderer, His body groweth and is fruitful, All his sins disappear, Slain by the toil of his wandering.14

There is, however, no mention of specific sacred destinations associated with itinerancy in this reference, which may refer to a renunciate life-style rather than pilgrimage. But the general concept of specific benefits for particular ritual acts inherent in the developed understandings of pilgrimage can be found in the earlier Atharvaveda. This listed benefits accruing to those who offered shelter to vratyas, whose practices, as we have seen, included ritual travel. The

11 See, for example, MhB. 3.(33).88.1–10; van Buitenen (vol. 2: 402). Also see, Elstink (1979). 12 Tīrtha, from the Sanskrit root tr/tarati (“to cross over”), appears in the Rgveda predomi- nantly in the sense of a “ford” or even a “road”,but may imply “transition”,“transcendence”, or a “liminal place”. On which see Eck (1985: 325); Singh (1995: 9); Kane (1953: 554). 13 See Eck (1985: esp. 327–328). 14 Bhardwaj (1973: 3–4; quoting Aitreya Brāhmaṇa, vii.15). 48 chapter 2 benefits ranged from receiving “possession of the pure worlds that are on this earth” for those who offered one night’s hospitality, to receiving “possession of the pure worlds that are unlimited” for those whose hospitality was likewise unlimited.15 With the Epics however, and in contemporary Buddhist16 and Jain literature, a phenomena not previously clearly articulated emerges as fully-formed and integral to that world. Pilgrimage appears as a religious undertaking that was systemised, culturally accepted, and widely practiced by various social groups. The advantages of pilgrimage are clearly stated in the Mahabharata:

A poor man cannot rise to the sacrifices, for they require many imple- ments and a great variety of ingredients … But hear to what injunction even the poor can rise, equalling the holy rewards of the sacrifice. This is the highest mystery of the seers [rsis]—the holy visitation of sacred fords, which even surpasses the sacrifices. Poor indeed becomes only he who has never fasted for three nights and days, has never visited the fords, and has not given away gold and cows. One may sacrifice with the Laud- of-the-Fire and other rites of rich stipends, yet not reap the reward that visiting the fords brings.17

Pilgrimage is thus (and, in the textual sense, suddenly), proclaimed as a reli- gious practice of the highest value, feasible for all classes of people. We do not know why pilgrimage became specifically defined and endorsed in this man- ner, no spontaneous revelation or act of devotion is suggested by any source. Rather a number of factors suggest themselves, not least the transparent eco- nomic interests of the Brahmanical class who authored the Epics, for pilgrims are consistently urged to gift “gold and cows” to Brahmans resident at pilgrim- age sites. The association of pilgrimage sites with locations linked to specific territorial formations does support Makhan Jha’s conclusion that pilgrimages originated “with the development of ideological legitimation for empires.”18 In that they

15 Atharvaveda xv.13.1–5. 16 Huber (2008: 21) locates “the first more systematized literary accounts of an extended pilgrimage network which constitutes something akin to a larger Buddhist ritual circuit and religious territory in India”, in versions of the life of Asoka, the Āśokāvadāna or Divyāvadāna, which dates from “about the second century ce and contains material that may be somewhat older.” 17 MhB. 3.(33).80.35–45; van Buitenen (vol. 2: 373–374). 18 Jha (1985: 7). “beyond the course of humans” 49 articulated a particular concept of certain specified territory as purified and sanctified, pilgrimage sites and the routes between them tended to act as cul- tural, and ultimately political, unifiers.19 Sites locally considered sacred might thus be drawn into regional prominence, with those regional sites taking on wider significance as the rise of empires created multi-regional conglomera- tions. As the tribal territories in the Gangetic plains matured into political units, Magadha (centred around present-day Patna), emerged as the dominant power in northern India,20 reaching its height under the great emperor Aśoka (reigned c. 268–239bce.21). The Mauryan empire collapsed within 50 years of his pass- ing, fragmenting into smaller, generally unstable kingdoms which were not reunited until the rise of the Gupta dynasty in the 4th century ce. Thus for some six centuries, north India comprised a series of regional centres and its history was shaped by regional interests, although certain shared cultural inheritances linked these regions, or at least their elite classes. But the Mahab- harata describes an interconnected network of pilgrimage sites covering vast expanses of north India (a description for which there is no textual precedent). This network echoes the Mauryas’ territorial expansion, with the later rise of regional traditions expressed in the Puranic literature, the regional diversity of which suggests the subsequent fragmentation of that territory. Mauryan encouragement of pilgrimage is indicated by the erection of mon- uments with religio-ethical associations (such as the inscribed ‘dharma pillars’ and Buddhist stūpas which multiplied in this period), and Aśoka’s sponsoring of travellers’ resthouses (dharamsalas).22 Superimposing a sacred geography onto their territory could stimulate a sense of unity and encourage a certain degree of cultural incorporation of minority groups, while the development of a wider geographical consciousness would also have stimulated the movement of capital and trade goods throughout their kingdom. The Mauryans apparently recognised that pilgrimage, like war, is good business. But if that model seems sustainable, it does not support the existence of a sacred Kailas at that time; the upper Himalayan region was far from any Indic empire, and was not yet a site for pilgrimage.

19 In much of the Indian Himalayas, as Channa (2013: 119) notes, it is the local devtās (deities) who are considered to undertake pilgrimages, with humans seen only as accompanying the deities. 20 For a textual study of the intellectual influence of Magadha that proposes a common source there for the Brahmanical and Śramaṇic traditions, see Bronkhorst (2007). 21 Aśoka’s precise dates are controversial, on which see Eggermont (1965–1966). 22 Tambiah (1976: 17). 50 chapter 2

We might note here that historically the value of pilgrimage has not been uncontested in Indic, or indeed most other religious traditions. Not only did its critics claim that any benefits could be more easily obtained by non-itinerant ritual practice, but the later internalisation of sacred geography became a recognised Indo-Tibetan practice (indeed in Tibetan literature it is often dif- ficult to tell if a described journey was real or visionary). Such doubts were wide-spread, and one of the implications of the proverb, “it is wrong to … follow up a river to its very source.”23

The Discovery of Kailas

While the identification of the Epic Kailas with the modern cannot, as will be seen, sustained, the Epics do reveal a greatly amplified knowledge of the Himalayas compared to earlier texts. Sacred sites were not situated in any fixed hierarchy, but particular features were becoming known although given different toponyms by different groups. The level of knowledge suggested is, if we ignore the possibility of coded knowledge, that of a few individuals who had travelled there, or of information obtained at considerable remove.24 There were at least three possible sources for this knowledge: traders, migrat- ing tribes, and renunciates. The lure of gold, not gods, may well have first drawn attention to the Kailas region. Gold is found around Lake Manasarovar and has been mined at the nearby Thok Jalung goldfields since the earliest recorded times. Herodotus, writing more than a century before Alexander the Great’s campaign brought India more fully into Greek knowledge, famously recorded a legend about “gold-digging ants” which may relate to those gold-fields.25 One explanation for this legend and for the paucity of precise geographical refer- ences to Kailas in early sources is that economic and security concerns would have given those who found gold in the mountains good cause to conceal or encrypt their knowledge. But we cannot be certain this reference is actually to gold from the Kailas region. Gold was passed by middlemen on its passage from the mountain mines to the courts and markets of the plains and traces of gold are also found in many Himalayan rivers.

23 Cited in Selvanayagam (1992: 59). For a similar Jain injunction, see Flugel (2010: n. 92 428). 24 “Theories of cognitive mapping suggest that we acquire [knowledge of] first landmarks, then the routes between them, and finally the area surrounding them.”; Leyerle (1996: 120). 25 Herodotus iii: 102; (Beloe: vol. 2; 88–91). Pliny’s account of the gold-digging ants is noted there. Michael Peissel has proposed identifying the “ants” as marmots, whose diggings throw up sand which may contain gold, see Osmaston (1996: 8). “beyond the course of humans” 51

From other Greek writers such as Ptolemy (2nd century ce.), we find the gold-digging ants associated with an Indic mountain tribe, the Darada, or ‘Dards’. This tribe was known to Herodotus as tributary to the Achaemenian (Persian) empire and was later recorded as fleeing deep into the mountains in the face of Alexander’s forces. In the Mahabharata, we read of pipīlika (‘dug by ants’) gold being presented as tribute by Himalayan tribes, including the Dards.26 Linguistic evidence indicates that the Dards were Indo-Europeans, who probably migrated through Swat during the second millennium bce.,27 (and are therefore contemporary with, or part of, the early tribes whose sagas are recorded in the Rgveda). Their migrations through the Pamirs, Hindu Kush and western Himalayas into areas of Kashmir and Ladakh suggest they could have been a source of Indic knowledge of mountain formations. It is uncertain whether the Dards ever formed a unified kingdom or migrated in significant numbers as far east as the Kailas region.28 The geography of the western Himalayas, with its rugged terrain and isolated habitable valleys, sug- gests that they probably fragmented into numerous smaller groups with shift- ing patterns of alliance and allegiance. Evidence for their subsequent cultural isolation is that they are among those mountain tribes described in Hindu texts as having lapsed from the Ksatriya caste, due to their neglecting the Brahman- ical rites.29 Their caste status thus declined along with other mountain tribes such as Kinnārs, Kiratas, and Cīnas, whose origins and identity are now simi- larly difficult to ascertain.30 The western Himalayas were thus home to numerous small tribal groups, including those sharing cultural links with the Indo-European tribes to the south and descendents of those who had settled there in Neolithic times. The mountain regions were isolated from many trends and developments in the north Indian plains, allowing localised cultures to evolve and autonomous political organisations to grow in relative seclusion. But despite the difficult terrain, long-distance migration and trade through these regions was possible along rivers such as the Indus and Sutlej and it does appear possible that some

26 MhB. 2.48.5–10; van Buitenen (vol. 2: 118); also see Joshi & Brown (1987: 309); Vohra (1983: 54–55); Tucci (1977: 18–20). 27 Petech (1977: 5–6). 28 But see Walter & Beckwith (1997: 1037). 29 Tucci (1977: 37). 30 Tucci (1977: esp., 36–37): Vohra (1983: esp., 60–61). The tendency to associate groups such as the Kinners with modern territorial units of similar name, i.e. Kinnaur, is extremely common, but ahistorical without supporting evidence, which is almost invariably lack- ing. 52 chapter 2

Indo-European groups reached the Kailas region and that contact with them allowed a knowledge of Kailas to permeate into India. The third possible contender for the title of ‘Discover of the Kailas region’ is the renunciate movement. Esoteric groups such as the vratyas were involved in territorial expansion from the earliest Vedic period and in the Epics renun- ciates are recorded as journeying into the mountain regions in search of herbs and to the sources of rivers which they held sacred. Their journeys clearly went beyond established routes through known territories; there are prosaic refer- ences in the Mahabharata, for example, to a sage travelling into the mountains “by the path of the Siddhas”, and to those journeying there as “living on roots and fruit.”31 Many aspects of the esoteric knowledge of the renunciates can be viewed as intended solutions to very practical problems such travellers would face. Atharvavedic mantras against serpents and wild beasts, for example,32 or yogic practices to raise the body temperature, would complement existing practices of inner Himalayan travel. These included trade in small items of high value and the practice expressed in the local saying; “load the sheep with grain, finish it first, then kill and eat the sheep; that is the way we travel on the uninhabitable tracts.”33 While difficult and dangerous, a trans-Himalayan journey would not have been impossible for renunciates. Tribal migrations and trade took place through these regions, and renunciates would have followed those routes. Given that the Himalayas are rising at an average of two feet (.6 mtrs) a year, the passes would have been several thousand feet (c. 600 mtrs) lower than they are today, and thus easier to traverse.34 Certainly there is one consistent aspect of Epic references to the upper Himalayas; their statements that beyond the ter- ritories of the hill-tribes lay a sacred Himalayan heartland where renunciates carried out religious practices among the gods who dwelled there. Individual identities, routes, specific practices, and actual locations may be uncertain or

31 MhB. 3.(33).143.1–5; van Buitenen (vol. 2: 494); MhB. 3.(33).145.5–10; van Buitenen (vol. 2: 496). 32 See, for example, the Atharvavedic mantras (largely taken from Rgvedic verses), in Bhatt (1987: 38–39, 72, 76); for driving away wolves; against tigers, diseases from food, “obtaining herbs from an enemy’s herbal field”, and “to secure release from a wild forest”, etc. 33 Gerard (1993: 150). 34 The Swiss geologist, Augusto Gansser, who visited the region in 1936, concluded that there was “recent and intensive convolution” of the rock formations around Kailas and that the Indo-Tibetan passes had risen by 3,000 feet or more since the coming of man to the area; Heim & Gansser (1939: 96, 104, 218). For similar conclusions, see Wadia (1968: 36); and Mithal (1968: 45, 50, 76). “beyond the course of humans” 53 unstated, but we have considerable evidence of the cultural milieu from which these renunciates came, the ideals they espoused, and the manner of their asceticism. We can assume close links between that evidence and the actuality of a renunciate presence in the high Himalayas. But whether that presence reached as far as Kailas-Manasarovar is another question. What we can conclude is that in the mountains the renunciates’ companions and informants must have included those from the fringes of society; outlaws, runaway slaves, and travelling performers, as well as local tribes. Ordinary Brahmans, bound by purity concerns, did not venture there, nor could Ksatriya armies do more than raid into the foothills. Merchants from the plains dealt with intermediaries among the hill-tribes, while farmers and low-caste workers had no business there. Neither had pilgrims. As will be seen, the Epics, which demonstrate that pilgrimage had become a widespread phenomena by this time, clearly and explicitly stated that the Himalayan region was not a place for the ordinary pilgrim. The pure land of the gods lay beyond a zone of impure wilderness, and those who wished to travel through that impure zone had to be equipped to deal with that impurity. Thus those who travelled into the upper Himalayas must have included those who followed the traditions of the Atharvavedins and the vratyas, the munis and the rsis. Their numbers included Jains and Buddhists, particularly when those were primarily renunciate movements, but they became most closely associated with some form of Śiva, whose character echoed that of Rudra, the fierce mountain god of the Vedas.

Kailas-Manasarovar in the Epics

There have been critical studies problematising the extent to which the literary Kailas can be understood to represent a real place, as will be seen. But most attempts to identify places mentioned in the Epics with modern toponyms have proceeded on the assumption that the literary Kailas is actually manifest in an earthly location and that geographical reality was thus concealed within the texts. The identification of the literary Kailas with the site in western Tibet has meant that toponyms and geographical references have commonly been situated in relation to that one ‘established’ fixed point. But no real consensus has emerged from these studies, which have generally foundered on the internal inconsistencies in individual texts and the overall incompatibility of the diverse accounts of the region in the classical works of Hinduism, as well as from a lack of chronological analysis and context in regard to literary traditions dating back around two millennium. 54 chapter 2

When we locate the Kailas of the Epics in its wider historical and social con- text we can, however, determine certain conceptual and actual developments in regard to the understanding of the location and nature of an earthly Kailas- Manasarovar. There are the usual caveats that apply to the historical study of Indic texts; as with the Vedic literature, dating the Epics is highly problematic in that they contain earlier material originally preserved in oral form as well as later interpolations. Their compilation was thus a process rather than an event which can be precisely dated. But for general purposes we may assign a date to the composition of the earliest sections of the Mahabharata of around 400–200bce., with its final expression taking form around 200–400ce.35 The Ramayana was contemporary, but slightly earlier.36 The great age of the Brahmanical sacrificial system had passed by the time the Epics were recorded, and many elements we now characterise as Hindu fully emerge only in these texts, which were intended to “fix the normative values of Aryan culture.”37 One aspect of that project was to delimitate the sacred geography of the Indic realm, and these sagas do imply an expanded knowledge or perception—whether empirical or mythical—of the Himalayas since the codification of the four Vedas. Testifying to an expansion of the Vedic cultural complex, large sections of the subcontinent that lay outside Vedic heartlands emerge within the Epics as sacralised territory, and the institution of a pilgrimage network embracing that territory is indicated. This is made clear in the section of the Mahabharata (3.80–88) which describes north Indian sacred places which pilgrims should visit. “The Tour of the Sacred Fords” (i.e.; the sacred tirthas, or places of pil- grimage),38 begins with Pushkar (in what is now Rajastan), described—in what became a standard trope—as a place where along with gods and various other deities dwelled Brahman “seers” who had achieved divinity through ascetic practices. The text states that pilgrims who bathed in the lake at Pushkar, “are not reborn in lower forms” and that whatever evil they had done was destroyed, whether they were “great-spirited Brahmans, barons, commoners, or serfs.” While the pilgrim is warned Pushkar was “hard to reach … gifts … are hard

35 Hiltebeitel (2004: 203–227), argues for a much shorter chronology. But Brockington (1998: 135, 144–145), considers the relevant parts of the Epics to be late. 36 Brockington (2003). 37 Varma (1988–1989: 13). 38 The “Tour of the Sacred Fords” (tīrtha-yātrā parva) section of the Epic is a sub-section of the “Forest Section” or “Book of the Forest” (araṇya-parvan): I am indebted to Dominic Wujastyk for this clarification. “beyond the course of humans” 55

[expensive?], to live there is very hard”,39 it is clearly stated that Pushkar was a pilgrimage site open to all.40 The “Tour of the Sacred Fords” goes on to describe numerous other sacred places—towns, rivers, and mountains—to which pilgrims should travel. It specifies the offerings they should make at each site, the rituals to under- take there (such as the previously unspecified rite of circumambulation; Skrt: parikrama), and the benefits accruing to them from those actions. Particular religio-historical associations, attributes, and specific benefits were ascribed to each site. Thus, to select an example at random, pilgrims were advised to visit Dirghasattra;

where the Gods led by Brahma, the Siddhas [those of magic power], and eminent seers attended a long Session [eg; sacrifice] with stipends; [there the pilgrim may obtain] the reward of a Royal Consecration and a Horse Sacrifice.41

The sacred area delineated in the “Tour” seems to embrace the Upanisadic- era heartland in the Indian plains, bounded in the north by the Himalayas but reaching up the river courses into the mountains. The ‘sacred fords’ of the northern region to which the pilgrim is recommended thus centre on sites along the Sarasvati, Yamuna, and Ganges rivers. These sites are frequently iden- tified as those where Vedic sacrifices were undertaken in the past, confirming at least part of the conceptual and/or actual process by which places were incor- porated into this sacred schema. Although the description is confused, one of the places mentioned is Badarī, commonly identified with the modern Badrinath, close to the source of the Alakananda branch of the Ganges and now a major Vaisnavite sacred site. The god Narayana (associated with Viṣṇu) is hailed as supreme there; “He is the holy, he is the supreme Brahman, he is the ford, he is the forest of austerities. And with him are the Gods, seers, Siddhas, and all the ascetics.”42 If this Badari is the modern site it indicates a knowledge of the upper reaches of the Ganges

39 MhB. 3.(33).80.40–60; van Buitenen (vol. 2: 374). 40 In Pushkar I was informed by a resident that until the financial benefits of tourism were made clear by the late 1960s influx of foreign visitors, residency in Pushkar was open only to Brahmans, and one temple there remains restricted to Hindus. 41 MhB. 3.(33).80.115–120; van Buitenen (vol. 2: 377). 42 MhB. 3.(33).88.20–30; van Buitenen (vol. 2: 403). The northern ‘Ford’ sites are described in MhB.3.(33). 88.1–30; van Buitenen (vol. 2: 402–403). Badari is here, however, a river synonymous with the Ganges. 56 chapter 2 and a Vaisnavite claim to that site. But Mount Kailas (and Manasarovar) are not mentioned in the “Tour” section of the Mahabharata. Thus—and the point is of major importance to our enquiry—there is no indication in the text that Kailas was a pilgrimage site. Kailas does appear in the Mahabharata in another context. The central nar- rative of the Epic concerns the power struggle between the five royal Pandava brothers and their cousins, the Kauravas, who usurp the Pandavas’ kingdom. The principal hero is Arjuna, one of the Pandava brothers, who is both a great warrior-hero carrying his bow, and a skin-clad master of ascetic practices.43 (This contradiction, which seems to suggest the earlier vratya model of the “consecrated warrior”, is remarked upon within the text.) In the course of the narrative, Arjuna travels north into the Himalayas, and there he hears of Kailas, a mountain ruled over by the deity Kubera and a place where “the Gods assem- ble.”44 Arriving at the “great and wondrous” Kailas mountain, the Pandavas find there Narayana’s hermitage, described as a;

blessed place … free from afflictions … crowded by hosts of great seers and filled with Vedic luster … a divine place of refuge for all creatures … [where] Lordly scholars of the brahman lived … great seers and ascetics who had regenerated their spirits and mastered their senses in quest of release, who subsisted on fruit and roots, self-restrained … the equals of the sun and the fire by the power of their austerities … devoted scholars of the Veda.45

The text actually says little else specifically about Kailas, although we read of an unnamed lake;

on the beautiful wooded crest of Kailasa [is a] lovely pond … which sprang from mountain falls close to Kubera’s dwelling, very lovely, of ample shade, and surrounded with all kinds of trees, covered with yellow lilies and divinely floating with golden lotuses—enough to purify the world, holy and of wonderful aspect … water, tasting like Elixir … the celestial pond … visited by seers …46

43 On Arjuna in the Mahabharata, see in particular, Katz (1989). 44 MhB. 3.(33).140.10–15; van Buitenen (vol. 2: 490). 45 MhB. 3.(33).145.15–30; van Buitenen (vol. 2: 497). 46 MhB. 3.(33).151.1–10; van Buitenen (vol. 2: 510). “beyond the course of humans” 57

But while commonly read as a reference to Manasarovar, that conclusion is not supported by either geographical realities or analysis of the text. In an important article (in German47), the Sanskritist Reinhold Grunendahl prob- lematises the identification of the Kailas in the Mahabharata with Kailas- Manasarovar. He demonstrates that the names of two mountains in the region, Kailas and Gandhamādana,48 are used interchangeably in the text. In addi- tion, not only do the two accounts of Arjuna’s travels seem to describe two separate mountains as Kailas, but the geographic descriptions of the region do not match the geography of the present-day Tibetan Kailas.49 Grunendahl con- cludes that the Kailas described in the Mahabharata is actually a mountain located around the Badrinath region of the Garwhal Himalaya, close to that source of the Ganges. This is a conclusion we shall return to.50 Grunendahl has also demonstrated that the descriptions of Arjuna’s trav- els in the Kailas region occur in two versions in the text (MhB. 3.140–153; & 3.155–162). These two versions date to different periods, with the second account actually being the earlier. In that earlier account Arjuna’s journey is less complicated than in the later version, which highlights the inaccessibility of the region. The influence of the Nārāyaṇa school of Vaisnavism is apparent from the articulation of their views in various parts of the Epic, and Grunen- dahl identifies the emphasis on the inaccessibility of the region in the later account with Narayana attempts to claim the Badari-Kailas region as their sacred centre.51 This seems consistent with Indic textual strategies from the earliest period, and the known emergence of the Narayana sect in the early Christian era further indicates these Kailas sections of the Epic are from its final redaction.

47 See Grunendahl (1993): I do not cite page numbers but rely on notes made when Dr. Grunendahl kindly explained the main features of this German language article to me personally, and in later email correspondence, 26 August 1996. 48 Staal (1990: 282) notes that this may be the mountain known today as the Gurlha Mand- hata but that the names “are not related.” Situated on the Indo-Tibetan border, Gandhamā- dana (Tib: sPos-ngad-ldan) is sometimes distinguished from the modern Gurla Mandhata (Tib: sMan-nag-snyil), see for example, Huber & Rigzin (1999; 125–153). Textually promi- nent in one form or another, the mountain sometimes has spiritual associations and sometimes not. 49 See for example, MhB. 3.(35).155.10–15; van Buitenen (vol. 2: 519), where Mount Kailas can be seen before reaching the foothills of Gandhamādana and Meru, until as Arjuna travels “ever higher … he came to the plateau of the Himalayas …” 50 Suryavanshi (1986: 59) reaches a similar conclusion. 51 A study of Laxmi-Narayana origins is a lacuna. 58 chapter 2

Turning to the Ramayana, we again find that the hero, the god Rama is a great royal warrior, depicted as the ideal Ksatriya, strong, noble, truthful, learned, and just. But like Arjuna, he wears the matted hair and clothing of a renunciate and carries his weapons with him. The Kailas-Manasarovar region appears briefly in this saga in the context of the search for Rama’s wife Sita, abducted by the demon Ravana. The search encompasses the Himalayas, and there are descriptions of the route her pursuers must follow, including a path to the north. This description is reasonably consistent with the landscape on the approach to the actual Kailas mountain of the Tibetan plateau after crossing the Himalayan passes to the south. Rama is told that after crossing the mountains he will come to;

an open space, a hundred leagues on every side, without mountains, rivers, or trees, devoid of any living thing … you will be delighted once you reach the white mountain Kailāsa. There, bright as a white cloud and embellished with gold, stands Kubera’s heavenly dwelling …52

But the route followed by Rama’s party takes them past various mountains, each in turn described as having fantastic characteristics. There is Mount Meru, “the king and greatest of mountains”,53 Mount Sudarśana, “a king of mountains filled with gold”, as well as, “the great golden mountain … Somagiri … king of mountains, for it reaches up to heaven … [where] dwells Brahma”54 and confusingly, there is a “wish-fulfilling mountain Mānasa.”55 A “vast lotus pond” is mentioned briefly, but Manasarovar is mentioned by name just once in the Ramayana. In a curiously colourless myth, Rama, enquiring as to the source of the river Sarayū, is told that it originates “on Mount Kailasa [where] there is a lake that Brahma produced from his mind, manas, because of this … it is called Lake Mānasa,” [Skrt: sarovar = lake].56

52 Ramayana 4.42.18–20, Goldman, vol. 4, 152. The version cited here throughout is the four volume edition published by Princeton University Press (1984–1994) under the editorship of Robert P. Goldman; for details of individual volume editors/ translators, see Bibliogra- phy. 53 Ramayana 4.41.32; Goldman (vol. 4: 149). 54 Ramayana 4.42.16 & 4.42.53–56; Goldman (vol. 4: 152). Ramayana 4.42.58; Goldman (vol. 4: 152), states that “Somagiri cannot be reached even by gods.” 55 Ramayana 4.42.26–27; Goldman (vol. 4: 152). 56 Ramayana 1.23.7; Goldman (vol. 1: 170). The later account of this myth in the SkandaPuraṇa as given by Hedin is only slightly expanded, see Hedin (1909: 198–199). “beyond the course of humans” 59

The Epic references to Kailas-Manasarovar are thus brief. The Mahabharata is one of the world’s longest works of prose. It contains more than 90,000 stanzas, of which only a handful contain any meaningful allusion to Kailas- Manasarovar, as is the case with the Ramayana. These references, taken out of context, can be used to construct a pre-eminent status for Kailas, but when we consider them in the context of descriptions of other sites given in the Epics, a very different picture emerges. As well as the Meru, Sudarsana, and Soma- giri mountains already noted, we must take into account the “most majestic golden-peaked mountain … sixty-four leagues high … [on which there is] a city of pure gold … that king of mountains”—Mount Varaha,57 as well as, “a mountain sacred beyond all other mountains, the holy mountain of Mahen- dra.”58 There are actually dozens of mountains referred to in such exalted terms, including Gandhamadhana and Himalaya. What these references demonstrate is not a specific knowledge of Himalayan geography, but henotheism, the worship of deities each praised in turn as endowed with supreme characteristics when it is the focus of praise. So Kailas is praised as the supreme mountain in the Epics, but so are many others. The entire Himalayan mountain chain now had sacred status and as the “Tour of the Sacred Fords” section of the Mahabharata indicates, many other mountains elsewhere in India were also lauded as sacred. There is no indication in these texts that any particular mountain was then regarded as paramount. Thus while the Epics do indicate a knowledge of a mountain known as Kailas, that knowledge is entirely unsystemised. The mountain is not specif- ically located, nor is it endowed with any particularly elevated status and descriptions of its location are geographically inconsistent. Gandhamadana is actually the most prominent peak in the relevant sections and both Epics men- tion Manasa as a mountain.59 There is a Lake Manasarovar mentioned in the Mahabharata, but it is one situated near Prayag. It is also notable that while the Ganges, Sarasvati, and Jamuna are the most important rivers in the Epics, the Indus, Sutlej, Tsangpo-Brahamaputra and Karnali rivers—all famously arising in the Kailas region—are notable by their absence. The divine centre for these texts is in any case not in the mountains, but the north Indian plains. In most Epic references, Kailas is not an actual mountain, but a metaphor for strength and purity; characters are “as tall as the Kailāsa peak”, or “white-

57 Ramayana 4.41.24–27; Goldman (vol. 4: 149). 58 MhB. 3.(33).85.15–20; van Buitenen (vol. 2: 400). There is a Mount Mahendra in Orissa, on which see, Rath (1976: 80–98). 59 MhB. 3.(37).213.1–20; van Buitenen (vol. 2: 647); Ramayana 4.42.26–27; Goldman, (vol. 4: 152). 60 chapter 2 complexioned as the peak of Kailās.”60 In other passing references it is men- tioned as a source of herbs,61 or a place of refuge for a king,62 and while the mountain is identified on occasion as the residence of a deity or deities, the scattered references are not predominantly to Śiva, but to Kubera.63 This deity (see Chapter 14), leader of the yaksa spirits and originally associated with out- casts and criminals,64 later became associated with treasures and found a place in the pantheons of each Indic religion. In sum, Kailas and Manasarovar have a very minor role in the Epics and, we must assume, in elite or Sanskritic Indic culture in the early centuries of the common era. Not only Kailas’s location, but its status is unclear at this time. It is not Meru or a cosmological centre, it is not particularly exalted in comparison to other mountains in the region, and it is not closely associated with Śiva. The latter features but little in these sagas, and is usually referred to under such epithets as the ‘Three-eyed Lord’ or the ‘Blue-Throated God.’ Although a King Sagara, performing austerities at Kailas, approaches Śiva for a boon, and in another instance, a King Bhagirath travels to Kailas to placate an angry Śiva, there is no specific statement in either Epic that Śiva actually resides, or at least resides exclusively, on Mount Kailas.65 References to Kailas in the Epics do not, therefore, refer to the mountain that we know today as Kailas (-Manasarovar). Not only are they incompatible with geographical realities, but Kailas is Kubera’s mountain, not Śiva’s. It is not the cosmological centre of the world and although the reference to it as the gathering place of the gods could be an indication of its growing status other locations such as Meru are also referred to as the gathering place of the gods in other sections of the text, as in a brief version of the famous creation myth of the ‘churning of the oceans’.66 Kailas is, at this stage, only one of a number of mythical mountains in a heavenly realm.

60 MhB. 1.212.20; van Buitenen (vol. 1: 408); MhB. 5.154.15; van Buitenen (vol. 3: 472). 61 MhB. 2.48.5–10; van Buitenen (vol. 2: 118). The reference is to a “northern” Kailas, seemingly implying the toponym is known elsewhere. 62 MhB. 3.(33).104.10; van Buitenen (vol. 2: 424). 63 Kubera is also stated in the Mahabharata to reside on at least one other prominent mountain, Mount Mandara; MhB. 3.(33).140 1–5; van Buitenen (vol. 2: 490). For a complete list of references in this Epic to Kailas, Manasarovar, etc., see the relevant sections of Sorensen (1904). 64 Śatapatha Brāhmaṇa, 13.4.3.10; quoted in Sutherland (1992: 63). 65 MhB. 3.(33).104.5–15; van Buitenen (vol. 2: 424); MhB. 3.(33).107.20–25; van Buitenen (vol. 2: 429). 66 MhB. 1.(5).15.5–10; van Buitenen (vol. 1: 72); also see re Meru, MhB. 3.(41).247.5–10; van Buitenen (vol. 2: 703). “beyond the course of humans” 61

The Epics are ultimately works of fiction. A coherent geography is not essential to their narrative, which deals with mythical heroes on a mythical landscape. But in one respect the Epic accounts of the Himalayan region are entirely consistent. They clearly and distinctly state that this was the abode of gods and demi-gods; it was not a place to be visited by ordinary (i.e. caste Hindu) pilgrims. While, as noted, the difficulty of travelling there is accentuated in the later of the two Mahabharata accounts of Arjuna’s Himalayan sojourn, both versions agree that, “this road cannot be travelled by mortals … this is a divine pathway, no humans travel it.”67 As Arjuna’s party were told when they reached this, “play garden of the gods, beyond the course of humans”,68 “mortal man cannot desport himself here.”69 The Himalayas, like the sacred mountains of China noted in the first chapter, were thus demarcated as beyond the known, the safe, the pure. As the gods’ heavenly abode they were naturally a radiant paradise;

golden and bright as fire, full of all kinds of jewels … [with] trees crowded with birds, always laden with fruit and flowers, whose fragrance, taste and touch are heavenly, yield every desire.70

Yet paradoxically, being located on the physical plane, the mountains were simultaneously wild and dangerous, remote and difficult of access. Thus to get there required crossing “horrifying wasteland”, encountering “flies, gnats, mosquitoes, tigers, lions and snakes”, avoiding or fighting “awesome beasts of prey”, “hordes of barbarians” and “swift and terrible” demons; gandharvas, yakṣas and rākṣasas. There could be no cause to idly traverse such a place, it was a destination, and beyond it lay only the void where, “we know not what lies beyond, boundless and untouched by sun.”71 The frequent references to these dangers were not mythological. The moun- tain wilderness really was a place of altitude sickness, tigers, snakes, and land- slides, as well as robber bands and tribes hostile to outsiders (tribes who might be conceptualised as demons). Thus it really was a place of danger and sudden death,72 and we know that in recent centuries those who ven-

67 MhB. 3.(33).147.40; van Buitenen (vol. 2: 504). 68 MhB. 3.(33).156.20–25; van Buitenen (vol. 2: 524). 69 MhB. 3.(33).152.5–10; van Buitenen (vol. 2: 517). 70 Ramayana 4.42.42–44; Goldman (vol. 4: 153). 71 Ramayana 4.42.59; Goldman (vol. 4: 154). 72 The paradoxical images of the forest are explored in Ramayana 2.24.10–25; Goldman (vol. 2: 56). Sita, wanting to accompany Rama in his forest exile, invokes the romantic 62 chapter 2 tured there usually sought safety in numbers, travelling under the protection of a powerful aristocrat or with an armed caravan. A similar situation is indi- cated in the Mahabharata, which relates how a group of Brahmans plead with the Pandavas for permission to accompany them on their tour of the Sacred Fords;

Pray take us with you, great King Pandava, for we are unable to go to them without you … The straits and the perilous places are infested with beasts of prey and the fords cannot be reached by small companies of travellers … [But] under the protection of you heroes we too might reach them.73

Further evidence that the realms Arjuna and Rama visited were not places of pilgrimage is that whereas various benefits and boons are described as accruing to those who visited pilgrimage sites such as Pushkar, no such profit is mentioned in regard to a Himalayan journey. It was the home of gods, not to be trespassed upon by ordinary men. But while to many these mountains were “inaccessible even in their imag- inings”,74 they are not described as devoid of those who had apparently once lived a human life. The land of the gods was accessible to certain individu- als, for the gods shared it with divinised beings who had lived a human life. These were “great seers radiant as the sun … [who] look like gods and are wor- shipped by the gods themselves.”75 These “seers” were renunciates, presumably once Brahmans dwelling in the world, who had renounced worldly existence and mastered tapas, gaining that ultimate magical power that could only be acquired through the practice of austerities. It was that power that qualified them to live in the land of gods. Without it they were excluded, for as Arjuna was told when he arrived at the sacred realm, “no one who has failed in auster- ities [tapas] can reach that region.”76 In the “play garden of the gods” the skills of the ordinary individual human were worthless. Arjuna, famed warrior, was told that;

image of “honey-scented forests.” The reply of Rama, who knows the reality, repeats the refrain “the forest is a place of utter pain” as he describes the hardships and dangers of life there. 73 MhB. 3.(33).91.1–10; van Buitenen (vol. 2: 406). 74 MhB. 1.(5).15.5–10; van Buitenen (vol. 1: 72); the reference is to Meru. 75 Ramayana 4.42.25; Goldman (vol. 4: 152). 76 MhB. 3.(33).142.25–30; van Buitenen (vol. 2: 493). “beyond the course of humans” 63

Weapons are of no use here; this is the land of the serene, of ascetic Brahm[a]ns who control their anger and joy.There is no use for bows here, nor for any fighting. Lay down your bow, you have reached the end of your journey.77

The Epics tell the sagas of Heroes, they are archetypal myths of an individual who journeys to an unknown land and returns with his power enhanced or regained. The typical Hero of Indian mythology must be capable of great feats of asceticism, the basis of his power.78 Arjuna is such a Hero, a great warrior, famed as a bowman. But when he leaves the realm of men, his worldly skills are of no account. Though he is tested in battle, it is understood that it is Arjuna’s power of tapas which enables him to prosper in this heavenly land. Only that power enables him to venture beyond the usual social boundaries into this realm of the gods. Only his dual nature enables him to confront the wilderness,79 to conquer it, and to return to society. In the Ramayana, King Rama, the ideal Hero, is a similarly contradictory figure, bow-wielding Ksatriya king fused with Brahman renunciate, wearing the symbols of the ascetic. That duality is essential, for the return, for the ideal ‘Hero’, and as a literary device. Incorporating the dual nature of warrior and renouncer, Arjuna and Rama could act with the necessary “authority and self-restraint”, or “fire and austerity”,80 the dichotomic qualities they would need for the journey beyond the worldly realm. We have seen how liminal figures such as the vratyas ventured beyond the ‘known’, and through their ritual practices brought previously unknown areas into Indo-European knowledge, sanctifying them as Aryan landscape. Only they were equipped to deal with the impure wilderness and the polluting out- casts who lived there. As Nancy Falk, analysing the concept of the ‘wilderness’ in Indian mythology, concluded, “[o]rdinary people avoided such spaces … In general, only ascetics took on the wilderness directly.”81

77 MhB. 3.(31).38.30–35; van Buitenen (vol. 2: 297). 78 Katz (1989: esp., 90). 79 Re the wilderness, see in particular Falk (1973: 1–15); also see Bharati (1978: 77–82); Katz (1989: 90–103); and Feldhaus (1990: 97), who notes that while considering the forest- dwellers as impure, the Brahman authors of the Deccan river mahātmyas also had, “an almost romantic idealisation of their religious and moral qualities”, an Indic parallel to the European Enlightenment idea of the noble sauvage. 80 MhB. 3.(33).140.10–15 & 141.1–5; van Buitenen (vol. 2: 490–491). 81 Falk (1973: 1). 64 chapter 2

As mastery of the wilderness brings worldly power over it, there also existed a complex relationship between kings and wilderness, with the king having “to transact with it to acquire or hold kingship.”82 There are many examples in Indic mythology of royals who carve a realm out of the wilderness or who lose their kingdom and only regain it after a sojourn in the wilderness. In the historical sphere too, many kings made journeys to Himalayan sacred sites that were heavy with the symbolism of their right to rule.83 Renunciates’ Himalayan journeys, crossing normal social boundaries into the realm of the gods, were likewise demonstrations of spiritual power. In sur- viving and conquering the wilderness, they proved to their culture that they possessed the “power of their austerities which lights up the land devoid of sun and moon.”84 Renunciates acted as pioneers, able to go beyond cultural fron- tiers, negotiate with the ‘wilderness’, and eventually bring it within the domain of ordinary citizens, whom the renunciates could protect by their power as intermediaries with the gods. And if they failed to return, the renunciate was understood to have entirely severed their ties with the world, disappearing from it, transcending it; he or she (and this symbolic phrase is used today to describe the passing of a renunciate), ‘went to Kailas.’ The Mahabharata and the Ramayana recount the journeys of Heroes. Ar- juna and Rama are Hero figures, indeed they are gods. As possessors of ascetic powers they are equipped to ‘transact’ with the wilderness. Quite clearly they are no ordinary pilgrims, and there is no suggestion that their journeys are within the capabilities of the ordinary pilgrim; on the contrary it clearly states the reverse. Thus there is an Epic Kailas, but it is a mythical place, an imagined heavenly realm which exists as a literary device. It is not part of a cohesive body of knowledge and if it contains some basis in reality that basis does not seem to be related to the modern Kailas-Manasarovar in Tibet.

82 Falk ibid. 83 See, for example, Lecomte-Tilouine (2009: 183); also see Sharma (2009: 26, 53). 84 Ramayana 42.31–36; Goldman (vol. 4: 153). chapter 3 Recreating the Divine Order: The Puranic Kailas

Introduction

Around the time of the final formulation of the Epics, a new Indic literary genre began to emerge, the Purāṇas. These texts presented a smorgasbord of philo- sophical positions and conceptions of the Absolute, and advanced increasingly detailed descriptions of the heavenly realms, including several versions of the cosmological model of the universe centred on Mount Meru. The Puranas reflect a culture in which references to Kailas multiplied. It acquired a set of signifiers and as a literary and mythical device Kailas proved pliant enough to be widely utilised in Sanskritic cultural conceptions and imaginings of the uni- verse. Like the Epics, the Puranas were popular texts appealing to a wide audience, in contrast to works such as the Vedas which were the preserve of the Brahmani- cal caste. Puranas therefore played a major role in spreading an understanding of Kailas in wider Indic society. What is crucial to understanding Kailas as it emerges in the Puranas however, is to analyse these texts as what Ronald Inden called “different claims contesting for hegemony rather than as expressions of a unitary tradition.”1 While broadly ecumenical in their acknowledgement of a range of deities, they were essentially sectarian documents intended to pro- mote the primacy of particular deities, their theology, worship, and associated sacred sites. Most Puranas therefore made no mention of Kailas, presenting competing sacred geographies and world-views. Even those that did mention Kailas offered differing visions of its sanctity and its location; visions which were subject to no classificatory authority. The disparate representations indi- cate that in the pre-modern period there was no single Puranic Kailas, or any pan-sectarian shared vision of the identity or even location of a Kailas moun- tain. In this chapter we discuss the pre-modern Puranic Kailas with particular reference to the ‘Central Mountain, Divine Lake, Four Rivers,’ mythology. This association of Kailas with a lake and four rivers, which emerges fully in the Puranas, is fundamental to its modern status. Yet Puranic citations in support of Kailas-Manasarovar’s ancient sacrality are invariably selective and lacking in

1 Inden (2000: 91); also see Inden (1992: 556–577).

© koninklijke brill nv, leiden, 2015 | doi: 10.1163/9789004306189_005 66 chapter 3 wider context. That context demonstrates that these texts do not support many essential aspects of the modern construction of Kailas.

The Political Context

While the Epic texts (and, as will be seen in Chapter 5, contemporary Buddhist literature) suggest some knowledge of actual Himalayan geography and tribal populations, there are no reliable records concerning the nature or holders of political power in the Kailas-Manasarovar region before the 7th century ce.2 From as early as 200bce., however, much of Kumaon and part of the Garwhal areas to the south of the Kailas region was influenced by the Kuninda dynasty and, for several centuries more after their collapse, by their successors the Katyuris. The Mahabharata refers to Kunindas as one of the mountain tribes offering gold, as well as fly-whisks (commonly derived from yak’s tails), and to their having neighbours who offered herbs and holy waters.3 These commodities all associate them with the source of rivers such as the Ganges and even the Kailas region, or at least with their having trading links to those areas. According to copper plate evidence and the testimony of Chinese pilgrim Hiuen Tsang, the Katyuris ruled as far north as Tapovan (above Joshimath).4 This would have given them a knowledge of the headwaters of the Alakananda branch of the Ganges that flows past Badrinath to Joshimath. In the case of the Kundindas, Maheshwar Joshi concludes from numismatic and epigraphic evidence that their authority reached from the Sutlej to the Kali river, and that the kingdom gained prosperity by acting as an exchange point for trade goods between the mountains and the plains. But while that role is consistent with that of later regional polities, ascribing the forms of power manifest in later polities to such pre-modern formulations, particularly in the hill regions, is problematic. Claims to authority in that period refer more to centres than peripheries and the Kunindas and Katyuris were probably loose confederacies rather than kingdoms or empires, with their population centred in the lower hill districts, as with later polities in that region.

2 Pranavananda (1983: 218) refers to an [unlikely] account of the Kumaon Raja annexing Kailas under Asoka’s orders; his sources are not given. 3 Joshi (1988: 73–86); (1989: 22); (2009: 327); Joshi’s latest work dates the Kuninda collapse to around the 3rd century, whereas his earlier work favoured a date later by several centuries. 4 Walton (1911: 164); Pranavananda (1984: 218); the copper plate is from Pandukeshvar, a village between Joshimath and Badrinath. This is not to be confused with the Tapovan above Gau- mukh. the puranic kailas 67

Culturally, however, these dynasties were influential. Their elites at least held to Sanskritic values, demonstrated when as late as the 3rd century ce., a Kuninda king performed a horse sacrifice (aśvamedha), the Vedic rite of kings. Kuninda coins also indicate that their primary focus of worship was a Sanskritised local deity, Chatreśvara.5 Their realm may therefore, have had an intermediary role in the increased knowledge of Kailas that emerged in the late Epic period, while the length of their reign also suggests that over that time some knowledge of their values was transmitted to the peoples of the western Tibetan plateau.6 There were other groups in the upper Himalayas about whom we know lit- tle, such as the (non-Indo-European) Kiritas, mentioned in the Atharvaveda and identified in the Mahabharata as a people dwelling near the sources of the Ganges.7 In addition, there were population groups moving into the hill regions such as a branch of the Licchavis, a republican tribe originally linked to the vratyas in the Magadha region.8 The Kuninda and Katyuri periods thus saw multiple encounters between different population groups and belief sys- tems, but there is no evidence from this period unquestionably concerning the upper Himalayas. Of later sacred centres Badri(nath?) was most frequently mentioned. But like Kailas, Manasarovar, and many other places, the corre- spondence of such toponyms with modern sites of those names is uncertain and none was articulated as more sacred than another. In addition, we know that at a later date certain regions on the Indic side of the Himalayan water- shed were under Tibetan control, which may have been the case in this ear- lier period. Kailas-Manasarovar itself certainly remained outside of Sanskritic political authority.

5 Joshi (1986: 196–197, 209); Joshi (1989: 22–23, 32–33, 50–75, 100–107); also see Handa (2002: 193–208); Gupta (1988: 17–20, 26); Powell-Price (1930: 5–16); Powell-Price (1945: 213–223). 6 This could suggest a period in which communications between aspects (cosmology? ritual?) of the later Hindu and Bön religions be sought. 7 A “little girl”, (i.e., a virgin?) of the Kiritas is referred to in Artharvaveda 10.4.14 as digging out herbs with a golden shovel, “upon the ridges of the mountains.” The commonality of such references may suggest that ridges held some greater cultural meaning. The Kailas glacier is similarly distinguished from the mountain in certain contexts that suggest wider meaning. Aside from the death of Gri-gum (see n. 26 chapter 11), see for example, Martin (2001: 118). Also see Hein (2007: n. 8 238) who cites a song collected by the missionary A.H. Francke in Namgya (Kinnaur): “Kailasa is the king of glaciers, Purgyul [a mountain above Nako in Kinnaur] is the king of mountains, Manasarowar is the king of lakes.” 8 Kosambi (1952: 182); also see Chandola, (1987: 37), this latter work should however, be used with caution. 68 chapter 3

Religious Developments

One significant factor shaping the multiplicity of accounts of Kailas-Manasaro- var in the Epic and early Puranic period was the development of sects as distinct formations in Indic society. Some elements of this process are relevant to our histories, as will be seen. By definition a sect held to a particular body of knowledge concerning a deity and the means by which it should be worshipped. Sects developed dis- crete conceptions concerning not only the character, history, and associated sacred geography of their particular deity but also philosophical issues such as the nature of existence. While there were cultural structures and understand- ings superimposed across these diverse bodies of knowledge, and a general tendency towards their systemisation, their distinctness was in itself a currency, preserved as a valuable heritage within families and traded or sold to outsiders. Earlier gotras and other teaching lineages had also maintained distinct tradi- tions in regard to deities and their worship, thus sectarian distinction may be best seen as a systemised temple-centred development appropriate to a new socio-political setting, and an effective strategy by which to broaden the appeal of a set of teachings. Probably the earliest reference to sectarianism is from around the second century bce., with the term Śivabhagavat used to refer to itinerant Śaivite ascetics wearing animal skins and carrying lances,9 and the first identifiable sect were the Śaivite Pāśupatas. Mentioned in the Mahabharata (12.335.40) in the context of a doctrinal system revealed by Śiva, the Pasupatas were a Brahmanical group that included those whose spiritual path embraced impure practices and deliberately seeking “dishonour.”10 Thus an association with the Atharvavedic tradition might be expected and we have later evidence of both Pasupatas and Śaivite Atharvavedins being centred in western India (Gujarat, Malwa) in the latter part of the first millennium ce.11 While no evidence sur- vives of their being in the upper Himalayas,12 Ronald M. Davidson gives a list of probable Pasupata sites that indicates their presence in Almora distict and the Garwhal Tehri from the 8th century, as well as at sites in Kulu and in Kashmir.13

9 Dyczkowski (1988: 4). 10 Dyczkowski (1988: 19–26); also see Hara (1994: 323–335). On “dishonour”, see Ingalls, (1962: 281–298). 11 Bisschop and Griffiths (2003: 320, 324). White identifies this region as a later Nath alchem- ical centre, see White (1996: 76–77). 12 Although Sanderson has found Kapalika texts in Nepal; Urban (1995: 68). 13 Davidson (2002: 341–343). the puranic kailas 69

Kāpālika renouncers, first mentioned around the mid-first millennium ce,14 were also predominantly Brahmans of orthodox Vedic background, but wor- shippers of Bhairava (a fearsome manifestation of Śiva). They cultivated impu- rity through cemetery and antinomian practices.15 Their lifestyle meant that in the popular mind they became synonymous with anti-structure and orgy. Kapalikas were, in David Gordon White’s memorable phrase, “the Hell’s Angels of medieval India.”16 Both Pasupatas and Kapalikas lost their separate identity by the 13th/14th century.17 Neither founded lasting structures or institutions,18 nor attracted major patronage in their unrefined forms. Both sects were however, of lasting influence in terms of models and parameters of behaviour. Other groups repro- duced their imitation of the lifestyle and appearance of their focal deity, cod- ified and developed their antinomian practices—generally in forms we now understand as ‘Tantric’—or inherited aspects of their traditions as in the case of the later Nāth siddhas19 and cemetery-dwelling Aghori sect.20 Whether particular sects were concerned with a Kailas mountain depended on their regional focus and sacred geography; most were not. But sects active in the upper Himalayas formulated understandings of the sacred geography of that realm and some sense of territorial competition between them seems apparent, particularly from toponyms. Certainly the Laxmi-Narayan sect were significant agents of Vaisnavite influence, while Śaivite influence seems to permeate from itinerant Atharvavedic-derived traditions such as the Pasupatas and the later Naths and Giris, as will be seen. Whatever may have been the case, the result was the development of the disparate and distinct bodies of knowledge concerning the region which were articulated in the various Puranas.

14 Lorenzen (2002: 27–29). 15 Lorenzen (1972: 80); Lorenzen (1989: 231–235); also see Davidson (2002: 211). 16 White (1996: 306). On such ascetics in medieval Indic literature and popular culture, see Bloomfield (1924: 202–242). 17 Lorenzen (2001: 83) notes that “their association with a specific vow, a specific doctrine, and a special dress code suggests some definite institutional tradition.” Samuel (2008: 243) however, concludes that the term does not seem to refer to “a distinct order of renunciates”, but to a mode of practice. 18 Dyczkowski (1988: 27). 19 Dyczkowski (1988: 6–29, esp., 28). 20 For a modern practitioner’s account of this sect, see Svoboda (1986). 70 chapter 3

Puranas

Modern understandings of a ‘traditional’ Kailas draw heavily on the Puranic literature, the nature of which greatly problematises analysis. Puranas exist not only in Sanskrit but numerous regional languages, are often of only local circulation, and continue to be produced into the modern era. These issues, combined with the sheer volume of these texts, means that although numer- ous individual Puranas have been studied (and translated into European lan- guages), the critical study of the Puranas as a whole has been neglected. Later tradition recognises eighteen major or Mahāpurāṇas, works whose “essence is an account of the divine origin and ordering of the world.”21 But any overall organising principle is absent in practice,22 and Puranas generally comprise cycles of myths and other layers of material added at different times by different interests. Demonstrating that these texts represent distinct bodies of knowledge held by diverse groups and traditions, each Purana contains “an infinite variety of religious trends and combinations of trends.”23 The richly varied array of cosmologies, primary deities, and philosophies they contain within and alongside a monumental profusion of myths and legends means that the immense unbounded corpus of Puranic texts largely defies collective analysis. It is, therefore, misleading to speak of a ‘Puranic worldview’ or to assume these works maintain a consistency of conceptual understanding, or even internal consistency. Parts of the earliest Puranic texts probably overlap with the final compo- sition of the Epics, but “dating a particular purāṇa is highly speculative if not impossible.”24 Yetif we cannot construct a reliable Puranic chronology,the find- ings of one study in particular are relevant to our enquiry. Puranic specialist Rajendra Chandra Hazra convincingly argued that the foundational layers of the early Puranas were concerned with rites and customs on the dharmaśāstric model of earlier texts such as the Manusmṛti, with the variety of their con- tent increasing from around 500ce. That later content included the sections on sacred geography none of which he concludes, are earlier than 700ce.25 Taken with evidence for the late appearance of the Epic descriptions of the Himalayas, these findings indicate that the first tentative systemisation of a

21 Inden (2000: 32). 22 Rao (1993: 87); also see Inden (2000: 31–32). 23 Doniger (1993: 104). 24 Rocher (1986: 103). 25 Hazra (1975: 174–189). the puranic kailas 71 sacred Himalayan geography dates to little earlier than the mid first millen- nium ce.26 Even if we are able to locate the Puranic references to Kailas in a very provisional time framework, the texts must also be located in religio-social context. But as Inden has observed, while the bulk of extant Puranic texts are actually comprised of later medieval and sectarian material, “[t]o a very large extent, the study of the Purāṇas as a genre has consisted of various attempts to suture the texts, at the expense of their medieval contents, to a classical Hinduism as its genuine, original author.”27 An obvious consequence of this approach, and the reconciliation of diverse tendencies within the construction of a unified Hindu religion, has been to obscure the importance of both the non-Sanskritic local and the regional Sanskritic traditions behind the imagined Pan-Indic.28 This approach has been accentuated by philological reduction of the various regional manuscripts to a standard text, a ‘critical edition’ that is probably without historical basis in actual usage and which standardisation only increases the tendency to obscure the fragmented and regional nature of pre-modern Hinduism.29 The Puranas were ultimately sectarian instruments to promote a particular body of knowledge as a universal teaching. They represented traditions deriv- ing from diverse (Brahmanical) gotras or (renunciate) lineage teachings, and generally emerged from or were characteristic of a particular region.30 Hence at the core of most Puranic texts is a claim to centrality and sacrality by a par- ticular regional tradition and its intellectual purveyors.31 Knowledge of Kailas

26 Gombrich (1975: 118) dates the arising of classical Indic cosmology to “after about 500bc”; thus the construction of cosmology and beginnings of its transference to an earthly location (as implied in the construction of sacred geography) seem essentially unrelated processes, with the latter being much later. 27 Inden (2000: 29). See however, Shulman (1980: 4) who cautions that the Tamil texts “are nevertheless, a part of the wider world of Hindu mythology; however different their orientation, however local their concerns, they are by no means independent of the classical Sanskritic tradition.” 28 Local ‘voices’ are apparent in the names and characters of deities that have been inducted into the Puranic pantheon from local belief systems (for example, Manibhadra, who appears in the Matsya Purāṇa 71.9 as the “formidable general” of a class of other-worldly beings). We also see hints of their presence in early forms of toponyms, and in the depic- tion of local tribal peoples who were conceptualised in different ways, sometimes as hos- tile demons, sometimes as demi-deified citizens of a higher realm, or as yoginis in disguise. 29 On which see Witzel (2014); Dodson (2007: esp. 38–44, 54–60, 144–148). 30 On which see, in the context of Tantra, Davidson (2012). 31 For an example of this regionality and of how cosmologies, “are appropriated … to con- 72 chapter 3 and its location within particular sacred geographies is manifest in just such constructions and contexts. Rather than a unitary understanding of a region, its location and significance varied according to particular traditions of knowl- edge.

As place-centred constructions, the category of Purana to an extent overlaps with that of mahatmyās, texts that serve as pilgrimage guides to specific sacred places and which, like Puranas, continue to be produced into the modern era.32 Mahatmyas describe such elements as the mythological origins of the site, legends concerning the activities of the presiding deity there, the rituals for its worship, and the benefits of pilgrimage to that place.33 Both texts are intended to attract worshippers to the sacred centres of a region and the beliefs of a local tradition, and both ultimately sought to locate those features as Pan-Indic. Thus we find the Brahma Purana is largely a mahatmya of Orissa, the Vayu Purana a mahatmya of Gayā, the Nīlamata Purana a mahatmya of Kashmir, and so on, while the Skanda Purana (to which we will return in Chapter 6), was originally centred on Kaśi (Benāres/Vārāṇasī) but came to be a highly organic collection of different mahatmyas.34 (This regional focus explains why numerous Puranas make no mention at all of Kailas—it was simply not relevant to their claims to centrality.) The Puranas and mahatmyas were never guide-books in the modern sense. They neither intended nor were expected to provide geographically realistic descriptions of sacred sites such as Kailas or the routes there. As Gregory Schopen has observed, Indian scriptures actually represent, “nothing more or less than carefully contrived ideal paradigms.”35 Thus these texts deployed a range of cultural metaphors and recycled mythologies to enhance the appeal of their visionary imaginings and sectarian interpretations of the divine. Analysing the Puranas from this understanding gives us fresh insight into their textual references to Kailas and Manasarovar. It confirms that there was

struct the instruments of legitimisation that paves the way for temple-centred pilgrim centres” (Sharma 2008: 138), see Inden’s study of the Viṣṇudharottarapurāṇa, a late first millennium Pāñcarātrin Vaisnavite text centred on Kashmir. This shows a specific regional teaching lineage competing for primacy and court patronage with other such lineages, while incidentally indicating the transition possible from renunciate to temple Brahman; see Inden (2000: esp., 26–40, 57–64). 32 On modern mahatmyas of Gangotri, see Pinkney (2013). 33 Gonda (1977: 277–278). 34 Rocher (1986: 71). 35 Schopen (1992: 5). the puranic kailas 73 no collective Puranic understanding of these sites, rather there were discrete and competing understandings, with Puranas that implicitly denied their cen- trality by entirely omitting mention of them, others that acknowledged them on a periphery, and those—ultimately successful—that sought to centralise them. In addition, we need to remember that although Puranic references to Kailas represented states or constructions of knowledge about the site at the time they were recorded, such aspects of that knowledge as were empirical must have been compiled from disparate sources—renunciates, traders, nomadic tribes, slaves, outlaws, gold-seekers, wandering bards, soldiers of fortune and other such characters who left no direct traces of their experiences. Collated by one or many agents and authors, permeated through different layers of cultural and sectarian perception, and strategically associated with earlier mythological cycles, the resulting representations were transformed into myth. Reading Puranic texts as representations of different conceptual world- views enables us to see the historical diversity behind modern unities. We may thus analyse them alongside texts of the Buddhist, Jain, and Tantric traditions, each of which preserved and represented discreet understandings of Kailas that were only to coalesce in the colonial period.

A Shifting Puranic Kailas

Kailas—and an associated lake—does enjoy some prominence in what is often considered the oldest Purana, the Vāyu.36 But it retains there the vague and geographically imprecise aura of the Epic period, consistent with Hazra’s conclusions concerning the late development of a Pan-Indic sacred geography. Chapter 41 is devoted to a “Description of Kailāsa” (described as having not one, but two peaks, connected by a ridge, a description entirely incompatible with Kailas-Manasarovar37). The description is as follows:

Kailāsa is the abode of the devotees of the lord who are of meritorious souls. It has isolated, charming summits. It is full of vegetation. It has the lustre of the conch. In the central ridge, as beautiful as [the] Kunda flower, there is the invincible city of the noble-souled Kubera, the presiding deity of wealth … Lord Kubera the companion of Mahādeva [Śiva] lives there

36 Rocher (1986: 245) dates the text to circa 5th century in its earliest sections. 37 Though it describes the Kinnaur Kailas; see Chapter 8. 74 chapter 3

… Similarly this mount Kailāsa is the abode of Indra, Agni, Yama, Devas and Apsaras-s where Kubera, the Lord of the Yaksas, is ruler … The great and charming water-reservoir “Mandākinī” is there. It has abundance of water. Its embankment has steps plated with gold and studded with gems. There are golden lotuses … There are excellent rivers, the Nandā and the Alakanandā … endowed with excellent qualities and … used by celestial sages. On the eastern peak of this lordly mountain, there are ten famous cities of Gandharvas endowed with prosperity … On the western peak of Kailāsa … is the abode of each of the Yaksas … The place is resorted to by Siddhas and celestial sages …38

The chapter goes on to describe the world south of Mount Himavat, identified as an “excellent mountain, the abode of many Siddhas … the source of origin of thousands of rivers”,39 as well as various other mountains east and west of this. But as noted, other Puranas were not concerned with Kailas. In the Varāha Purana, a somewhat later text with its earliest sections dating to around the 9th century,40 Mount Meru, located between the continents of Bhārata to the south and Kuru to the north, is more significant. “Kailāsa” appears only in a list of 19 “other mountains around Mānasa” and in a list of 19 “minor mountains” of the continent of Bharata.41 Later Mahapuranas do not necessarily represent conceptual progression in regard to such aspects as cosmology or geographical knowledge. Despite observing the prosaic point that “it is not easy to go from one settlement to another since each is surrounded by steep mountains and rivers”,42 the Varaha ascribes fantastic dimensions to the heavenly region and its peaks. This became characteristic of references to that world, the Viṣṇu Purana (11.2), for example, describes Meru as being 84,000 yoganas43 high and with a summit diameter of 32,000 yoganas. The Brahmāṇḍa Purana, dating as well as can be estimated to the late first millennium, was another that did not advance Kailas’s status. It gives a list of the holiest places of the world according to the god Brahma which does

38 Vāyu Purāṇa 41.1–23. 39 Vāyu Purāṇa 41.44–47. 40 Hazra (1975: 180). 41 Varāha Purāṇa 78.14–16: Varāha Purāṇa 85.3. 42 Varāha Purāṇa 75.30. 43 The precise distance of a yogana is open to interpretation, but perhaps around 8 miles / 13 kilometres. the puranic kailas 75 not include Kailas, Manasarovar, or even Meru.44 Nor is Kailas the abode of Śiva there for elsewhere it states that, “[i]n the midst of the Himalayan ridges, there is a mountain named Kailāsa. The glorious and prosperous Kubera lives there.”45 Kailas as the abode of Śiva was to become the defining characteristic of that mountain, but while the Matsya Purana does describe Kailas as Śiva’s residence,46 the Śiva Purana, a text from around the 9–11th centuries, refers to the deity as residing at Himavat (which may refer to a single peak or to the entire Himalayan mountain range), as well as at Kailas.47 Interestingly, the Śiva Purana, in listing the 12 major sites of Śaivite pilgrimage does not include Kailas or Manasarovar, while the chapter entitled the Kailasāsamhitā largely concerns the duties of renunciates. Clearly the term ‘Kailas’ had a symbolic function beyond the toponymical, but the sacred associations of the site were poorly developed. We might continue to survey the Puranic references to this region or to examine them in more depth in an attempt to plot them on a modern map. But it is clear from many earlier attempts to understand the geography of the Puranas that there is no agreement on the location of different places and that entirely different geographies can be constructed from the texts. S.M. Ali, for example, in perhaps the most detailed study of Puranic geography, concluded that Meru and its surrounds were located on the Pamir plateau (from which region four major rivers, the Oxus, Indus, Tarim and Syr Daria, also descend).48 Given that Puranas referring to the Himalayas, or more specifically to the Kailas region are actually only a small fraction of the Puranic corpus. Kailas in this early period remains an other-wordly site. Elements of both later under- standings and scientific geography are present in the texts, but are neither consistent nor systemised. The later model of Kailas-Manasarovar can only be supported by selective textual quotation entirely removed from context and shorn of inconsistencies and contrary statements. Only in later versions of the Skanda Purana can we unequivocally recognise the modern Kailas, and as will be seen in Chapter 6, those references are the product of very different histori- cal processes.

44 Brahmaṇḍa Purāṇa esp., 70.33–41. 45 Brahmaṇḍa Purāṇa 18.1–7. 46 Matsya Purāṇa 71.1–2. 47 Śiva Purāṇa 40.22. 48 Ali (1983: 49, 51–52). What characterises this and similar works are conclusions in which the Kailas region covers an immense territorial expanse far beyond modern Indian bor- ders. 76 chapter 3

Cosmology

We do not have space for consideration of the complex processes by which Indic cosmological understandings arose, but again there was no single tra- dition, rather there were variations around a general theme, each of which formulations claimed and contested acceptance as ‘true’. Vedic cosmological ideas may be distinguished from the later ‘classical’ cosmologies of Hindus, Jains and Buddhists,49 and these later formulations generally developed two particularly prominent conceptions. In one, the world was envisaged as con- sisting of seven continents with the circular landmass of Jambudvipa at the centre. Jambudvipa was surrounded by water with six alternating rings of land and water beyond. This world of seven continents was surrounded by two other rings which separated the world from the realms of eternal darkness beyond; these rings were Svarṇabhūmi, “the playground of the gods”, and the Lokāloka mountains. In the centre of Jambudvipa was Mount Meru, around which were six parallel mountain ranges running from east to west and two running north- south. The other prominent understanding was of the world as shaped like a lotus, with Mount Meru resembling the seedpod in the centre of the lotus flower, and four continents arranged like petals around that centre. In this version, the continent (varṣa) of Uttarakuru was situated in the north and that of Jambu, or Bharata, in the south. In a realistic note this Bharatavarṣa had oceans on three sides and the Himalayas to the north, like the Indian sub-continent. As well as what became the Hindu traditions, these cosmologies informed those of the Buddhists, Jains, and even Tibetan Bönpos, with the Abhidhar- makośa, composed by the northwestern Indian monk Vasubandhu (4/5th cen- tury ce) becoming the foundational exposition of this system in the Buddhist traditions.50 In the context of Kailas-Manasarovar, the key feature common to these cosmologies was the idea of a central mountain, Meru, which was the source of four (or more) major rivers, the identity of which varied in different accounts. That model did not remain solely attached to Meru, but also became a fundamental part of the geo-sacral associations of Mount Kailas. The distinc- tion between Meru the cosmic centre, Meru an actual Himalayan mountain, and Kailas in its various conceptions and manifestations became increasingly blurred over time. While most texts distinguished them, some referred only to

49 Gombrich (1975: 110–119). 50 Gombrich (1975: 132–133) notes that Vasubhandu’s work was “rather more elaborate” than the Theravāda cosmology of Pali commentaries and Buddhaghosa’s Visuddhimagga from c. 400ce. the puranic kailas 77 one of them and their given attributes were largely interchangeable. As a model Meru proved highly elastic, capable of being used as a symbol or metaphor and most importantly in this case, of being identified with an earthly location. But this central World-mountain was a heavenly one, and thus as with Kailas the Puranic descriptions of its beauty, immensity, lustrousness, etc., are literary devices, religious not geographical conceptions. Later texts developed various lengthy and intricate accounts of these heavenly realms and the often fantastic beings that dwelt in them.51 The authors seemingly sought to outdo each other in amplifying the glories of the sacred realm, for “it was apparently the level of detail that created an aura of verisimilitude sufficient to convince … audiences.”52 They also sought to associate their own sectarian affiliation more closely with that realm and thus elevate the teachings of their sect above its competitors. There was another significant aspect to this cosmology. In the Indic cultural context the divine order was understood to be revealed to man and ultimately accessible to man. Ideally it might be manifest or recreated on earth through the proper performance of dharma. But the fundamental concept distinguish- ing the Indic religious traditions—that man could become a divinity—implied a particular logic. If a man—no ordinary man but a renouncer and master of tapas—became a deity, then he could access the heavenly mountains, descrip- tions of which indicated that they were located at the heavenly source of certain earthly rivers. If the renouncer travelled to the river sources, then he journeyed from the earthly realm to the heavenly.

The Four Rivers

A prime signifier of the sacral status of Kailas-Manasarovar is its proximity to the source of ‘the four great rivers of India’, understood today as being the Brahmaputra, the Indus, the Sutlej, and the Karnali. But in pre-modern texts the river names vary considerably,53 and a study of the processes by which the

51 See, for example, the entry under ‘Meru’ in Desai (1968: 264–265). 52 Campany (2009: 141–142). Frederick Barth (1990: 640, also see 643) discussing the transfer of cultural knowledge by significant individual agents, states that they must “enhance its mass so as not to lose status by running out of materials.” Henk Blezer (2011a: 157) notes that in general, “texts and narratives tend to develop towards increasing elaboration and perceived completeness and consistency”; for an example of this, see perhaps, Mimaki (2000). 53 In reality there are numerous stream or river sources in the vicinity of Manasarovar, as there are numerous sources of the Ganges. 78 chapter 3 modern identification of the four rivers developed demonstrates how differ- ing bodies of knowledge articulated their understanding of Kailas, how those bodies contested acceptance, and how they were reshaped according to cir- cumstances, notably by the encounter with scientific modernity. We have seen the precedent of the heavenly river of early Iranian cosmology, and the importance of rivers to the Vedic peoples, who held them sacred. Their identity was shaped by their residence in the Punjab, the ‘Land of Seven Rivers’, and their explorations probably followed river courses where possible. By the Epic period the primary riverine focus of sacrality in Sanskritic India had become the Ganges, but each region had its prominent rivers. In the northwestern plains’ districts, for example, the Jamuna was prominent, while the Narmada was pre-eminent for the peoples of the Deccan. Regional Puranas exalted their local river(s) as supreme, often equating them to the Ganges and thus contesting its supreme status. Rivers and their sources were key components of Indic sacred geography. It is clear that the concept of a central mountain (and/or its associated lake), as the source of one/three/four/five or seven great rivers derives from archetypal mythologies common to many cultures. In the Biblical chapter Genesis (ii. 10–14) for example, we read of four rivers flowing from the Garden of Eden,54 and as we have seen Korea’s Mount Baekdu is associated with three rivers, while various concepts of such central mountain complexes are given in Chinese sources. Among their most prominent is Kunlun, the ‘snow mountain’ associated with the western quarter. First known from a text dating to at least 281bce,55 conceptions of Kunlun contained many parallels to Indo-Tibetan understandings of Meru/Kailas. In Chinese sources Kunlun, ruled over by the fearsome Queen Mother of the West, was sometimes a range and sometimes a specific peak, sometimes seemingly real and sometimes clearly mythical. It was a source of wealth and power for kings, and a site favoured by renunciates because, like Kailas it was understood to be accessible only to those possessing the necessary magical powers. And, like Kailas or Meru, with which it was often identified by the 4–5th century ce., Kunlun was envisaged as the source of the world’s great rivers.56

54 “And a river went out of Eden to water the garden; and from thence it was parted, and became into four heads. The name of the first is Pison: that is it which compasseth the whole land of Havilah … And the name of the second river is Gihon: the same is it that compasseth the whole land of Ethiopia. And the name of the third river is Hiddekel: that is it which goeth toward the east of Assyria. And the fourth river is Euphrates.” 55 The cited text is the Mu T’ien-tzu Chuan (“Biography of the Son of Heaven Mu”); Porter (1993: 73–106). 56 On Kunlun, see Porter (1993); also see Kleeman (1994: 226–238); Stein (1990: esp., 209–272, the puranic kailas 79

As a conceptual model derived from archetypal mythologies Kailas (or Meru) as the source of great rivers was articulated in various formulations,57 not least in regard to the number of rivers (the identification of which is often prob- lematic). Five or seven were particularly common in earlier sources, with a Pali Buddhist text referring to the Oxus, the Tarim/Yarkhand, the Rapti (in Nepal?), the Sarju, and the Mahi (near Gaya?), while Fa-hsien, the Chinese Buddhist who travelled through northwest India around 400ce., refers to the Ganges, Jumna, Sarju, Airavati (the Rapti? or Ravi?) and the Mahi.58 As late as the 16th century a Tibetan source refers to five rivers;59 the commonly listed Ganges, Sindhu (Indus), Sita,60 and Vaksu (or Chaksu; Tib: Paksu), with the addition of the Candrabhaga (Chenab). Within Brahmanical literature the Matsya Purana also advanced a concept of seven rivers; Sita, Vaksu and Indus flowing west, Nalinī, Hrādinī, and Pāvanī flowing east, with the Bhagirathi (Ganges) flowing to the south. The influence of the Abhidharmakosa meant that its river quartet of Ganges, Sindhu, Sita, and Vaksu became the dominant Tibetan Buddhist formulation and this model also passed into Bön.61 But Brahmanical texts such as the Visnu

n. 106 341). (Kunlun is also the Chinese term for a black person, on which see Don J. Wyatt. 2009. The Blacks of Premodern China. University of Pennyslvania, Philadelphia.) 57 Brashier (2001–2002: n. 47 177) notes that lists of the Five Marchmounts of China were similarly inconsistent. 58 Hoey (1907: 41–46). 59 See van der Kuijp (2004: n. 83 327). Van der Kuijp cites the Abdhidharmakosa five-fold listing of rivers (from L. de La Vallée Poussin (ed./trans) 1971. L’Abdhidharmakośa de Vasubhandu. Tome ii, Mélanges chinois et bouddhiques, vol. xvi. Brussels: Institut Belge des Hautes Études Chinoises; n. 4 147–148), in addition to a seven-fold classification (from J. Schneider. 1993. Der Lobpreis der Vorzüglivh-keit des Buddha. Udbhaṭasiddhasvāmins Viśeṣstava mit Prajñāvarmans Kommentar. Bonn: Indica et Tibetica Verlag; 205–207). The latter classification is identical to that in the Matsya Purana. 60 While the Sita river is commonly identified with the Tarim (e.g.: Newman 1987: 309–312) there is also a Śītā river in Kashmir that has been identified with the Alakananda; see the discussion in Hartzell (1997: 994–998). Hartzell favours the identification of the Sita with that in Kashmir (although if my reading is correct there is some confusion here over the two Srinagars, in Kashmir and in Garwhal). 61 Cutler (1996: 91) notes this is the schema in both the mDo ’dus … (late 11th century) and the mid-19th century Bön guidebook to Tise, Tenzin Rinchen’s ’dzam gling gangs Tise dkar chag …; also see for example, Martin (1995: 52–56). Clearly we are not dealing with geographical realities or even necessarily the immedi- ate Kailas-Manasarovar region. While the text clearly states that the rivers all originate in a lake south of Gandhamadana (Abhidharmakośa 3–57, cited in Pruden [1988–1990: vol. 2, 456]) the Abhidharmakosa quartet are most commonly identified respectively with 80 chapter 3 and Varaha Puranas more commonly identify them as the Alakananda, Sita, Badra, and Paksu rivers. In any case by the late medieval period the geomantic construction relating the rivers to the four cardinal directions emerged as the foremost conception in the Indo-Tibetan world. But there was no consensus over the sacred geography of the Kailas complex within the Puranas either as individual texts or as a genre. The Vayu Purana, for example, describes the “great and charming water-reservoir Mandākinī” (asso- ciated with the “excellent rivers, the Nandā and the Alakanandā”), as being near to Kailas. It also refers to a Mandakinī river that flows from the Manda lake at the foot of Kailas. But in addition it describes a sacred celestial stream that cir- cles Mount Meru and falls upon its “four northern peaks”, whereupon it divides into four. One of these divisions, the Alakananda, “flows over Gandhamādana, the lord of mountains in the south … It enters the northern lake Mānasa …” and then flows through various mountains including Kailas and Himavat, the latter identified as the source of “thousands of rivers”, including the Ganges.62 In contrast the Matsya Purana has seven great rivers formed by the heavenly river arise from a Lake Bindusara, situated north of Kailas mountain at the foot of a mountain “with golden peaks.”63 Further examples might be given, but it is already clear that although these texts referenced earthly toponyms they refer- enced a separate conceptual world to that of scientific geography. The modern ‘four great rivers of India’ only represents one strand of those conceptions. Nor do the associations of the four rivers with particular creatures shed light on geographical realities. Buddhist texts usually identified the Ganges with an elephant, the Sita with a lion, the Sindhu with a bull (or ox), and the Vaksu with a horse,64 although there were variations.65 The association of these creatures

the Ganges, Indus, Tarim/Yarkhand and Oxus. Neither the Tarim/Yarkhand or Oxus rivers have any apparent connection with Kailas-Manasarovar and the belief that the sources of these two rivers were related to Kailas suggests that the mythological schema embraces some very different imaginings. 62 Vāyu Purāṇa 41.14–18; 47.2–3. 63 Matsya Purāṇa 71.24–42; also see Agrawala (1963: 201–203). A similar understanding is found in the Brahmaṇḍa Purāṇa 18.25–27, and see 18.70b–71a, where Lake Manasa is identified as the source of two (otherwise unknown) rivers, the Jyotsnā and the Mṛgakāmā. 64 Staal (1990: 278) citing the Mahāprajñāpāramitāśāstra; also see Van der Kuijp (2004: 326). 65 See the Tang dynasty source, the (8th century)Ta-T’ang-His-Yu-Chi (“Records of the West- ern World”), cited in Allen (1982: 30–32); also see Hedin (1917–1922: vol. 1; 81–83). While giving the standard version of the rivers’ toponyms and the first two associations, it then links the Indus with the elephant and the Ganges with an ox. Similarly the much later 1896 Tibetan guidebook to Kailas used this schema, with the exception of the Sindhu becoming associated with the peacock rather than a bull; Filibeck (1988: 72). We may also note that the puranic kailas 81 and rivers was explained either as due to the rivers emerging from mountains resembling those creatures, or through a link to their qualities. In the case of the lion and the Indus, for example, it was said that, “those who drink the water of the Indus … become heroic like a lion.”66 But the four animals are actually a standard set of considerable antiquity. The same creatures were carved on the four sides of the base of the Asokan pillar that once stood at Sarnath,67 and thus they represent yet another common mythological trope applied to enhance the sacrality of Kailas. Mahesh Sharma, however, identifies a scheme of deity vehicles (vāhana) attached to each of the river goddesses in the region (who include Sindhū and Śatuldhārā, goddesses of the Indus and Sutlej, respectively).68 While his schema is apparently later it does raise the possibility that the creatures asso- ciated with these rivers reflect a World-religion overlay on older or local sacred identifications. (A link to tribal totems might also be considered.)69 We may conclude that Chinese, Tibetan, and Indic traditions preserve nu- merous versions of a basic myth; a central World-mountain (usually with asso- ciated lake) that is the source of a civilisation’s great rivers. Such sacred geogra- phies cannot be aligned with scientific geography or their internal contradic- tions reconciled. To attempt to do so is largely to miss the point of their inten- tion. Sacred geographies are not required even to be internally cohesive, for they were never intended to function as maps. They are the product of separate and competing traditions utilising cultural models amenable to many different productions. While their authors were operating within a general conceptual framework, their sources, intellectual and spiritual heritage, world view, and social determinants oscillated throughout their lengthy composition period. Yet, anticipating the theme of Chapter 15 we may explain why it is accepted today that the four rivers with which Kailas is associated are the Indus, Brahma- putra, Sutlej, and Karnali. During the course of the 19th and early 20th centuries,

W.W. Rockhill, the early American scholar-diplomat and Tibetan traveller, cites Chinese sources that describe four mountains, conterminous with Kailas, which are shaped like a horse, elephant, lion, and peacock; Rockhill (1891: 255–256). Rockhill’s Chinese source is unclear; it refers to “Kang-ti-ssŭ shan” (ie; Gangs Tise/Kailas mountain), as “unques- tionably the greatest of all mountains. In Sanskrit books, it is called A-o-ta (Anavatapta) mountain [sic]”; suggesting a confused Chinese (or author) rendering of an Indic source. 66 Pranavananda (1949: 16). 67 Dietz (1988: 114). 68 Sharma (2009: 39). 69 In that the creatures represent the cardinal directions such schema are found elsewhere, see for example Wayman (1990: 319). 82 chapter 3 the Kailas region became the subject of geographical enquiry by Western sci- ence. A series of British and Indian explorers mapped the region and estab- lished its physical geography in considerable detail. Although in several cases there was continuing controversy over which tiny stream could be properly termed their actual source, it was established that these four rivers all took their course within fifty miles of Mount Kailas.70 In determining that these were the four rivers with which Kailas was ‘properly’ associated, Western knowledge fixed their location in time and space. What was once myth became geography. Extraordinarily, although an earlier source cannot be ruled out without a survey of the entire corpus of Indo-Tibetan literature, it appears that the first reference to the quartet of Indus-Brahmaputra-Sutlej-Karnali as the four rivers of Kailas may be in modern sources. This quartet had one notable omission that distinguished it from most Indo-Tibetan versions—it did not include the Ganges. Although often absent from Puranic cosmology, the majority of later Indo-Tibetan accounts agreed on the Ganges’ place in the ‘four rivers’ schema. But by the early 19th century the Ganges source(s) were established as being on the southern watershed of the Himalayas. This had numerous political and religious consequences that will be discussed later, but we may note here that the Karnali, which flows from Tibet to Nepal before entering India and joining the Ganges at Doriganj (west of Patna) in Bihar, is rarely mentioned in early Indo-Tibetan texts concerning the Kailas region. It owed its place in this schema entirely to scientific logic being applied to the identification of a quartet of rivers, and was thus conceptually a very poor substitute for the Ganges, the great river of northern India. Even today guide-books and accounts of Kailas-Manasarovar largely ignore the Karnali. The Ganges was not entirely ejected from the sacred schema, however. Firstly, the channel that connected Manasarovar with Lake Rakas Tal was also known as the Ganga chu in Tibetan sources. Within the Indic world the Ganges was re-integrated into the sacred geography via another cycle of myths that, like the Meru model, were a device that could be applied to many different locations, or in this case, rivers. The myth was that at Kailas, at Meru,71 or at an associated lake, the Ganges issued from the foot of Visnu (in Vaisnavite accounts) or the hair of Śiva (in Śaivite accounts) before dividing into four rivers.72 Modern Indian accounts frequently follow this mythology, or suggest underground tunnels running from Manasarovar to Gaumukh.

70 See for example, Pranavananda (1946: 168–180). For evidence that such issues remain current, see Bellezza (1993: 41–44); also see Lahiri (2006: 238–254). 71 For a list of Puranas describing the Ganges descent on Meru, see Eck (2012: n. 49 506). 72 This is one of the most frequently cited myths in the regional sacred geography. The the puranic kailas 83

One important result of this process was that having established which four rivers had their sources at Kailas, the academic study of earlier references was now empowered to classify them, to explain and correct their ‘mistakes’, or to seek to demonstrate that earlier texts referred to other locations. In a 1986 English translation of the Brahma Purana by a team of Indian scholars, for example, the reference to the Mandakini as having its source at Manasarovar is ‘corrected’ into a reference to the Ganges.73 What is probably evidence of an attempt by a Puranic chronicler to increase the status of the Mandakini river, occurring also in the Vayu and Matsya Puranas as we have seen, was, in an attempt to fit the classical model, rendered an error! The authority to divine the intention of Puranic scribes was also claimed by modernist Indian scholars who accepted the scientific construction. In order to reconcile text with geographical reality, even the eminent explorer of the region Swami Pranavanada (see Chapter 10), was left to conclude that references in a Tibetan text74 to the Ganges, Sindhu, Vaksu and Sita rivers were actually to be identified respectively as the Sutlej, Karnali, Brahmaputra and Indus!

The Sacred Lake

As noted in the previous chapter, Kailas and Manasarovar are associated in the Ramayana, which states that, “on Mount Kailas there is a lake … Mānasa.”75 Later in the text, when Rama is told where he must travel, a lake, not specifically named but found at Kailas, is described in the following terms;

Near [Kubera’s heavenly dwelling] is a vast lotus pond filled with red and blue lotuses, crowded with geese and ducks, and frequented by hosts of apsarases. And there the majestic giver of wealth, King Kubera Vaiśrava- ṇa, king of the yakṣas, honoured by all beings, enjoys himself along with the guhyakas.76

details vary as to where the river actually descends etc., and it is frequently associated with penance, usually by or associated with a King Bhaigirath, to release the waters when they have been halted by a deity or a rsi; see for example Vayu Purana 47.32–35; Ramyana 1.37–43; or for a modern take, Sundaranand (2001: 48). 73 Brahma Purana 13.1–7; (1983: 166–167). 74 Not identified other than as Kangri Karchhak (i.e.; Kailas guide), although citing Bud- daghosa; see Pranavananda (1949: 15); but also see Snelling (1990: 56). 75 Ramayana 1.23.7; Goldman (vol. 1: 170). 76 Ramayana 42.21–22; Goldman (vol. 4: 152). 84 chapter 3

As noted in the previous chapter, a similar, again unnamed lake which is “the play-garden of Kubera”, appears in the Mahabharata.77 Shared combinations of particular attributes, such as swans and lotuses, imply a unity to these accounts. Yet the Epics did not firmly establish the connection of Kailas with Man- asarovar in the Indic tradition. In the Vayu Purana a Lake Manasa was men- tioned but said to be located to the southwest of Kailas where it was associ- ated with a Mount Vaidyuta.78 (This text also mentioned a Lake Manasarovar as watered by the Alakananda river and in a long list of tirthas described as suitable places to carry out funeral rites [śrāddha] during pilgrimage.) In the Varaha Purana, Manasarovar was identified only as the southern of four lakes in the “four great mountains” surrounding Meru.79 In the Matsya Purana it was again the Vaidyuta mountain which had “at its foot … the holy lake Mānasa, frequented by Siddhas”,80 and Manasa is also mentioned in a list of 18 pilgrim- age sites. In the Brahmāṇḍa Purana we read that the “sacred, splendid and chill water originating from the foot of the Kailāsa mountain has formed a lake named Mada … The auspicious river Mandākinī rises from that divine (lake).”81 The Matsya Purana, as noted, also refers to another lake, Bindusara, situated to the north of Mount Kailas at the foot of a “mountain with golden peaks.”82 Again we are dealing not with a search for meaning or place, but with different cycles of myths expressing competing visions and alternative sacred geogra- phies. One largely unconsidered aspect which is perhaps the most powerful argu- ment against the Epic and Puranic texts describing the area around modern Kailas-Manasarovar is that none of them refer to the existence of two adja- cent lakes to the south of Kailas. Yet in modern times the plain to the south of this mountain is dominated by Manasarovar and Rakas Tal. (The distance between the two lakes today is between two and five miles.) If the Epic and Puranic descriptions of the region really were informed by the geographical reality of the Tibetan Kailas region, the absence of any such reference seems inexplicable other than through geomorphic change. It is geologically possible that the two lakes were once one (as noted, they are linked by the narrow, intermittently flowing channel of the Ganga chu).

77 MhB. 3.(33).151.1–10; van Buienen (vol. 2: 510). 78 Vāyu Purāṇa 47.12.13–14. 79 Varāha Purāṇa 78.9. The model is reminiscent of the Chinese. 80 Matsya Purāṇa 71.16–17; also see the identical statement in Brahmaṇḍa Purāṇa 18.15. 81 Brahmaṇḍa Purāṇa 18.2–3. 82 Matsya Purāṇa 71.25–39; also see Agrawala (1963: 113, 201–203). the puranic kailas 85

This was the conclusion of Swami Pranavananda, who had some geological knowledge. Most recent secondary sources have followed his conclusion, sug- gesting that “Rakas Tal was originally part of Manasarovar but land movements separated the two water masses.”83 While this conclusion is consistent with the evidence for geological change in the region, the absence of any ‘myth of division’ to explain what would have been a dramatic and highly significant transformation is an obvious anomaly. In the Pali Buddhist accounts (discussed in Chapter 5), it is a Lake Anotatta that is most prominent. A Kailas is mentioned as one of the mountains sur- rounding Anotatta,84 and this lake has traditionally been interpreted as being Manasarovar.85 Frits Staal proposed that the earliest Buddhist visitors entered the region from the Himalayan passes to the south. Observing that the Sutlej river flowed from Rakas Tal into the impenetrable gorges to the west, these early Buddhists would, according to Staal, have concluded that this river was the source of the Ganges. They would, therefore, have regarded Rakas Tal, appar- ently the fountainhead of the great river of India, as highly auspicious. Later Buddhist travellers entered the Kailas region from Kashmir in the west, and as Kashmiri Mahayana Buddhists they held a less Indo-centric perspective on the world and would thus have been more concerned with the course of the Indus, the great river of Central Asia. Problematising this conclusion however, is that the most likely approach from the south was via Badrinath over the Mana-la, in which case such travellers would already have reached what they understood to be the Ganges source, and are thus unlikely to have confused it with the Sut- lej. But Staal also raises the issue of the inauspicious status now attached to Rakas Tal, a place few pilgrims visit. While its Sanskritic name means ‘Devils Lake’,suggesting its inauspiciousness in the Indian understanding, the origin of this status is uncertain. Staal concludes, although on little apparent evidence, that it is an association of comparatively recent origin, even as late as the seventeenth century. He suggests it derives from later Buddhists adopting the view of Manasarovar as a more auspicious lake from Saivite Hindus, although

83 Bedi & Swamy (1984: 5). They quote unnamed “ancient accounts” for this, but apparently rely on Pranavananda, whom they consulted prior to their visit to Kailas. Other commen- tators similarly appear to rely on Pranavananda in this regard; see for example, Allen (1982: 26); Johnson & Moran (1989: 41). 84 Lake Anotatta is located on Meru by Mabbett (1983: 66, 69); also see Obeyesekere (1987: 334). For an illustration and discussion of Lake Anotatta and its rivers, see Harley and Woodward (1987: 732–733). 85 See Staal (1990: 278–279). 86 chapter 3 there is no evidence for this and he does not venture to suggest how they formed that opinion.86 The only obvious geographical feature which might account for its malign associations is that the orthodox Hindu pilgrim who faces Kailas from a posi- tion south of the lakes has Manasarovar on his right and Rakas Tal on his ill- omened or impure, left. That is not, however, an understanding with obvious Indian pilgrimage site parallels, nor would it necessarily be the understand- ing of Tantric practictioners. Rakas Tal is known in mythology as the site where Ravana, the demon-king of Lanka in the Ramayana, carried out penance to pro- pitiate Śiva, and Rakas Tal is consequently also known as Ravana Sarovar.87 But given that in Hindu mythology Ravana is the half-brother of Kubera, ruler of Kailas in the Epics, his being allocated Rakas Tal is consistent, and provides a basis for the development of an understanding of it as a place of demons, and hence a sinister locale. Yet even that explanation is problematic, for a place associated with demons would be expected to attract renunciates such as those following cremation- ground practices, who were eager to confront their powers, and Rakas Tal is not renowned for attracting such renunciates, although the small islands on the lake are recorded as housing Buddhist monastics in retreat. Thus its inauspi- ciousness and overall status remains inexplicable, although we might consider that it relates to the failure of actual geography to manifest the ideal divine dyad of heavenly mountain-lake. Seekers after that ideal were thus required to account for the presence of an ‘extra’ lake in the heavenly realm, and deemed it demonic. Yet Staal’s article reminds us of another important point; that the focus of early pilgrims to the region was on a lake, be it Manasarovar or Rakas Tal, rather than the mountain, Kailas. Given that Vedic and later Indic culture revered river sources above mountains, and that Vedic and Brahmanical sources hint at a knowledge of the headwaters of the Indus but do not mention Kailas, the Pali sources suggest that this understanding of sacred water sources contin- ued at the time the Buddhist accounts were recorded. It was known that the Himalayas sheltered a majestic lake, a sacred place which was believed to be the source of the great rivers of Indian civilisation. But Kailas was not then regarded as a mountain of any particular sanctity. Staal also implies that Indic travellers to the Kailas region came in “waves”, interrupted by intervals when the region regained its isolation (and hence its

86 Staal (1990: 275–291). 87 Pranavananda (1984: 17). the puranic kailas 87 mystique). The “wave” theory has much to commend it, for we can accept the probability that eras of social breakdown occurred during which travel to the region became too difficult or dangerous to undertake. Certain routes to Kailas through the mountains could well have been forgotten by outsiders and rediscovered later, with layers of knowledge settling and then being readjusted and reinvigorated by each new wave. We must conclude that the textual lake Manasa/Mada/Bindusara/, like Kai- las, Meru, and even the myriad of rivers mentioned in these sources, was an imagined place, situated where the believer envisaged it to be. Its location was not fixed, but applied to the appropriate place depending on context. As one observer (favouring old English spellings), noted in 1801; “The few Hindus, who live towards the Indus, infift that the lake near Bamyan [in Afghanistan] is the real and original Manfarovara[sic].”88

Poetry

The Kailas of the great Epics, an imagined heavenly realm in the upper reaches of the Himalayas, seems to immediately predate the emergence of Kailas as a common and even prominent feature of a wider Indic culture, though chronol- ogy here is complex. But prior to the Epics there are only scattered indications of a conception of the upper Himalayas, while after the Epics Kailas emerges as a concept that if not particularly prominent is none-the-less manifest in a num- ber of cultural fields. We might locate this emergence to the 4/5th centuries ce, which is broadly consistent with the dating for its slightly later well-known appearances in architectural and poetic form and the still later emergence of more complex mythologies in the Puranas. There is no obvious explanation for this phenomena, no sources that shed light on process. But the tentative chronology of textual references to Kailas that we have advanced allows another conclusion, that the ideal of Kailas predated its identification with a geographical reality. That is, that a heav- enly Kailas was imagined by Vedic rsis and their renunciate successors who practiced in the Himalayas. No specific place was meant, although specific places may have impacted upon the construction. That ideal Kailas, the heav- enly model, was then applied in or envisaged to be manifest in, many differ- ent realms of culture and society, ultimately becoming a cultural reference point.

88 Wilford (1801: 491). Perhaps Band-i-Amir (which has 5 spectacular lakes) is meant? 88 chapter 3

It seems clear that it was the imagined realm that the great Sanskrit poet Kalidasa drew on for his famous poem “The Cloud-Messenger” (Meghadūta).89 Kalidasa’s dates are unclear, but he probably wrote in the immediate post-Epic height of Gupta court culture, around the late 4th—early 5th century ce.90 The Cloud Messenger tells of a yakśa, a divine attendant on Kubera in his Himalayan home, who is exiled to the Vindhyas for a year. The lonely yaksa calls upon a cloud to bear a message of fidelity back to his new bride, who awaits his return. Images from the Epic Kailas flow through the lines:

Bring her my message to the Yaksha city, Rich-gardened Alaka, where radiance bright From Shiva’s crescent bathes the palaces in light. … The swans who long for the Himalayan lake Will be thy comrades to Kailasa’s peaks, With juicy bits of lotus-fibre in their beaks. … Fly then where Ganges o’er the king of mountains Falls like a flight of stairs from heaven let down … As thy dark shadows with her whiteness blend– Would be what Jumna’s waters at Prayaga lend. Her birth-place is Himalaya’s rocky crest Whereon the scent of musk is never lost, … Seek then Kailasa’s hospitable care, … Drink where the golden lotus dots the lake; … Then, in familiar Alaka find rest, Down whom the Ganges’ silken river swirls, …91

89 The idea of Kailas as a poetic model seems consistent with Basham’s observation re “the very numerous and universally accepted stock epithets” in Sanskrit poetry, which, “say far more than their bare meaning and induce a whole series of emotions by a single brief verse.”; see Basham (1954: 419). 90 The prominent Gupta Emperor Harsha (r. 606–647: Lorenzen 1993: 212) was from Garwhal and thus we might speculate that a court poet would cite that region, although Kalidasa is probably earlier. 91 Kalidasa, The Cloud Messenger, translated by Arthur W. Ryder, modified by Antihubris the puranic kailas 89

This is the Kailas of the Epics, the home of Kubera, with his divine city Alakā, the Kailas peak only briefly mentioned, vaguely located north of Prayag and near the Ganges source, near an unidentified lake with its golden lotuses. This is a poetic vision of a heavenly abode.92 Even its tantalising glimpses of reality, its swans and its scent of musk-deer, are from the pen of the poet not the geographer. This is Kailas as a growing legend, not a geographical reality. But the great popularity of Kalidasa’s verses and its visionary implication ensured its lasting cultural impact.

Kailas Temples

The Kailas phenomena was part of a wider tendency aimed at replicating the model of the divine order in the earthly realm.93 This tendency (one of many such propensities within Indic society), followed certain cultural logics. It was understood that the cosmic order (ṛta) was maintained by each individual performing the role appropriate to their caste status (their dharma). It was also understood that the fully enlightened individual could control, and even become, a deity and that deities could appear in earthly forms. The distinction between the realm of deities and the realm of men thus became a matter of degree. With kings claiming sacral status, and renouncers recognised as capa- ble of attaining it, the heavenly mountain at the centre of the universe was not, conceptually, unattainable. It could be visualised through meditation or other practices and even experienced by certain cultural Heroes. This, concep- tually, meant that the divisions between worlds could be dissolved, and that the attributes of heavenly realms could be reproduced on earth. That tendency also involved the imaginal development of elaborate icono- graphies and mythological biographies for the pantheon of Sanskritic deities and their promulgation in art and literature. From around the 2nd century bce, with the rise of sects and fixed structures for worship, the demand for artis- tic expressions of the deities must have greatly increased. Temples, with their images of the deities, became increasingly associated with local and regional political authority. Their architecture came to be envisaged as replicating heav-

.com: from http://www.antihubris.com/books/india/cloudmes.php, consulted 17 April 2008. The poem was translated into Tibetan in the 14th century; Smith (2001: 193, n. 663 319, 370). 92 Granoff (1997: 176–184) observed that there were stock descriptions of heaven, which were drawn on by disparate authors. 93 On which see Miller (1985: 203). 90 chapter 3 enly vistas, with the temples as palaces for the deity’s manifestation on earth conceptually replicating the palace of the royal ruler, and vice-versa. Gods, rulers, and communities of worship were thus linked with the divine order.94 In seeking to portray heavenly vistas, the idea of a divine mountain—a Meru, a Kailas, or a Himavat—became a central conceptual device, and it emerged in architectural form with the earliest extant architectural text, the Bṛhat-saṃhitā (c. 587ce).95 This refers to temple models termed Himavat, Meru and Kailas. Notable Kailas temples were those at Ajanta and Ellora in Maharashtra. While earlier dates have been proposed for Ajanta, both seem to represent Pallava dynasty styles, and the great Kailāsanātha rock-cut temple at Ellora is convincingly datable to the mid-late 8th century.96 Their location also suggests their divine inspiration. A temple modelled on a mountain in a heavenly location was surely a more impressive conception than one modelled on a mountain in a known earthly region. Had reality been intended more localised regional sacred mountains would presumably have been used. What was being implied was an earthly manifestation of the heavenly realm. This use of Himalayan mountain toponyms for temple types appears to predate any reliable textual exposition of the attributes of these sites. In other words, the temple types are ideal models which precede the actual knowledge of the sites (just as yoginī temples represented ideals of flying yoginis97). They draw on cosmology rather than empirical knowledge and are thus signifiers of the idealisation of Kailas and other Himalayan sites rather than of their actuality. As depictions of Kailas-Manasarovar concerned a heavenly, not an earthly, realm, there was no cohesive development of the understanding of Kailas. Ele- ments of the construction were sometimes unique to a particular tradition and

94 On the temple as the actual location of the god, see Granoff (1997: 170–193, esp., 175, 184, 187); also see Eck (1985). 95 Inden (1985: 54–60); also see Inden (2000: 86–87). 96 Malandra (1993: 5–12) concludes that its development was in phases; Hindu from 550– 600ce, Buddhist from 600–730, and a Hindu/Jain phase, from 730–950. The Kailas temple was built under Kṛṣṇāraja the 1st, who reigned from c. 757–772ce. Buddhist inscriptions date Ajanta to 460–486ce. Also see, Havell (1924: 25–27, inc., plate viii a & B) who makes the novel argument, on the basis of the resemblance of the side elevation of the temple to that of Kailas mountain, that the designer of Ellora must have seen the mountain. We may note the rock-cut temples of Bamian also date to the 5–8th centuries. Also see Bäumer (2010). 97 Yogini temples were roofless, to allowing access to their flying consorts; see White (1998: 198). the puranic kailas 91 sometimes shared or present as part of a wider cultural conceptual device. But terms that carried cultural implications could also have wider applica- tion within society and pass between completely different categories. Sacrality attaching to a term made it a cultural device, just as Kailas may be used today as a personal name, one with subliminal associations, and deliberate correspon- dences existed or were created through the use of such terms. A “free-floating”98 term might become fixed as a category or as a symbol; Kailas was one such term. Partly because it was neither precisely defined or geographically located, it most commonly became a metaphor for strength, purity, and heavenly man- ifestation in the earthly realms. Just as such a metaphor could be a literary device in the hands of a poet such as Kalidasa, so too could Kailas be a suitable model for temple architecture or an appropriate subject for a classical narra- tive. Such metaphors carried considerable meaning within the cultural system.

98 Von Falkenhausen (1995: 281, citing a 1931 article by Bernard Karlgren in the Bulletin of the Museum of Far Eastern Antiquities) uses the term “free” textual material, a useful concept for discussing toponyms, as well as intertextuality. chapter 4 A Tantric Kailas: The Alchemical Trail

The Medieval Context

While a Sanskritic Kailas-Manasarovar emerged as a formulated ideal around the mid-first millennium ce., the actual site does not properly emerge into the Indic historical record until well into the second millennium. Even the Tibetan sources are reticent as to whether it was visited by Indic followers of the Brah- manical traditions before that time. There are references to tīrthikas (Tib: mu stegs pa), a term often used to refer to the Bönpo, but also to Hindus.1 Yet it is notable that the Guge-Purang history (see Chapter 11), which refers to Buddhist practitioners at Kailas-Manasarovar in the 11th–15th centuries, do not mention any non-Buddhist visitors from India. Indeed, their only reference to Hindu pil- grims is a rather cryptic legend—which takes several forms—concerning seven “yogins” refused hospitality who turned into wolves before disappearing.2 There are no clear historical references to Hindus reaching the Kailas region from India until the 17th century when the kings of both Garwhal and Kumaon dispatched forces across the Himalayan watershed, notably in 1670, when King Baz Chand of the Almora-centred Chand dynasty sent forces as far as Taklakot (Purang), purportedly to protect pilgrims.3 As will be seen, this king also patro- nised Badrinath temple,4 and is credited with the Sanskritisation of the local

1 Das (1983: 967) defines tīrthika as, “a heretic … or one indulging in pilgrimage … a non- Buddhist.” Davidson (2005: 49) states that the term “may indicate Śaiva or Śākta yogins, Brah- mans, or others outside a Buddhist affiliation” and sometimes (Davidson 2005: 137) clearly indicates Brahmanical traditions/Hindus. Its use is attested in the Dunhuang sources, see McKeown (2010: 53, 247); and the term is also used for Muslims, see Kapstein (2011: 347). Martin (2001: 36) cites a life of Padmasambhava containing a list of ten countries (in each of which Padmasambhava stayed 200 years). In addition to India, China, Magadha, etc., it includes a “tīrthika country.” Again, as Martin (2001: 176, 183–190) indicates, Buddhist texts sometimes list Bön and tīrthika separately. 2 See for example, Vitali (1996: n. 491 318); also see Cutler (1996: 211–212). 3 Walton (1911); Pande (1993: 252, 259); also see Atkinson (1883: 67–71). I am advised by Roberto Vitali that he has not sighted any mention of the 1670 incursion in Tibetan sources. These incursions were obviously military rather than religious in intent, and the pilgrims needing protection may have been to Badrinath rather than to Kailas-Manasarovar so these are not proof of an Indic Kailas pilgrimage at that time. 4 Handa (2002: 82–84).

© koninklijke brill nv, leiden, 2015 | doi: 10.1163/9789004306189_006 a tantric kailas 93 cross-border trading communities (‘Bhotias’/‘Jads’).5 Thus his forces’ incursion over the watershed was probably part of Sanskritising the region, and strength- ening his northern frontier. But if we cannot conclusively date a Hindu presence at Kailas-Manasarovar even to the 17th century, earlier visits seems certain. Travel there by renunciates probably predates this by four or five centuries, or much earlier if we accept that early sources reflect some knowledge from actual journeys to the region. Peripatetic groups such as the Pasupatas travelled widely in the Himalayas and as there were also early sects that focused pilgrimage on the Ganges,6 that they at least reached Badrinath is probable. Other routes to Tibet were also open. A Nepali inscription from 695ce., records the royal grant of a village to support Pasupatas en route to Tibet.7 But there is no evidence that they were bound for Kailas; Lhasa is a more likely destination.8 From the 18th century onwards however, there is evidence of Hindu renun- ciates travelling to Kailas-Manasarovar, and in this chapter we shall consider why certain renunciate groups were attracted to the site.

Tantra, Renunciates, and Their Communities

We have seen that renunciates as a category were prominent from the earliest period of recorded Indic history, and that conceptually they complemented their fellow Brahman ritualists, with the fulcrum of division between the two being their pure/impure status. The renunciates, equipped to travel beyond the purified norms and realms of Indo-European society, acted as agents of that culture in a two-way process of cultural encounter and expansion. The strictures against travel to heavenly realms such as the Himalayas, so prominent in the Epics and Puranas, did not apply to them for they were understood to be equipped with the tapas-derived magical powers that qualified them to travel there. During the first millennium ce., new ritual technologies, or systemised sets of disparate existing technologies, became part of the armoury of the renunci-

5 Brown (1984: 9). 6 On the Gangayatrahs, see Habibullah (1976: 434). Habibullah also notes the existence of a Hindu temple at Bamian (Afghanistan) which was visited by renunciates; (cf; n. 88 87, this work). 7 Davidson (2005: 132). 8 See Dargyay (1981) on Yarlung dynasty support for “yogis” etc, although these seem to project a Buddhist identity. 94 chapter 4 ates.9 These were articulated in a particular genre of texts, the Tantras. Central to this category was the developed understanding that the individual could become their chosen deity, or at least acquire the powers of that deity. Tantra was a means to achieve this state through self-identification with the deity by a highly ritualised series of practices potentially applicable to any deity.10 Tantra thus transcended religious boundaries such as those between Hindu, Buddhist, Jain, or Bönpo, and, “came to pervade almost all areas of Indian religion.”11 Our concern here is not with properly defining or exploring the complex his- tory of Tantra, subjects on which there is now a considerable literature and a certain degree of consensus.12 But we must note that aspects of Tantric prac- tice violated social order, not least Brahmanical purity norms. Most obviously transgressive were practices such as ritual killing and sexual rites which were explicitly ordained in numerous texts.13 Tantra has consequently been viewed as unorthodox both in emic and etic perspectives, and its teachings and prac- tices have frequently been subject to a process of censorship and even suppres- sion by various religious and secular authorities. But it was not denied within Indic society that Tantra was a valid means to attaining spiritual liberation (mokśa). Reformation was thus sought rather than suppression, particularly the domestication of transgressive rituals within a revised system, rendering, or attempting to classify for example, erotic ele- ments as symbolic or entirely conjugal. As Alexis Sanderson states, the “… gen- eral trend—and this was also the case in Tantric Buddhism—was to purify the rites by taking in everything except the elements of impurity.”14

9 As with earlier texts in the Atharvavedic lineage, these included protective rituals relevant to renunciates travelling beyond society; see Dyczkowski (1988: 39–40) regarding the (lost) Gāruḍa Tantras, which dealt with magical remedies for snakebite and poison (cf: the hymns of Garutman in the av.). 10 The origins of this concept are uncertain, see Samuel (2008: 238) on the possibility that vratya ritual involved self-identification with the deity. 11 Sanderson (1988: 130). 12 On Tantric historiography, see Wedemeyer (2001); also see Urban (2003). On Tantra, see for example, the works of Ronald M. Davidson, Mark S.G. Dyczkowski, Alexis Sanderson, David Gordon White, and Teen Goudrian cited in the bibliography. 13 Given the Tantric initiate sought identification with the deity, might one conceptual agent behind the sexual rites be the practitioner’s question of how to make love to the Goddess/consort? We might note in this regard Arthaśāstra 4.13.41 (Rangarajan 1992: 485), where a penalty is imposed on those having sex with an image of the goddess. This presumably means that was known to occur, a point White (2001: 17) seems to have observed. 14 Sanderson (1988: 129). a tantric kailas 95

A major stimulus to the survival and development of Tantra was that its prac- tices, with their echoes of Atharvavedic magic,15 came to be central to the ritual armoury of the royal purohita—almost invariably an Atharvavedic Brahman. In the royal setting these Brahmans could deploy the full Tantric armoury in sup- port of the king, and might provide Tantric initiation to equip him to deal with impurity (for like the Vedic sacrificer, the Hindu king was inherently involved in killing). Certainly the power of killing through ritual magic, or of concealment or other such classic rites of the Atharvavedic tradition, were of considerable appeal to Indic rulers, and with the increasing systemisation of Tantra from around the seventh century ce.,16 Tantric concepts came to shape the articula- tion of the conceptual basis of society. While the multitude of forms which Tantra takes make its classification problematic, what is characteristic is that the (Hindu) Tantric alternative was within the Brahmanical traditions.17 Its apparently subversive elements were ultimately a means of strengthening the charismatic power of the Brahmani- cal caste. Whereas temple Brahmans used Tantric rituals and understandings within Brahmanical purity norms to cater for orthodox household clients, the renunciates used impure ritual forms as a means of confronting the impure world in which so many of them operated—a world of tribal, lower, and out- caste groups. Yet when we identify the leaders of communities that formed around charis- matic individual renunciates, particularly at sacred sites, the question of whether they were or are Tantrics is not necessarily significant. Tantra in this context is simply a weapon in the renunciate’s armoury. By no means all renun- ciates used this weapon, although they were popularly understood to possess it.18 The nature of such rites, however, ensured their prominence and both a

15 The issue of continuity between Atharvaveda and Tantra remains controversial, see how- ever, Sanderson (2007). 16 Many of the strands of belief and practice that make up Tantra can be traced to the Vedic period and even earlier. I would argue that once the belief that the gods can be controlled by the sacrifice is accepted within Indic society the conceptual basis for Tantra exists. But the date by which Tantra can be said to be a fully-formed system is much later. That system as expressed in texts only emerges in the second half of the first millennium ce., and thus as Davidson (2002: 117) concludes, “[i]t is only in the second half of the seventh century that the definitive esoteric system emerges.” 17 Evidence for this is the use of Sanskrit and the Brahmanical authors of many such texts. An unpublished paper by Mayer (1990), makes a convincing argument for the Brahmanical origins of Tantra. 18 On the Tantric ritual consumption of sexual fluids, which White places at the centre of South Asian Tantric practice, see White (2003). 96 chapter 4 specialised and a popular understanding of renunciates as performing their rites, “… in secluded places such as lonely forests, mountains, deserts, crema- tion grounds or sacred centres where adepts, male and female (siddhas and yoginīs), traditionally assemble.”19 We have seen that liminal spiritual figures associated with magical and transgressive rituals were characteristic not just of Indic but of Pan-Asian soci- ety. Tantra (which is indisputably of Indic origin20), provided Indic renouncers with a powerful new weapon of considerable appeal to the wider world, and it was rapidly spread by them to central and southeast Asia. As noted, Tantra had considerable appeal to royal elites, and in the hands of a charismatic indi- vidual, Tantric rituals could appeal to all social groups, just as the Atharvavedic traditions that were at the roots of Tantra appealed across the social hierarchies to all who sought the power to transform the circumstances and relationships of their immediate world. But as will be seen, the Tantric worldview not only impacted on gender and society, but through its empowered formulation and visualisation of landscape it produced new codes of meaning and understand- ings of place.

Pithas

A distinct tradition of sacred sites—pīṭhas—developed within the Tantric tra- dition, alongside and occasionally overlapping the development of sacred sites expressed in the Puranas. The mythological foundation of the concept of the pīṭha (‘seat’; i.e., of the goddess: śākta), which was added later,21 is the well- known Puranic account (e.g.: Kālikā Purana, 16.18) of the grief-stricken Śiva car- rying the corpse of his wife Sati across Indian skies. As the decomposing parts of her body fell to earth, these formed the sacred places of the goddess. The pitha network was the central feature of Hindu Tantric sacred geography,but was also assimilated into both Puranic and Indo-Tibetan Buddhist thought22 (demon- strating the overlapping elements within the diverse Puranic and Tantric texts and traditions23).

19 Dyczkowski (1988: 59). 20 “The origins of Tantra are … Indian. All authentic Tantric lineages—of deities, scriptures, oral teachings, and teachers—claim to extend back to Indian sources.”; White (2001: 20). 21 The earliest texts to include this myth are the Kālikā Purāṇa (11/12th–14th c.) and Devīb- hagavata Purāṇa (13th/14th c.); Bakker and Entwhistle (1983: 33). 22 Dyczkowski (2004: 97); also see Davidson (2002: 206–211). 23 Thus the pitha scheme also appears in Puranic texts; see van Kooij (1972). a tantric kailas 97

In texts from the late first millennium the number of pithas was generally given as four. That numeration was shared for example, by the Hevajra Tantra24 and the Kālikā Purana, which both identify the sites as Jalandhara, Oddiyana, Karmarupa, and Purnagiri.25 But as we have seen with the model of four rivers, precise identification was a pliable concept reflecting different interest groups. Thus the Cakrasamvara Tantra lists the sites as Jalandhara, Oddiyana, Arbuda, and Pulliramalaya,26 and the related Samvarodaya Tantra names Jalandhara, Oddiyana, Arbuda, and Purnagiri.27 Though a hierarchy is not commonly sug- gested, other texts such as the Kulacūḍāmaṇi and Vāmakeśvara Tantras (from the 9th–11th century period at the earliest), indicate their eastern regional ori- entation by including Karmarupa as the supreme pitha of a quartet with Jaland- hara, O/Uddiyana, and Purnagiri.28 None of these early texts situate Kailas (or Manasarovar) as a pitha, further evidence of its then obscurity. Later texts enumerate a greatly expanded pitha network; 24 sites are recog- nised by the Kubjikāmata Tantra, the earliest manuscripts of which are from the 11th century,29 while the early/mid-16th century Jñānārṇava Tantra listed 50 pitha places. While there were occasional lists of 108 such sites, the concept reached its fully-developed form by the 17–18th century when the number was semi-stabilised at 51. These pithas seem to be a device for recognising and link- ing local sites, including many in tribal or peripheral areas, into a larger network of Tantric sacred geography. But no standard list developed, the texts contin- ued to represent regional perspectives on the identity of the sites (if indeed all 51 actually existed outside of visualisations, which is doubtful, for again there is a blurring of earthly and heavenly locations).30 Pitha toponyms include Kashmir (Kaśmīrā), and Nepal (Nepāla), regions rather than specific places, which may imply they were only vaguely known to the authors or that they were considered vast sacred realms. Again the texts represent distinct traditions and bodies of knowledge that would have been

24 Snellgrove (1976: 70). Bharati (1965: 88) dates this text to 690ce and states this is the earliest such list. 25 Discussing possible locations for this latter site, Dyczkowski (2004: 105) favours a moun- tain of this name in central India. 26 Sircar (1973: 12); Gray (2007: 67). Gray (2007: n. 22 332) identifies Arbuda as Rajastan’s Mt Abu. 27 Tsuda (1974: 272). 28 Finn (1986: 12, 118, 126–127). 29 Goudriaan and Schoterman (1988: 24; also see 126 which lists 24 pithas, apparently not including any Himalayan sites). 30 Sircar (1973: 3–4, 18–22). 98 chapter 4 of very limited circulation, perhaps restricted to initiates of a particular cult. Thus they seem intended for an internal audience, in contrast to mahatmyas intended for a general audience. Further evidence for this is that most identi- fiable sites were outside the boundaries of medieval Indic culture, located in tribal or other peripheral and ritually impure areas. They therefore provided centres for renunciates and their followers outside of settled Brahmanical soci- ety. It is in this context that we must assess the addition to the pitha lists of sites in the Himalayas. The Kubjikamata Tantra has Jalandhara and Jvalamukhi as separate sites, and it also includes Prayag and Badari (Badrinath?). The Jnanarnava Tantra includes in its 50 pitha sites not only Prayag (although not Badari), but Kedara, Merugiri, and Kailasa (although not Manasarovar). If we accept the dating of this text as not later than the mid-16th century we have evidence for a distinct Tantric conceptualisation of the upper Himalayan region existing at least a century prior to Baz Chand’s initiatives there.31 (It is not, however, conclusively a representation of the Kailas-Manasarovar region, for the four named sites all exist as toponyms in the modern upper Gangetic regions, as will be seen.) We thus have another layer of sacred geography encompassing the Himalay- as. To the Epic and Puranic geographies, and those of specific traditions such as the Śaivites or Buddhists, we may add Tantric sacred space; the layers separate, albeit overlapping. What we do not have is any change in the understanding that these were sites restricted to advanced renunciates for Tantric texts also clearly state this to be the case.32 The pithas were gathering places for renunci- ates, understood within each tradition honouring the site as places where the deities and their worlds were manifest (at least to those of enlightened mind). They were places ideal for spiritual practice, and, no doubt places where the necessities of the renunciates’ life and practice were available.33

31 Dyczkowski (2004: 158–160) maps 50 pitha sites from the Manthānabhairava Tantra and its associated Ambāmatasaṁhitā, as well as the Yoginīhṛdaya. These include Kailas but not Manasarovar as pitha sites. If my reading of Dyczkwoski is correct, the inclusion of Kailas is late. A very modern addition to the pitha network in the Himalayas—testifying to the dynamic and revelatory nature of such identifications, is that by a renunciate resident at Kailas in the 1990s. He identifies “Dolma devi”, a large rock on the Drolma la (the highest point of the Kailas circuit), as one of 52 Shakti pithas (apparently where Parvati’s right thigh fell); see Giri (2000: 128). 32 Zangpo (2001: 30), citing the Sakya Pandita, “A Thorough Delineation of the Three Vows”. 33 A modern parallel might be the ‘Hippy Trail’ network of sites such as Rishikesh, Kath- a tantric kailas 99

A pitha site was the centre of the sacred realm of the deity. While in the primary myth this was the wife of Śiva, the model could be applied to any goddess, with the landscape understood as her body and her physical features perceived to be manifest there. In the Tantric practice of identification with the deity, the conceptualisation of the body by the fully-realized Tantric was of correspondence with that of the deity. Thus the sacred geography of the pitha, the body of the goddess, could be inscribed on the body of the Tantric. Each feature of the goddess’s body, and thus of the sacred landscape of her pitha, could be visualised within the body of the practitioner.34 Mark Dyczkowski has stated that by the 11th or 12th century the development of this technique meant that Tantric pilgrimage to the pitha places became “redundant.”35 While that may be theoretically correct, in reality, being social actors the renunciates continued to travel to these sites.36 Itinerancy was very much a part of their identity, and as Dyczkowski notes, the texts also stated that these sites, or “power places” as Keith Dowman termed them, were ideal for the performance of Tantric rituals.37 The inscribing of the pitha on the body of the Tantric practitioner led not to such places becoming redundant, but rather contributed to their infinite replication, with the practitioner able to reproduce that internalised sacred geography onto their vision of other land- scapes. In regard to Kailas-Manasarovar, as will be seen, this concept may have been most significant within the Buddhist tradition, allowing the installation of a Buddhist deity onto the sacred landscape. But it is not clear that this pro- cess was common among Brahmanical renouncers. Although the Jnanarnava Tantra indicates that at least the upper Himalayan realm was within Tantric conceptions of sacred space it is not prominent in known texts until a very late exception discussed in Chapter 6. Kailas-Manasarovar seems late and periph- eral in pitha enumerations. Thus we might consider alternative motivations for Hindu Tantrics crossing the Himalayas to Kailas-Manasarovar.

mandhu, Varanasi, Goa, etc, which is similarly a circuit dictated by such factors as eco- nomics, climate, and the timing of particular festivals. The central sites offer the neces- sities of the Trail and its culture, consumption, and so on, while guidebooks ritualise behavior. 34 Dyczkowski (2004: 141 & passim). 35 Dyczkowski (2004: 117; also see, 146 where he reiterates the latter). Pal (1988: 1049–1050) similarly argues these sites “exist only in the mind and the body of the tantric”, and thus cannot be identified with precise geographical locations. 36 For a discussion of actual and visionary journeys in the context of the Cakrasamvara Tantra, see Templeman (1994). 37 Dyczkowski (2004: 141). 100 chapter 4

An Alchemical Kailas

If early renunciates travelled to Kailas-Manasarovar their motives presumably included the search for the source of rivers or spiritual experiences and encoun- ters in mountains considered the home of the Vedic deities. But a different motivation is apparent for one group of renunciates who emerge as a distinct entity in texts dating from around the 13th century (although claiming earlier antecedents). Here we follow the work of David Gordon White, who has inves- tigated a particular strand of siddha practice, the alchemical path involving the ingestion of substances considered physically and/or mentally transformative. While Tantric systems are all traceable to Indic foundations, alchemical spiri- tual practice appears to originate in Persia,38 or Taoist China.39 Its passage to India, as with the transmission of Tantra to China, indicates some communica- tion with the esoteric practitioners of these lands. From around the 5th century ce, if not earlier, alchemical substances began to be drawn into the spiritual practice of certain Indic renunciates. White identifies three types of siddha order as developing from this period: Kaulas, seeking immortality through erotico-mystical practices; Rasas, or alchemists, seeking immortality through ingesting minerals equated to the sexual fluids of Śiva and his goddess consort; and the Nāths, whose hatha yoga projected these ideas onto the body. These were overlapping and even integrated paths, approaches rather than specific sects, and influenced both by “mystic and metaphysical speculation on the one hand and … developments within the medical schools on the other.”40 Closely associated with Tantric developments, what White calls “an alchem- ical synthesis” emerged in the 10th century in western India or Andhra through the interaction of Rasa and Nath siddhas with followers of the Kubjika tradi- tion, resulting in a specifically Indian Tantric alchemy.41 This ideology may have led renunciates to the Kailas-Manasarovar region, for it was a bountiful source of alchemical agents they could consume or trade. John Clarke has demonstrated the existence of renunciate trading networks throughout India and Tibet in the 18–19th centuries and perhaps earlier. The maths (temple/monastic centres) of north Indian renunciate sects were heavily

38 White (1996: 187) states that Persian traders to Chang-an were regarded as alchemists, “due to the magical powers and precious stones in which they traded.” 39 On which see, Strickman (2002). On alchemical practices by the Nath order in the context of Chamba, see Sharma (2009: 102–103). 40 See White (1996: esp., 55). 41 See White (1996: x, 2, 54, 76–77, 142–143, 176); also see Schoterman (1992: 313–326). a tantric kailas 101 involved in business and finance by his period, and travelling renunciates played an important part in these trading activities. The renunciates followed annual trading and pilgrimage cycles that coincided with fairs (melas) at sacred sites and with the accessible seasons on the high Himalayan passes. Among their cycles were routes across north India, through south India in March/April for the pearl fishing season, and others that reached through Nepal, the western Himalayas, and Tibet.42 While trading in items such as pearls would generally have required institu- tional investment, itinerant renunciates could also fund their individual trav- els in this manner. They generally carried small items of high value—gold, musk, and so on—that could be easily concealed and, while their religious sta- tus offered a certain protection from robbery, renunciates usually travelled in groups for added protection. One feature of their trade goods was that they were often items that could obtained for nothing in one region, but which were of great value in another. Water from the Ganges, particularly the source, is one obvious example, as is at least in the modern period, cannabis, and alchemical substances would have come into this category.43 Evidence for (Tibetan) renunciates visiting sacred mountain sites in search of these agents comes in Toni Huber’s definitive study of the pilgrimage to Tsari (“Pure Crystal Mountain”) in southeastern Tibet. Huber demonstrates how the Tibetan Tantric understanding of the sacred landscape itself as empowered means that the centre of the sacred site is perceived as the palace of the deity. There, “the vital bodily substances of the chosen meditational deities can be found and extracted from the landscape itself, as their bodies are the body of the summit.” Huber gives a number of examples of these “empowered” substances at Tsari, along with a variety of herbs and plants are a thick white matter found floating in a lake, which substance is considered the semen of the presiding deity of the site, and a type of red ochre suspended in a lake and considered to be the menstrual blood of the deity’s consort (which “can be consumed to obtain direct enlightenment”).44

42 Clarke (1998: 52–70). 43 Tapovan (2001: 130) notes pilgrims on return to Almora selling plants taken from the Indo-Tibetan passes. 44 Huber (1999: esp., 94–100); also see Huber (1999a: n. 11 99); also see van Kooij (1972: 25–27) concerning the Karmarupa pitha site where reddish water derived from red arsenic is attributed magical visionary power. (van Kooij’s study of an 11th century work centralising this site contains a number of interesting parallels with the textual strategies employed to promote Kailas.) Items taken from sacred sites are not necessarily specifically ‘Tantric’, however, see Buffetrille (2014) and Huber (1999: 100, 114–117, 249–250) re bamboo. 102 chapter 4

There are a number of other such Himalayan sites accessible from India. Mercury, the alchemical substance par excellence, seen in this context as divine semen, is found in northern Kashmir.45 Mica “the sexual fluid of the goddess”, and sulphur, “her menstrual blood”,46 along with other alchemical agents such as calcium carbonate and ochre (sindhura), as well as gold, borax, arsenic sul- phates, and various medicinal stones, are all found on the Kailas-Manasarovar pilgrimage route. Tirthapuri (Tib: Pretapuri), around 38 miles (60 kilometres) northwest of Lake Manasarovar on the Sutlej route is a site of sulphur springs and is partic- ularly rich in alchemical substances.47 Of Tirthapuri, Pranavananda states that it, “is believed both by Indians as well as Tibetans that the pilgrimage to Kailas is incomplete or does not bear full fruit unless one visits Tirthapuri also”,48 and most Indian pilgrims do seem to have visited this site. Despite certain basic mythological structures, its sanctity is otherwise difficult to account for except as a site for the collection of empowered substances.49 As a Tibetan guide-book to the region states, “… as for the central power place there, the hot springs itself is taken to be the flow of the white yogic awakening fluid.”50 Observers note that the transactional collection of these substances by pil- grims is common, with offerings left in return. At Tsari, however, Huber records that the collection of empowered substances in the inner heart of the sacred site was restricted to certain recognised Tantric practitioners (which may have been the case at Kailas until the modern period). Their collection of these sub- stances was, however, different to that of the ordinary pilgrim, a distinction

45 White (1996: 65–66). On the alchemical mythology and traditions concerning mercury, sulphur, etc., see White (1997: 73–80). 46 White (1996: 5). 47 Pranavananda (1983: 49–51, 103). On mercury and sulphur in the western Himalayan and Manimahesh context, see Sharma (2009: 117–118). 48 Pranavananda (1983: 103–104). 49 The literature on Tirthapuri is limited and I have not visited the site; I am grateful to Felix Dawe, a Buddhist practitioner, for providing me with an account of “The Tantric land of Trita Puri [sic]”: email of 16 November 2005. This states that it is the site of worship of Vajrayogini/Dorje Pagmo, divine consort of Cakrasamvara/Demchok, and the site of a number of meditation caves associated with Padmasambhava and other seminal Tibetan Buddhist figures including Götsangpa, the ‘opener’ of the Kailas pilgrimage in Tibetan Buddhist understanding. The account states that in the cave of Vajrayogini, “pilgrims collect a type of soil called Jang-Sem-Karpo” as prasad; cf., Huber (1999: 96) “changsem marpo”, the red ochre tinted lake water. Also see, Pranavananda (1983: 103–104). 50 Huber & Rigzin (1999: 139). a tantric kailas 103

Huber attributes to “the yogin’s practice of ‘essence extraction’ (chulen) and the use of special preparations drawing on alchemical theories.”51 As well as alchemical substances at Tirthapuri, the shores of Manasarovar offered such materials to the practitioner. We may, therefore, envisage a circuit via Manasarovar and Tirthapuri being undertaken by Tantric renunciates col- lecting empowered substances. From there they headed westwards through the Chamba valley to Jvalamukhi, near Jalandhara in present-day Kangra (h.p.).52 As will be seen, this route was taken by a number of renowned Tibetan renunci- ates in the medieval period and while we do not have specific records of similar journeys by Indic renunciates, there is the apparently incongruous reference to this circuit in the Manasakhanda, the mahatmya of Kailas-Manasarovar (dis- cussed in Chapter 6). That text lays out the route from Almora to the sacred site, and the description of the return journey indicates in close geographical detail how the pilgrim should return south from Rakas Tal to Badrinath. But it then states, “[t]hence to Jwála-tírtha [i.e.: Jvalamukhi], where he should worship the sacred fire and bathe in the Padmávati. Thus is the pilgrimage completed.”53 The reference is otherwise inexplicable, with no sites mentioned between the Kailas-Manasarovar region and the otherwise unmentioned Jvalamukhi in Kangra, which is some 300 miles (500 kilometres) to the west and totally outside the region with which the text is concerned. There are no mythological or obvious historical links between the two sites, and—if our source is reliable (see Chapter 6)—the mahatmya thus seems to be encoding a different body

51 Huber (1999: 97–98). 52 The present day Jvālāmukhi (mukhi = ‘mouth’) temple complex is 24 miles (38 kilometres) from Kangra in the Beas valley (h.p.) The temple complex includes a shrine to Goraknath. For local mythology see Jerath (2001: 22–27). Jvālā/Jvālāmukhi is a local Kashmiri goddess and a pitha site in that context; Sanderson (2003–2004: n. 29 361). Jvalamukhi is also cited as a Tibetan goddess ruling over cholera; Chhodak (1978: 32). While the uncritical acceptance of early toponyms as identifiable with modern day sites is criticised here, the distinct nature of the place, with its flames emerging from the rocks, would suggest that early references may be safely identified with the present-day site. This is also the tentative conclusion of Barkhuis (1983: 58–69). That conclusion does not however, preclude the use of the toponym for other sites (not least at Mukhtinath, Dullu, and Baku, sites of a similar phenomena). On Jvalamukhi, see Shastri (2009; who notes [11] that it is also a Cakrasamvara site); also see Hutchinson & Vogel (2000: esp., 99–109). On Dullu, see Lecomte-Tilouine (2009). On Baku, see Stewart (1897). The toponym of the nearby town of Jalandhara is often (and for our purposes generally) interchangeable with Jvalamukhi. In some sources they are separate, in others identical. 53 The Manasakhanda, as cited in Atkinson (1884: 45). It must be borne in mind that we cite an Anglicised précis, not a translation (on which see Chapter 6). 104 chapter 4 of knowledge to that at the heart of the text. Intended as a guide for pilgrims, it apparently acknowledges the travels of the alchemical traders and ritualists existing at a more advanced spiritual level. Jvalamukhi was and remains an important site for renunciates, and it seems that many chose to travel there via Lahaul and Chamba rather than returning to Badrinath, from where the journey to Kangra may have been more difficult, particularly in times of political instability. Chamba, with its abundant supplies of lime, mica, gold, and gypsum,54 had its own attractions to the alchemist. Nath siddhas were prominent there,55 and more so in Jvalamukhi.56 Thus, not least because of the renunciates’ need for patronage to travel outside the Indic world, the most probable impetus to renunciate travel to the Kailas-Manasarovar region was the demand for alchemical agents that arose by, if not earlier than, the 13th century ce. As will be seen, this was a period of major developments in the Tibetan Buddhist understanding of the region, and that seems likely to have drawn a response that involved the identification of the site as sacred within the Indic traditions. We shall therefore, after discussing the renunciates’ use of psychotropic agents and aesthetic perceptions, look more closely at the Nath traditions that were at the forefront of these developments.

A Psychotropic Lineage

Tantra represents a body of ritual and visionary means by which the individual may obtain the powers of the deity worshipped. Included in those means are approaches antithetical to Western understandings of religious paths; not only sexual but also psychotropic substances are among them. As one Tantra states;

How can there be sacred knowledge ( jñāna) for [a] man who is with- out intoxication; or how can there be worldly knowledge (vijñāna) for him[?]57

Among alchemical agents were various plants used in Tantric practice as well as in the related medical traditions (suggesting their shared Atharvavedic foun-

54 Rose & Hutchinson (1996: 16–32, 240–241). 55 Sharma & Sankhyan (1996: 46) note a Nath Bairagi gotra at “Bharadwaj” (Bhadrawah?— see Chapter 7). 56 Sharma & Sankhyan (1996: 668); also see White (1996: 193). On the mahatmya of Jvala- mukhi, see Sharma (1995). 57 Tsuda (1974: 315). a tantric kailas 105 dations). Light-weight, free to obtain, and easily transportable, these plants were another natural product which renunciates could collect to use and trade. Pranavananda refers to several such botanical items at Kailas-Manasarovar, including cooking spices and plants used in incense-making, along with thuma (described as an “excellent aphrodisiac”), and at a village just south of the Dharma-la, ephedra vulgaris.58 Pranavananda identifies this ephedra by the name soma, that given to the divine hallucinogenic of the Vedas. As noted, the identity of soma has long been the subject of academic debate and recent schol- arship has tended to favour Pranavananda’s identification, one still extant in the vernacular in some regions of India.59 In his discussion of the collection of herbs at Tsari, Huber draws attention to a particular class of plants sought there by Tantric practitioners, those that “either possess or are identified with the paranormal powers of the divine res- idents of the mountain.” Foremost among them at Tsari is “tsa ludud dorje (‘adamantine serpent deity-demon herb’).”60 While I have seen no reference to this occurring at Kailas-Manasarovar, another strong candidate for identifica- tion with soma, Syrian ru (Peganum harmala), is found there, as is henbane (Hyoscyamus niger), a well known psychotropic plant used in witchcraft in medieval Europe.61

58 Pranavananda (1983: 44–45, 148, 199). 59 A segment of the admirable History of India television series by Michael Woods includes his obtaining “soma” (ephedra) from a herbalist in north India. 60 Huber (1999: 98–100); the latin name is not given, see perhaps, Ricard (1994: 246) concern- ing “Black Naga’s Devil’s herb (Tib: klu bdud nag po)”; “This herb cures leprosy and other ailments, and brings forth both ordinary and extraordinary siddhis.” Also see Ricard (1994: 268–269) on the bonnet bell-flower (Codonopsis), a “creeping plant with grey-blue flow- ers and an unpleasant smell … said to cure all diseases, especially leprosy and epilepsy; to enable one to fly … and walk on water; and to bring forth all ordinary and extraordinary siddhis. It is used in the preparation of a sacramental substance (…) called the “rainbow light pill” (’ja ’od ril bu) the mere taste of which liberates one from rebirth in the three lower realms of samsara.” 61 On Peganum harmala at Kailas, personal information courtesy of Carol Dunham; as soma, see Flattery & Schwartz (1989). For Ephedra as soma, see Falk (1989: 77–90). For Amanita Muscaria as soma, see Wasson (1968). On henbane, a member of the nightshade family, see Müller-Ebeling et al., (2000: esp., 18), where it is stated that; “The Tibetan henbane on Mount Kailash … grows unusually tall and produces seeds that are nearly twice as large as the European ones”. This popular work discusses a veritable pharmacopoeia of psychotropic plants used in the Himalayas, but must be treated with some caution. (Codonopsis [note 60] is not mentioned in the list of psychotropic plants given on pp. 154–155.) 106 chapter 4

Although the collection and use of psychoactive plants was within the be- havioural patterns of certain renunciate groups, climatic and environmental changes make identification of available flora in the Kailas-Manasarovar region difficult to reconstruct historically. But while the identity and disappearance of soma remains one of Indology’s more controversial issues, the study of the role of other consciousness transformative substances in Indic history has been neglected. Yet we can trace a historical progression in the ritual use of var- ious such substances over several millennia. Beginning with the early Indo- European use of mead,62 through the ideologically dominant Vedic cultural agent of soma and the intoxicating darbha grass utilised in the sacrifice,63 it was marked by the prominent place of alcohol in Tantric ritual in the first mil- lennium ce. But this latter period also saw the Tantric use of datura, before cannabis became the renunciates’ drug of choice around the beginning of the second millennium ce. The popularity of alcohol among renunciates may have been affected by its requiring greater preparation and being less easily transportable than herbal agents. Datura, however, is widely available in India and has a powerful effect on the human consciousness. E.B. Havell noted the use of datura as a floral symbol of the Himalayas in the Vaya Purana. He also identified “the long white trumpet-shaped flower of the Datura alba, a poisonous plant sacred to Siva”, in such places as;

[T]he gargoyles which carry off the water used in the temple … for the rit- ualistic bathing of the god [which are] … discharged by a long projecting gargoyle—[resembling] a conventional datura flower—into a stone tank in the courtyard.64

Datura is cited in numerous texts; the Cakrasamvara Tantra, for example, refers to the “five intoxicants”, which Buston’s commentary identifies as the root, stem, leaves, flowers, and fruit of datura.65 The later Vajramahābhairava Tantra hints at the ultimately poisonous effects of the plant; datura is an ingredient in the potion for use in ritual killing (māraṅam).66 But while datura continues

62 Witzel (2004: 594–597). 63 Deeg (1993: 99–112). 64 Havell (1924: 36–37). 65 Gray (2007: n. 12 352). 66 Siklós (1993: 71–76). Siklós suggests datura seeds may have been traded. For further sources on datura use (and cannabis) see, http://people.tribe.net/sahajananda/blog. a tantric kailas 107 to be taken by some Indian renunciates today, its poisonous effects—both physically and psychologically—have removed it from common use.67 The first mention of the use of cannabis as a psychotropic in Indic soci- ety seems to occur in texts from the late first millennium ce.68 It becomes explicit and its use widespread from the beginning of the second millennium, suggesting its spread was stimulated by Islamic groups whose influence on the Indic world began around that time.69 While the leaves and flowers of the dried cannabis plant could be smoked, it became common in the Himalayas to pre- pare the compressed dried resin of the plant (charas), which was then easily transportable. This was the common mode within the Arabic world. Again, we may be dealing with technologies that were only accepted by other groups over time, but from this period on, cannabis rather than alcohol became increas- ingly associated with renunciate groups, particularly of the Śaivite and Tantric persuasion.70

67 In my fieldwork I found renunciates using datura tend to be mistrusted and exiled to the fringes of renunciate society. It is commonly considered to bring actual rather than spiritually-inspired madness. For this reason it has also failed to find a common place in the consciousness-altering sections of Western society. 68 The circa 8th century Cakrasamvara Tantra mentions hemp (Śaṅa = cannabis sativa), as part of a recipe for abundant life as a free ranging yogin; Gray (2007: 373); White (1996: n. 220 412) states that cannabis is called vbijayā (“victory”) in a number of texts and is “according to the TaraTantra (3.1.-11) essential to ecstasy.The 7th c. south Indian Tirumular sings the praises of marijuana in his Tirumantiram.” See however, n. 69 (below). 69 Wujastyk (2002: 45–73, esp., 55–57). Wujastyk disputes (following Meulenbeld: 1989; n. 4 62), the identity of vbijayā, and concludes that the first “incontestable” reference to cannabis use is from the 11th century. Also see, for an extensive survey of (not always reliable) secondary source material which indicates the earlier use of cannabis in Chinese society rather than in Indic, Touw (1981: 23–34). There is considerable literature on the subject; for an early perspective see O’Shaughnessy (1839: 732–742). 70 See for example, Mallinson (2001) concerning The Khecarīvidyā of Ādinātha, a text on Kaula ritual and yoga, probably dating to earlier than 1400ce. This praises cannabis as the supreme drug. Mallinson (2001: 5–15) concludes that ritual alcohol use has been toned down in some recensions of the text. Also see Sanderson (2003–2004: n. 43 365–366), who states that the use of cannabis rather than alcohol derived from eastern India and was subsequently transmitted to Kashmiri Śaivism; the mantra of its ritual empowerment was the same as that for alcohol. While the eastern tradition proclaimed the superiority of cannabis, Kashmiri commentators contested this view. Sanderson supports the theory of Muslim ascetic influence in the introduction of cannabis, although the evidence he presents for east-west transmission seems contrary to that theory. The issue may be complex and multi-directional. 108 chapter 4

From the earliest period the use of cannabis seems to have been associated with Śiva, who was popularly considered to consume vast amounts of the drug. Its consumption was regarded as a sacrament, both an offering to Śiva and prasad from the deity.71 While its use in wider society was common,72 charas became particularly associated with Śaivite renunciates. It became “the necessity of their lives.”73 While alternative psychotropic substances were/are used on occasion,74 and some of the most significant modernist renunciates in the Himalayas were opposed to its use,75 sustenance with “chai, chapatti, and chillum” is a typical daily regime for many renunciates, with each aspect of the “three c’s” preparation highly ritualised. The use of cannabis must have been a fundamental element not only in an economic world but in the transformation of the worldviews of the intellectual and creative classes of the time, and an important factor in shaping under- standings of Himalayan sacred geography. Its appeal to renunciates is partly due to its physical effects, staving off hunger and pain, increasing endurance, and improving sleep in harsh environments. But it is consumed particularly for its visionary properties, which are seen as replicating the state of mind enjoyed by Śiva. Cannabis also produces different understandings of time and space, altering perceptions of landscape and the place of the individual within it. Just as deity visualisation may be enhanced by its use, so too is the experience of the sacrality of landscape.76

71 See, for example, oioc, Emerson, Treatise, chapter xvii, 6–8, which records that, “[l]ower down the valley in the range of mountains separating Kulu from Bhushahr is Shivji ka pind, or Shiva’s phallic emblem, a bare pinnacle of rock … over 18,000 feet”, to which “passers-by offered charas in ‘Shiva’s pipe’,a natural cavity in a rock by a pool of water—now dried up.” I have not sighted any textual sources for the mythology associating Śiva with cannabis. 72 Though abundant in India, the considerable demand for cannabis was also met by large- scale importation from Central Asia to Kashmir and India in the 19th and early 20th cen- turies. Details may be found in the Kashmir Residency reports in the Oriental and India Office Library, l/p&S/ series. 73 Banerjee (1937: 151–152). 74 In addition to substances already noted, it is of interest that Fraser (1820: 353) observed that musk was smoked as a stimulant; the practice was probably too expensive to be common! 75 Although the renunciates discussed in Chapters 9 & 10, Swami Sundaranand and (Sun- daranand advised me), Swami Tapovan, strongly opposed the use of cannabis, it is a defining element of Giri lifestyle and was used for example, by the highly regarded Sri 108 Krishnangiri-ji Maharaj (‘Nanga baba’; d. 1962) an archetypal Giri renunciate in Brahmaur (Chamba Dist. h.p.); see Yadav (1988: passim). 76 “Large oral doses of marijuana are truly hallucinogenic. Vast and celestial visual halluci- a tantric kailas 109

Yet a relationship between renunciate cannabis consumers and Kailas-Man- asarovar should not be overstated.77 It does not grow in that region although it is a common feature of the environment at all of the Indian Kailas mountains (which problematises the popular understanding of Śiva as spending his win- ters smoking cannabis at the Tibetan Kailas). Furthermore, at least two, if not all three of the renunciates most responsible for the modern Indic perceptions of the wider Kailas region (see Chapters 9 and 10), did not use cannabis. Para- doxically however, their successors who partake in their vision of the landscape and follow in their traditions tend to be heavy users. While there may be gen- erational factors involved, some modern renunciates acknowledge the higher mental and spiritual power of those predecessors in such matters, while oth- ers attribute it simply to different modes of practice. All would consider their understanding of the landscape to be, at most, properly appreciated through consumption of cannabis, and certainly not created by it.

Aesthetics

This brings us to another factor in the construction of sacred landscape that is difficult to quantify, the aesthetic factor. Although poets such as Kalidasa and Milarepa celebrated the beauty of the Himalayan realms, Tibetan appre- ciation of the landscape has been described by one commentator as “reserved for sacred phenomena”,78 and as noted, early Indic civilisation regarded the ‘wilderness’ as hostile and forbidding. But in the Indic perspective, at least in the case of the Himalayas, the residence there of numerous major deities implied a heavenly and consequently a beautiful Himalayas. Part of that con- struction rested on the concept of purity. While impure in the sense of being

nations occur … Suggestibility is increased, … [along with] time and space distortion: … and every activity is imbued with a sense of timeless grandeur … It seems an ideal way to attain a sense of one’s own divinity through euphoric experimentation with the powers of one’s mind … The role of cannabis in Tantric ceremony is thus to enable the worshippers to feel the divinity within and without themselves.”; Aldrich (1977: 227–233). 77 It should also be noted that as a rule Buddhist renouncers do not seem to use cannabis or other psychotropic elements, at least in modern times. This raises the question of whether we may isolate elements of a world-view that are not influenced by these substances. 78 Daniels (1994: 181). Given that most, perhaps all of Tibet is embraced by sacred landscape constructions, the issue seems more complex. It has been noted that Tibetan spiritual biographies largely pass over the actual journeys made by their subjects and do not describe either their difficulties or the landscape through which they travel. Further enquiry is needed. 110 chapter 4 beyond the known, the landscape, like the deities and the advanced Brahman- ical renunciates who were equipped to travel there, was ‘pure’ and unsullied by polluting substances and beings. While texts counselled against ordinary pil- grims travelling there due to both worldly and other-worldly dangers that only the spiritually advanced could overcome, those dangers were juxtaposed with descriptions of the beauty of the heavenly landscape. The renunciates who trav- elled into the mountains must have carried those understandings with them and thus been pre-disposed to viewing the environment in a positive way, just as modern visitors do. Appreciation of the upper mountainous regions in Western history largely arose with the Enlightenment, prior to which such areas were considered hos- tile and forbidding. That Enlightenment perception dominated British impe- rial perception of the Himalayas from the late 18th century onwards.79 Travel accounts seemed to compete in superlatives to describe them, with only the occasional dissenting voice. In Tibet however, outsiders did not always appre- ciate the landscape. A British officer wrote of the Kailas region that, “the climate is bleak; the wind is a blizzard; there is fear of dacoits; the road is very bad, and the country is eerie and almost uninhabited.”80 We have one account of renunciate appreciation of the sparse beauty of the sacred realm in its actuality. In 1846, a British official recorded a renunciate who had just returned from Kailas-Manasarovar describing it as a “very beautiful place, no trees, no grass, nothing but rock and snow.”81 But a contrasting beauty confronted those who reached the high-altitude meadows of the Himalayas, such as those near the Ganges source. In summer these isolated plateaus high above the river are well-grassed, irrigated by crystal-clear brooks, carpeted with wild flowers and fragrant herbs, and home to numerous birds and animals. There are snow-capped mountains in the background and plentiful timber for fuel. In the Western aesthetic—as in the modern Indo-Tibetan which it has influenced—these are extraordinarily beautiful places, a view modern renun- ciates share. Their beauty also reinforces the association of this landscape with the realm of the deities and draws renunciates to these regions.

79 On the 18–19th century landscape aesthetic in the Himalayan context, see Bishop (1989). 80 oioc; l/p&s/7/207–1873, W.S. Cassels to Chief Secretary, Government of the United Prov- inces, 23 September 1907. 81 Strachey (1848: 1–2); The Jogi’s description is given in Hindi, translated in a footnote. The original reads “Bahut sundar jagah; per nahin—ghás nahin,—siwá pathar aur baraf kuchh nahim! [sic]”. One might suspect irony in the description, however, and it juxtaposes with the understanding that enlightened individuals can see the heavenly realms of the deity beneath the external appearance of wilderness. a tantric kailas 111

Naths, Giris, and Mountains

The historicity of reputed founders of Indic renunicate traditions often dis- solves under critical analysis and they might be better considered as archetypal or ideal figures rather than specific individuals. As archetypes they are, like sacred toponyms and origin myths, a legitimising device available to various regional and sectarian traditions who may authenticate their status through reference to such common cultural and even cross-cultural symbols. Thus numerous traditions claim to preserve the birth and death places, as well as the fields of achievement, of Gorakhnath, credited as the founder of the Nath renunciates.82 The sect emerged however, as a distinct tradition in 13th century western India, which, as White states;

witnessed a major realignment of a number of preexisting Śaiva religious orders, the creation of a number of new orders, and the appearance of a ‘canon’ of literature on the technical and experiential aspects of hatha yoga.83

The Naths seem to have inherited the earlier traditions of the Pasupatas and Kapalikas, and to emerge from a fusion of such Śaivite tendencies with those of the siddha traditions; those semi-divine figures to which status certain Brah- manical and Buddhist renunciate groups aspired.84 The Naths are prime candidates for having been early renunciate visitors to Kailas. With their advanced ritual technology and inheritance of peripatetic and other features of earlier renunciate traditions they were conceptually equipped to transact with the mountains and the addition of the suffix nath to Himalayan toponyms at sites en route to the Ganges’ sources such as Badrinath, Kedarnath, Rudranath, Baijnath, and (on a route from Kailas-Manasarovar to Chamba), Triloknath, seems to testify to a powerful presence of Naths in this landscape.85 Furthermore, they seem to be the first group of Indic renunciates for whose travels we can identify an economic base.

82 See Briggs (2001: 228–250). 83 White (1996: 93). 84 See White (1996: esp., 2–3, 97–99, 324–327); also see White (2003: 160–187). 85 Badri, Badrinath, and Badrinarayan are all extant names for that place, but Badrinath is accepted by the Government of India and appears on maps (indicating a Nath conceptual victory). 112 chapter 4

Fundamental to renunciation is the need for economic support, without which a renunciate lifestyle cannot be sustained. Indic renunciates commonly relied on patronage from one or other social class, and tailored their practices to sustain their appeal to various patrons. No doubt there were practitioners with sufficient personal charisma to travel beyond Indic cultural boundaries to places such as Tibet, but in most cases travel beyond the social networks necessary for most Brahmanical renunciates to obtain hospitality required alternative means of support, such as that obtained by Pasupatas who served as armed guards for trading guilds.86 The demand for alchemical substances meant however, that Nath alche- mists could create transactional structures that freed them from bonds of patronage. If Tibet’s Kailas-Manasarovar had not previously been identified as a sacred Indic site (and there is no real evidence that it had been), the moti- vation for a journey there seems largely absent. But the rise of an alchemical economic world could support a trade route to Manasarovar and Tirthapuri, then on via Triloknath and Chamba-Brahmaur to Jvalamukhi. Kailas would ini- tially have been an incidental part of the landscape of such a journey, but it appears that a new sacred geography was developed in which the ideal, heav- enly Kailas was identified with the Tibetan site, conceptualised as the abode of the Naths’ predominant deity Śiva. That Śaivite sacred geography later became part of Pan-Indic understanding. As will be seen, there had been considerable regional cross-border religious exchange for several centuries before this as a result of the development of Buddhism in Tibet and in the early 13th century Kailas had emerged at the centre of a new sacred geography advanced by one sect of Tibetan Buddhism. That new construction surely impacted on whatever bodies of knowledge were held by the Naths about the site and at least partly inspired that Kailas-centred Śaivite sacred geography which fully emerges by around the 17th century. There was also a specific Nath identification with mountain practice. White, noting the Taoist’s explicit “identification of … semidivine figures and alchem- ical apparatus with [the] sacred mountains themselves”, identifies “behind the medieval Indian cults of divine Siddhas and Vidyadharas as denizens of moun- tains … a more archaic cult of these mountains themselves as a group of demi- gods.”87 Yet ultimately the Naths, like so many such groups, are with present sources impossible to fix in many details necessary for Western historical anal-

86 Davidson (2002: 80); also see Lewis (1993; esp. 173) who draws attention to merchants as possible sponsors of religious travelers. 87 White (1997: 85–86). a tantric kailas 113 ysis. As a peripatetic group without fixed centres the early Naths float free of such definition, textual references to them invariably seem several centuries after the fact. By the late 16th or early 17th century, numerous peripatetic renunciate tra- ditions were systemised under the umbrella of the Daśnāmī (‘ten names’) order,88 which structural body came to dominate north Indian Śaivite renunci- ation. Although an independent Nath presence in the Himalayas continues to this day, their influence, practices, and trading circuits were largely inherited by a Dasnami order whose name specifically identified them with mountain practice, the Giri (Skrt: ‘hill/mountain’) sect.89 While the vratyas had ventured beyond the boundaries of Vedic civilisation to conquer and sacralise the wilder- ness, and Pasupatas and Naths expanded that territory, in the modern era it is the Giris who have become the principle Sanskritising agents in the upper Himalayas and in particular the Kailas mountains. As our records of renounc- ers reaching Kailas-Manasarovar become historical, the individuals noted are predominantly from the Giri order.90 According to traditional histories of Indic renunciation, Sankaracharya (Śaṅ- karācārya; trad: 788–820ce.) was the key figure in the establishment of the Dasnami. But Dasnami chronicler Mathew Clark has found, “no reference in any text to the ten names before the sixteenth century.”91 He attributes the founding of the Dasnami as a response to the orthodox Muslim regime in Delhi, which determined the military character of the new renunciate formations (akhāṛās).92 (That character also reflected, as we have seen, the militant strand in renunciate traditions from the era of the vratyas.) But the Himalayan realms were effectively beyond the reach of Mughal authority. Thus they provided a space in which religious processes, including claims and contestation around sacred space, could continue outside of the Islamic, and later British, states.

88 On which see Clark (2006). This work is fundamental to the study of the history of renunciation in India. 89 Dating the transition is problematic, but see van der Veer (1988: 146–147) which gives a list of abbots of the Dasnami matha in Ayodhya. From 780–1374, the names all end in nath, after that they all end in giri. The list may not be historical, and the take-over seems unsustainably early, but it is the Giris’ articulation of their take-over from the Naths. 90 Two renunciate informants cited by Duncan (1801: 49) are an exception. One is identified as a member of the Puri (order of Dasnami sannyasis) and the other is identified as “a Brahman of the Yajurveda sect” and I am not aware of Giris following that tradition, though it is not impossible. 91 Clarke (2006: 148–176, & passim, quotation from 174). 92 Clarke (2006: passim, esp., 3, 25, 227, 246). On the militarisation of renunciation in this period, see in particular, Pinch (2006). chapter 5 An Early Buddhist Kelāsa

Introduction

The modern understanding that Kailas is sacred to Buddhists as the home of the Tantric deity Demchok/Cakrasamvara is entirely absent from the early Indic Buddhist material preserved in the Pali sources and other texts1 predating the introduction of Buddhism to Tibet. Most notably it does not occur in the Abhidharmakośa of Vasubandhu, the foundational text of later Tibetan Buddhist cosmology. In this chapter we discuss the Kailas complex as it appears in the early Buddhist sources, which represent both similar and distinct bodies of knowledge to those found in the Epic and Puranic texts. The Pali material comprises the canon, commentaries and other paracanon- ical texts which form the doctrinal foundation of Theravada Buddhism. While the core of this material is traditionally considered attributable to the Buddha himself, or to have been recorded within a century or two of his passing,2 the earliest written texts probably date to around the first century bce. But it is Buddhaghosa’s 5th century ce. commentaries that mark the actual establish- ment of the later canon, and while much of it is certainly far earlier, it is only from that era—and on the basis of much more recent manuscripts—that the specific content of the canon can be properly established.3 Thus in discussing the Pali references to Kelāsa (Pali: Kailas) or a related sacred lake, we face problems very similar to those we encounter with the Sanskritic material. Firm dating is usually impossible and informed estimates may be several centuries out. Traditional datings actually suggest the Pali texts represent a somewhat earlier discourse concerning the Himalayan realms than the material in the Sanskrit Epics. But there are too many variables to permit

1 Including those in Buddhist hybrid Sanskrit, etc; for simplicity I use ‘Pali’ in the general sense to describe the early Indic Buddhist texts of the Theravadin tradition. 2 An attribution that has academic support; see for example, de Jong (1993: 25); Gombrich (2006: 20). Given my unfamiliarity with Pali, I gratefully acknowledge the assistance of Lance Cousins in providing me with a list of the principal Pali sources on Kailas, and of Greg Bailey and in particular Professor Richard Gombrich for assistance in translating and understanding the Pali material. 3 Schopen (1997: 3, 23–24); also see Davidson (2002: 146–147).

© koninklijke brill nv, leiden, 2015 | doi: 10.1163/9789004306189_007 an early buddhist kelāsa 115 us to safely conclude that a Buddhist body of knowledge preceded other such bodies, not least that most of the relevant references in Pali sources appear in the later commentarial literature, and thus cannot be proven earlier than similar references in the Sanskritic tradition. G.P. Malalasekara for example, recounts a Pali myth in which Kelasa is mentioned.4 This involves an evil Himalayan king, Alavaka, who is in league with a demonic yaksa. On hearing of the Buddha’s arrival the king (who is eventually converted by the Buddha), was angered, and placing his left foot on Manosilatala and his right on Kelasakūta,5 uttered a shout heard throughout the continent of Jambudivipa. This myth appears however, only in the later commentarial literature. Also probably late is the description of the seven great lakes and the five surrounding mountains of the Himalayas, which identifies Kelasa as one of five mountain ranges surrounding a Lake Anotatta. In seemingly earlier sources Kelasa is of little prominence, and again there is no evidence to elevate the sanctity of the site above many other features of what is a heavenly rather than earthly landscape. Nor do the locations referred to seem to match modern locations with similar toponyms.

A Jataka Parallel

Most of the earliest Buddhist references assumed to refer to the Kelasa region occur in the various biographies of the lives of the Buddha and those of his disciples, the 500 Arhats.6 We have seen that early Indic references to Kailas were usually in the context of metaphor and this is generally the case in the Jatakas.7 A white elephant, for example, is as “tall as a peak of Mount Kelāsa.”8

4 Malalasekera (1937) Dictionary, consulted on: www.palikanon.com/english/pali_names/dic _idx.html, which does not give page numbers. 5 Malalasekera (1937) citing sna.i.223; sa. (Sāratthappakāsinī, Samyutta Commentary) i.248. kūta, according to Rhys David’s Pali dictionary carries a variety of meanings including “shoul- der”, “point”, “summit”, “peak”, “prominence”, “a jewelled top”, or “a pitcher” (c.f. “vase”); http:// www.abhidhamma.com/Pali_English_Dictionary_RhysDavids.pdf. Accessed December 2010. 6 These accounts are found particularly in the Jatakas, in the Apadāna (Pāli: ‘stories’) biogra- phies in the Khuddaka Nikāya section of the Sutta Piṭaka, and in the Mahāvastu. 7 Malalasekara (1937) states that Kelāsa is “used in similes to describe an object that is perfectly white …, very stately (e.g., an elephant’s head or a big building) …, or difficult to destroy.” 8 Cowell (1901: 145–146; Jataka No. 479: Kāliṅga-bodhi); also see Chalmers (1895: 104; Jataka No. 40: Khadiraṅgāra), where the Buddha’s forehead is “as the snowy crest of Mount Kelāsa.” 116 chapter 5

That Mount Neru/Sineru (ie: Meru/Sumeru) is also used in this way implies at least the basis of an existing cosmology later made explicit; we read of the earth, “which can support the mighty weight of Mount Sineru and its encircling peaks.”9 But there is certainly no suggestion that Neru/Sineru and Kelasa are identical. While implicitly or explicitly identified as within a sacred landscape, there appear to be no references in these texts that suggest Kelasa is of any partic- ular importance beyond the metaphorical. There are, however, a number of more detailed references in the Jatakas to sites that came to be associated with the later sacred geography of Kailas. There is the Lake Anotatta (‘not warm’: Skrt; Anavatapta),10 most commonly identified with Manasasarovar, Gandhamadana, which as in the Sanskrit sources seems only ever to be a mountain (and probably that known today as Gurlha Mandhata), and there is Manosilatala, which Frits Staal took to be a lake, indeed Lake Manasarovar. That enabled him to identify Anotatta with today’s Rakas Tal and implied a ref- erence to two lakes, the duo so oddly absent from early Indic sources. But Staal erred, for the references to Manosilatala clearly denote not a lake but a moun- tain or a prominent outcrop of a mountain on the plain beside Anotatta.11

(I cite the collection of Jatakas originally published in six volumes by Cambridge Univer- sity Press, 1895–1907; now available at www.sacred-texts.com. Their translation is seriously antiquated.) 9 Cowell (1901: 176; Jataka No. 72. Sīlavanāga). 10 See Francis (1897: 168; Jataka No. 382: Sirikālakaṅṅi); and Francis (1897: 230; Jataka No. 408: Kumbhakāra). 11 Staal (1990). Staal relies on Malalasekara’s work and his conclusion that Manosilātala is a lake rests on a misreading of Malalasekara’s dictionary reference to Anotatta, where it is mentioned once: “Other instances are given of goddesses bathing in the lake [Anotatta] and resting on the banks of the Manosilātala next to it (e.g., J.v.392).” Staal (1990: 276) repeats this citation (found at www.palikanon.com/english/pali_names/ay/anotatta.htm, accessed 9 July 2010). Malalasekara’s Dictionary (1937) reference to Manosilātala, however, clarifies this con- fusing statement. He does not identify Manosilātala as a lake, but states that, “Manosilātala [is] a locality in Himavā [emphasis added]. When Alavaka threatened the Buddha he stood with his left foot on Manosilātala and his right on Kelāsa (sna.i.223). Manosilātala was near Anotatta, and those who bathed in the lake dried and robed themselves there (e.g., J.i.232; iii.379).” The story of Alavaka makes little sense if Manosilātala is a lake—he is unlikely to be described as standing with one foot on a mountain and the other in a lake! That Manosilātala is indeed a mountain is confirmed by the Sudhābhojana Jātaka (v.392: v.535 in Jataka, Vol. v, [H.T. Francis; trans.], 1905. Available at www.palikanon .com/english/pali_names/ka/kancanaguhaa.htm), as well as by the logic of references to an early buddhist kelāsa 117

A more significant geographical account may be that in the lengthiest and last of the 547 Jataka stories, the Vessantara Jataka which in its final form is no later than the 5th century ce.12 It reflects many of the themes prominent in the Sanskrit Epics,13 with its exiled king Vessantara travelling with his wife to a Mount Vaṁka, where they live as renunciates before eventually regaining their kingdom. Mention of a Mount Suvaṇṇagiritāla suggests their exile is in the Himalayas,14 a location confirmed by their reaching “that rocky mountain … Gandhamādana.” They then travel north to a Mount Vipula from where they can see “deep” river Ketumatī. A Mount Nālika is reached, then;

To the north-east of that is Lake Mucalinda with its covering of lotuses and white water-lilies … the forest, dense like a cloud, green with grass, full of trees in flower and trees heavy with fruit … countless birds of many colours … with their beautiful staccato chirpings. When you have passed the difficult terrain of that mountain and the sources of the rivers, you will see a lotus pond surrounded by karañja and kakudha trees. Its copious waters … sweet and fragrant.15

Lake Mucalinda seems distinguished here from the other, unnamed spring-fed lake, although the heavenly descriptions of Himalayan flora and fauna are as mythologised as the Epic descriptions of Manasarovar’s surrounds. While there is a reference to an elephant, as “white as Mount Kelāsa with his snows”,Kelasa is not mentioned in the description of the voyage north from

Yakkhas holding regular assemblies there (sna.i.187; cp. d. [Digha Nikaya, 3 vols.] iii.201 and da. [Sumangala Vilāsinī, 3 vols.] iii.967); the Budda’s mother being placed by the gods under a sāla tree on Manosilātala (d.iii.52); and by the reference to a Paccekabuddha who “washed his mouth in Lake Anotatta [then] clothed himself in Manosilātala” (S-a [Sāratthappakāsinī, Samyutta Commentary] I 160). Nor does Manosilātala occur in the list of lake toponyms discussed in the text below. Staal’s proposed identification of Manosilātala with Manasarovar (and of Anotatta with Rakas Tal), thus cannot be sustained with this evidence. 12 Cone and Gombrich (1977: xxxi). 13 For example, just as Rama warns Sita of the hardship of life in the forest in the San- skrit Ramayana, so too does does Vessantara caution his wife Maddī; on the similarities between these two texts, see Gombrich (1985: 427–437) and on resemblance to the Mahab- harata, see Cone and Gombrich (1977: xxviii). 14 On an earlier ‘core’ story in which Vessantara’s exile is to mountains in Orissa, see Cone and Gombrich (1977: xxix). 15 The translation is from Cone and Gombrich (1977: 38–39). 118 chapter 5

Gandhamadana to Mucalinda, in practice an interesting omission if Mucalinda equates to Manasarovar (or Rakas Tal). Kelasa again seems an established metaphor for a heavenly place rather than an integral part of a Himalayan sacred geography. In this account an alternative toponymical strategy seems to be advanced. Toponyms such as Ketumati and Nalika are specific to these early Buddhist accounts. Even Mount Vamka,16 the oft-mentioned here sacred peak that is the object of Vessantara’s journey is otherwise unknown, and the separate insignificant mention of Kelasa seems to rule out its identification with that peak. Mucalinda as a toponym is almost as obscure as Vamka; the name is prominent as that of the Naga who shelters the Buddha under his hood,17 while in other Pali sources Mucalinda occurs as the name of a tree, a dragon king (or perhaps a Naga king—born in an unnamed lake near the tree),18 and as the name of a yaksa chief.19 But no other lake of that name, or use of it to denote Manasarovar, occurs in the Pali or Sanskrit sources consulted for this study. We are therefore dealing with a body of knowledge that might seem to reference the Kelasa region—imagined or experienced—but which is almost entirely toponymically distinct from the Sanskrit accounts. Only Mount Gand- hamadana is consistently common to both. A simple juxtaposition of toponyms may be proposed, but is problematised by Kelasa’s appearance in metaphor but not in descriptions of the landscape. Again we seem to be dealing with sepa- rate constructions, and/or mythological cycles that are only partly systemised in these sources. None-the-less, the description of the route from Gandhamadana to Lake Mucalinda is articulated in the Vessantara Jataka with a clarity lacking in any other pre-modern accounts that suggest the route to Kailas. There is actually a smaller peak (Mount Vipula?) north from Gandhamadana, and the Ketumati

16 I have not sighted any reference to this mountain outside of the Jataka context. 17 On which see, for example, Rawlinson (1986: 135). 18 Vinaya Piñaka, Mahàvagga, Mahàkhandhaka, bhàõavarà 1–4. Mucalindakathà, available at www.ancient-buddhist-texts.net/Texts-and-translations/Mahakhandhako/ Mahakhandhako.pdf, accessed 12 August 2009: Bollee (2007, n. 156 34) cites a Naga king Mucalinda, with his seven heads sheltering the Buddha from a heavy monsoon rain; from Vinaya I 3; Mahāvastu iii 301, etc. 19 d.iii.205; Ap. [Apadāna, 2 vols.] ii. 536 (verse 86), available at http://www.palikanon .com/english/pali_names/s/sineru.htm. There is also a reference to Mucilinda, a “dwelling place” in the city called “Pure Other Shore”, in the Avatamsakasutra: Cleary (1984: vol. 2, 219). an early buddhist kelāsa 119 might be identified with the Sutlej or one of the minor rivers visible in that region. Mount Nalika is difficult to identify unless it is a minor peak, but this does allow that “further still towards the north is Mucalinda Lake” if that be identified with Manasarovar or Rakas Tal. While most descriptions of the region in ancient texts suggest, on close examination, journeys in the upper Gangetic region, this reference is probably the closest to the actual geography of the Kailas-Manasarovar region we know today, suggesting an imagined realm informed by some reality. What the Pali, Epic and early Puranic literature shares is the imagined heavenly nature of the upper Himalayan region, the flowers, birds and forests that are entirely at odds with the actual landscape around Kelasa in modern times. Thus we read of King Vessantara and his family dwelling as renunciates at Mount Vamka and at Gandhamadana, in their hermitage surrounded by trees bearing various tropical fruits—mangoes, jack-fruit, figs, dates, grapes, and coconuts. There are honey bees among the scented flowers, rice and mustard grow there, and fish, crabs and tortoises abound in the lakes. Other renunciates abide there also, as do all manner of creatures; squirrels, monkeys, deer, lizards, tigers, yaks, and elephants. This is clearly not intended to describe an actual earthly realm. Rather it is a heavenly land, and while no specific prohibition on ordinary householders is articulated, the Kunāla Jātaka confirms that it can only be reached by those with magical powers. When the Buddha wishes to take the Arhats there, they reply that, “we have no supernatural powers; how should we go?” Reaching this realm is only possible for them when the Buddha, “by his miraculous power caught them all up with him in the air and transported them to the Himalayas.”20 Yet the portrait of a heavenly landscape is layered with prosaic observations. The Vessantara Jataka refers to the hero being aided in the miraculous realm by a hunter, who supplies him with deer meat to eat. The encounter between hunter and hermit is a common literary trope in Indo-Tibetan sources—in the life of Milarepa for example—but was one with a basis in reality, for the two types must have shared the remoter hill regions from the earliest period. A similar reality is glimpsed in the earlier Mahavastu, where Kelasa is iden- tified as the region of the Kinnari tribe.21 This tribe are depicted in ways some- times human, sometimes seemingly animal, reflecting the perspective of the

20 Francis (1905: 221; Jataka No. 536: Kunāla Jātaka). 21 Mahāvastu (vol. 2; 97, 109: see also vol. 3; 309). Nakamura (1987: 130) dates the Mahavastu (which includes 40 Jataka tales), to “perhaps around the 2nd century b.c.” 120 chapter 5 literate Indic elites. A renunciate is told by a hunter about a lotus pond north of his retreat and that, “the daughter of Druma king of the Kinaras [said to live on the summit of Kelasa] … is wont to go from mount Kailāsa to disport at the lotus pond.” There is a sense of ancient history here, as there is with the occa- sional acknowledgement of local Naga tribes and kings. But it is the mythical Kelasa that becomes more prominent in the Mahavas- tu. It is now, “the abode of the Yakśa hosts and the home of the Rāksasas”, and there is a great renunciate, Asita. He “dwelt in the region of the Vin- dhyas [the modern toponym of a range of mountains in western India] … [and now] … dwelt on Kailāsa’s summit, like a lord of wealth [cf. Kubera?], albeit a lord of a cave …”.22 This is a Kelasa more akin to the Epic Kailas, as if an earlier understanding had been overlaid by a culturally dominant narra- tive.

Anotatta and the Pali Lakes

The Pali sources demonstrate the early Indic primary concern with bodies of water rather than mountains, for they contain far more references to lakes than to peaks. If Lake Mucalinda is specific to a single Jataka narrative, Lake Anotatta is a far more prominent regional toponym in early Buddhist sources and its identification with Manasarovar is common in Western scholarship. Most of the references to Anotatta, however, occur in the commentarial (i.e., post-Epic) literature. Those in the canonical texts which are probably earlier are outside of worldly time and space. They reference the lake within the Tāvatimsa heaven, where the Buddha first taught the Abhidhamma to an assembly of devas.23 But he is said to have descended daily to the semi-human realm and we read that, like other renunciates and monks, he visited Ano- tatta.24 In this context lies what must be considered a very early reference to Ano- tatta, from the circa 3rd century bce., Mahavagga, part of the Khandhaka sec- tion of the Vinaya Pitaka. This single isolated mention of the site has been referred to by Staal using Malalasekara’s translation, which reads;

22 The Mahāvastu (vol. 2; 30, 91–111; vol. 3; 296). 23 DhA.iii.222; cited in Malalasekara (1937). 24 “Examples are given of other holy men doing the same. e.g., Mātangapandita, J.iv.379; see also DhA.ii.211.”; also see Dvy. 399 [Divyāvadāna]; Malalasekara (1937). an early buddhist kelāsa 121

The Buddha would often go to Anotatta for his ablutions and proceed from there to Uttarakuru25 for his alms, returning to the lake to have his meal and spend the hot part of the day on its banks.26

But the wider context of this passage clearly locates the citation in the realms of the miraculous, as part of the demonstration of the Buddha’s magical pow- ers. Uruvela Kassapa, the leader of a company of fire-worshipping Brahmanical renunciates who frequently appears as a spiritual competitor of the Buddha in the early Buddhist accounts, plans a ceremonial meal at which he hopes the Buddha will perform a miracle. The Buddha, reading his mind, transports him- self from Uruvela to Uttarakuru for the day, collecting alms there, eating them at Anotatta, and after resting, returns to Uruvela in the evening, astonishing his host.27 Anotatta is thus a metaphor for a far-off place, a heavenly site beyond the reach of those lacking magical powers. The legend indicates that Anotatta exists in the higher realms; any earthly location is neither clearly indicated nor relevant to the point being made. Sim- ilarly, the lake is the backdrop for the Buddha’s teachings in the Anavatapta- gäthä (Song of Lake Anavatapta),28 where again it is a heavenly place not a worldly one. These associations are literary devices to emphasise the Buddha’s status, not to advance a sacred earthly geography. Nor do these myths relate the lake to a Kelasa mountain. We may infer that both are within the heavenly realms, but neither are central to that realm, or linked in any way. The Jatakas and later commentarial texts expand the understandings of a liminal lake. The Buddha’s mother dreamt of bathing there on the day of her conception; Anotatta was stated to be “one of the last to dry up at the end of the world”,29 and, in a theme that recurs in various forms in association with Kailas mountains, regular assemblies of deities were held there, often attended by the Buddha. Anotatta seems to have been a conceptually essential component of an imagined landscape of heaven, necessary for the gods to

25 Uttarakuru (‘northern realm of the Kurus’), is, like Uttarakaśi (‘the northern Kaśi’) an undefined place of sacral associations, or at least one ritually purified, and probably refers generally to the realms where the Ganges descends onto the plains. 26 Staal (1990: 276). The cited passage is from the Vinaya Pitaka, 1.28. 27 I am extremely grateful to Professor Gombrich for this translation and explanation of the wider context of this tale. 28 On this and related literature, see Hofinger (1982: esp., 177–181 which notes earlier schol- arship concerning the lake); Salomons (1999; esp., 158–159) & (2008); Mejor (1987); Naka- mura (1987: 25–26). 29 Malalasekara (1937) citing a. [Anguttara Nikaya, 5 vols.] iv.101. 122 chapter 5 carry out the bathing rituals integral to Indic spiritual practice. As in Indic society, there were social or at least status divisions in that heaven, with the lake having separate bathing places for “Buddhas, Pacceka Buddhas, monks, ascetics, the Four Regent gods and other inhabitants of the deva-worlds, and for the goddesses.”30 Just as the earliest Sanskrit references to Manasarovar indicate it was the home of a Naga, so too does Anotatta have its resident Naga king, here named Pannaka.31 This is the clearest indication of at least a conceptual link between the two, but Naga lakes were regionally profuse and thus it may equally be a literary device. The lake-waters were considered medicinal,32 as are those of Manasarovar today, but were said to be obtainable only by those (such as the Buddha’s powerful novice Sumana), who possessed magical powers (siddhi). In a prosaic note, however, that water was supplied to great earthly kings such as Aśoka and Vessavana, whose water-fetching yakkhinis might die of exhaustion from their work.33 Again, therefore, we see elements of a real world projected into imagined heavens and vice-versa, but such imaginings were necessarily based within the cultural knowledge of their creators and cannot be isolated as evidence of a historical world.34 It is only in the commentarial literature, dating to at least the 5th century ce., that Anotatta assumes the more central role in the proto-cosmological pat- tern that was canonised in the Abdhidhammakosa. Even there however, Ano- tatta is not singled out for specific sanctity, only described as one of the seven great lakes of the Himalayas, along with Kaṇṇamuṇḍa, Rathakāra, Chaddanta, Kuṇāla, Sīhappapāta, and Tiyaggala,35 with the latter often substituted by the

30 Malalasekera (1937) citing ma.ii.918; VvA. [Vimānavatthu Commentary], 131–132, Ap. [Apa- dāna, 2 vols.], i.299. 31 Malalasekera(1937) citing DhA.iv.134, also ThagA. [Theragāthā Commentary, 2 vols.] 457. 32 Malalasekera(1937) citing DhA.iv.129. 33 Malalasekera (1937) citing DhA.iv.134ff., i.40; Sp. [Samantapāsādikā, 4 vols.] i.42; Mhv. [Mahāvamsa] v.24; 84; xi.30. For Sumana, the promising novice, bringing water “from the great lake, Anotatta”, see Thag [Theragāthā], 6.10 (Thag 429–434). 34 It is in that context that we might note a single reference in the early Buddhist sources that refers to two lakes. The Culla Nidessa, an early (pre-Christian era?) commentarial text included in the Sutta Pitaka mentions lakes named Mānasaka and Anotatta. The reference is, however to possible destinations, and the former is identified by Malalasekara with a village in Kosala on the banks of the Aciravatī river. Malalasekara (1937) da.ii.399. I am indebted to Lance Cousins for this reference and to Richard Gombrich for his comments on it. 35 See Francis (1905: 221; Jataka No. 536: Kunāla). Malalasekara (1937) states that the same list appears at Vsm. (Visuddhi magga, 2 vols.), 416. an early buddhist kelāsa 123

Mandakini, the only title common elsewhere.36 (This list omits Mucalinda, the lake toponym prominent in the earlier Jataka tale, and which now van- ishes from the texts.) Five mountain peaks said to surround Anotatta are also named in this literature. They differ from the earlier accounts and now include the obscure Citrakūta and Kālakūta, along with the familiar Kelasa, Gand- hamadana, and Sudassanakūta. Malalasekara, drawing on three commentarial texts, summarises these later accounts of the hydrology of the region as follows:

All the rains that fall on the five peaks and all the rivers that rise in them flow into the lake [Anotatta] … Four channels open out of the lake in the direction of the four quarters: Sīhamukha [‘Lion’s mouth’], Hatthimukha [‘Elephant’s mouth’], Assamukha [‘Horse’s mouth’] and Usabhamukha [‘Bull’s Mouth’]. Lions abound on the banks of the Sīhamukha; elephants, horses and cattle respectively on the others. Four rivers flow from these channels; the eastward river encircles the lake three times, waters the non-human regions of Himavā and enters the ocean. The rivers that flow north and westward flow in those directions through regions inhabited by non-humans and also enter the ocean. The southward river, like the eastward, flows three times round the lake and then straight south over a rocky channel for sixty leagues and then down a precipice, forming a cascade six miles in width. For sixty leagues the water dashes through the air on to a rock named Tiyaggala, whereon by the force of the impact of the waters the Tiyaggalapokkharani [lake] has been formed, fifty leagues deep. From this lake the waters run through a rocky chasm for sixty leagues, then underground for sixty leagues to an oblique mountain, Vijjha,37 where the stream divides into five, like the fingers of the hand. The part of this river which encircles the original lake Anotatta is called Āvattagangā; the sixty leagues of stream which run over the rocky chan- nel, Kanhagangā; the sixty leagues of waterfall in the air, Ākāsagangā; the sixty leagues flowing out of the Tiyaggala-pokkharanī and through the

36 Malalasekara states that the Mandakini, “[o]ne of the seven great lakes of the Himālaya”,is used rather than Tiyaggala in j.v.415; j.ii.92; a. [Anguttara Nikaya, 5 vols.] iv.101; sna.ii.407; d i.54; da. [Sumangala Vilāsinī, 3 vols.] i.164, 283; UdA.300; aa. [Manorathapūranī, Angut- tara Commentary, 2 vols.] ii.759. 37 “Vijjha: A horizontal rock on which the stream, flowing from the eastern mouth of Ano- tatta, divides into five rivers …: referred to in … sna.ii.4,39; aa.ii.760; ma.ii.586. [etc.]”; cited in http://www.palikanon.com/english/ pali_names/vy/vijjha.htm. 124 chapter 5

rocky gorge is called Bahalagangā, and the river underground, Ummag- gagangā. The five streams into which the river is divided after leaving the oblique mountain Vijjha are called Gangā, Yamunā, Aciravatī, Sarabhū and Ma- hī.38

There are strong conceptual similarities between this cosmology and those of the Sanskritic literature. Most of the toponyms are unique to this account however, and notably there is no central mountain involved, nor do the rivers flow from a heavenly source. But the Ganges is given five names depending on the part of the river referred to, something that does occur in later Indic geography and the toponyms of the Ganges after its division into five streams are those also cited in the circa 400ce. account of Fa-hsien,39 suggesting their wide currency at the time. But just as it is notable that Kelasa has no dyadic, indeed any particular link to Anotatta, so is any link to Meru-centred cosmology absent here. That cen- tral world mountain (Neru/Sineru) is in these sources topped by the divine world of Tavatimsa.40 According to texts including the Mahavastu it is sur- rounded by seven mountain ranges; Yugandhara, Isadhara, Karavīka, Sudas- sana, Nemindhara, Vinataka and Assakanna,41 of which only Sudassana is a toponym otherwise commonly associated with the Kailas region. The charac- ter of the Neru/Sineru central world mountain is briefly fleshed out in the Neru Jataka, where the Buddha in an earlier life is said to have settled there with his younger brother and to have become “golden of hue from its lustre.” This “undis- criminating Neru”,“Noblest of mountains”,was “home to many birds and beasts and honoured by all without distinction”,42 presumably because it lay beyond worldly divisions of good and evil. Like Kelasa, it was an imagined heavenly mountain.

Foundations of the Later Tibetan Traditions

The Abhidharma is a canonical body of Buddhist literature concerning the nature of existence. It exists in different collections that represent the teach-

38 Malalasekara (1937) citing SnA.ii.407; 437–439; ma.ii.585f.; aa.ii.759–760. 39 Hoey (1907: 41–46). 40 Malalasekara (1937) citing sna.ii.485f. 41 Malalasekara (1937) citing sna.ii.443; Sp. i.119; Vsm. 206; cp. Mtu. [Mahāvastu, ed. Senart, 3 vols] ii.300; Dvy. 217. 42 Francis and Neil (1897: 160; Jataka No. 379: Neru). an early buddhist kelāsa 125 ings according to different sects. Various versions reached Tibet, but most influential there was the Abhidharmakośa (Treasury of Higher Knowledge) of Vasubandhu, dating to around the 5th century. An account of the phenomenal world is only a very small part of this vast compendium of Buddhist lore, but describes a Meru-centred cosmology that includes a section concerning our sacred region. After describing the fantastic appearance of Meru and the mountains that surround it, the text describes Himavat (“the mountains of snow”) as being to the north of nine “ant” or “black” mountains.

Beyond that, this side [i.e., south?] of the Gandhamadana (‘the Mountain of Perfume’), lies Lake Anavatapta from whence there flows out four great rivers, the Ganga, the Sindhu, the Vaksu and the Sita. This lake, fifty yojanas wide and deep, is full of a water endowed with the eight qualities. Only persons who possess magical powers can go there.43

Thus even in this classic exposition of a lake and four rivers, there is no asso- ciation with Kelasa. While the existence of numerous different versions of the Abhidharma complicates any conclusions, not least because later Tibetan ref- erences to it cite elements that cannot be traced to this particular text, several other points stand out here. Again, there is a lake that is not reachable by ordi- nary persons, and it is notable that the rivers said to originate there include just one of the four now associated with Manasarovar. But as emphasised, this lake is (apparently) located to the south of Gandhamadana, and Kelasa is not mentioned. But there were other canonical texts later cited by Tibetan proponents of the sanctity of Kailas-Manasarovar and specified in guide-books to the site.44 One such text was the Satipaṭṭhāna Sutta, one of the most significant discourses in the Pali canon. Although there actually seems to be no explicit reference to the site there, the tendency of such works to attract a large body of commentarial literature means that it may be referred to in that associated material. There are however, relevant references in another text frequently cited as referring to the sacred site, the Avataṃsaka Sūtra (also known as the Mahāvaipulya Buddhāvataṃsaka Sūtra: Tib: mdo-phal-po-che). This text consists of a number of originally independent sutras compiled during the centuries after the passing of the Buddha and eventually combined

43 Pruden (1990: vol. 2, 456), Abhidharmakosabhasyam of Vasubandhu, (chapter 3–57; also see chapters 3.48–56). 44 Filibeck (1988: 70). 126 chapter 5 around the late 3rd or 4th century ce. It continued to be augmented with later material and was translated into Tibetan in the early 9th century.45 Its refer- ences to our site may thus be very late. But drawing on the World-mountain cosmology, it does refer to the mountains, rivers and ‘heatless’ lake of that construction, although in the context of metaphors for the achievements and powers of enlightenment beings, or Buddhas. Thus we read that;

Just as the great lake Heatless [i.e., Anotatta] pours forth four rivers from four mouths into the ocean, so also do great enlightenment beings pour forth various practices from the four powers of understanding, ultimately to enter the ocean of omniscience.46

While relevant metaphors and conceptual models are again apparent,47 the text does expand the Abhidharma description of the four rivers, adding the association of their origin with four mouths on the four sides of the lake, those of an elephant, lion, ox, and horse. It also names the “ten great mountains”, which include the “Fragrance mountain” (Gandhamadana?), and the “Snowy Mountains.”48 This latter designation is crucial to the history of the site because it allowed later Tibetan commentators to identify the mountain they called Tise as the one referred to in the canonical scriptures of their tradition. While the precise etymology of Kailāsa is problematic, it was, as will be seen, colloqui- ally used in India to mean ‘snowy mountains’, which was also the approximate meaning of Himavat. Thus a guidebook to the site could state that, “Tise … was also called Himalaya”.49 Such blurred meanings are important here, they enabled blurred identities, and allowed the toponymical flexibility to inter- pret Himavat, Gandhamadana, Kailas, Meru and other peaks as, according to context, one mountain, a range of mountains, or the same mountain(s). The Tibetan commentators, dealing with varied texts, translations, and Buddhist

45 Susumu (2007: 86–92). 46 Cleary (1986: vol. 2, 154). 47 Thus, for example, “Eulogies on Mt Sumeru” is a chapter title, but there is no mention of any mountain in the chapter; Cleary (1984: vol. 1). This is a common phenomena in Indic literature, an authority of place in which the sacred location of their transmission adds status to the teachings. 48 Cleary (1986: vol. 2, 112, 153 and see 313.) No explicit mention of the ‘Great Snow Mountain’ later cited by Tibetan commentators such as Jigten Gönpo (see Chapter 12) is actually apparent in this translation, although it cannot be ruled out as existing in one of the many commentaries or texts. 49 Filibeck (1988: 67). an early buddhist kelāsa 127 traditions must themselves have struggled to make sense of issues such as sacred geographies, but conversely they also enjoyed considerable flexibility of interpretation as a result of the many anomalies, inconsistencies, and inter- polations in the literature they were translating and expounding.

A Buddhist World-View

Despite the conceptual similarities between the early Buddhist and Brahmani- cal sources almost entirely distinct sets of toponyms are applied in the two tra- ditions. Gandhamadana (or some closely related toponym), is common to both sets, as is Mandakini, and if the latter is a shifting signifier the identification of Gandhamadana with the modern Gurla Mandhata may be the only consistent mountain toponym we encounter in early Indic sources, as the Ganges seems the only consistent river toponym. We must conclude that this is an indication of toponymical strategies by different traditions, competing terms as symbols of competing ideologies. To almost the same degree as the Brahmanical Sanskrit sources, the Pali accounts centre on one lake, not two, evidence for an imagined heaven rather than a worldly site. We cannot simply equate the two, or assume that was the original intention. Although an equation of Anotatta to Manasarovar often seems a reasonable hypothesis, the other-worldly aspect of the references de- nies a consistent identification. In accounts where Anotatta is mentioned in association with the Buddha, its other-worldly aspect is paramount and its earthly location much less grounded. None-the-less, the accounts of water from the lake being highly valued and indeed collected and transported to royal courts, do appear plausible and consistent with later practice at Manasarovar and other Himalayan water sources.50 Similarly,the prosaic accounts of hunters in the mountain realm are not part of any heavenly construct. While they serve literary purposes, their presence seems entirely realistic. But the realistic ele- ments cannot be taken as evidence for more than cultural foundations to leg- ends, strands of reality woven into mythological fabrics. The original distinctness of those mythological cycles is clear in that sets of five or seven rivers/mountains are common to the Pali sources (echoing perhaps, the Vedic seven rivers). In contrast, in the Sanskrit literature mountain references do not appear in standard number sets and references to four rivers are far more common (although as noted there are exceptions such as the Matsya Purana which retains a concept of seven rivers).

50 The water contains numerous minerals considered beneficial. 128 chapter 5

Ultimately, although they informed Buddhist cosmologies in east and south- east Asia, it is important not to overstate the significance of Kelasa/Anotatta in the Pali sources. The canonical references are sparse. Dozens, indeed hun- dreds of other toponyms occur as frequently. The two are not a dyadic com- plex, indeed they are hardly associated. Kelasa is not at all prominent in these Buddhist sources. It principally serves as a metaphor for strength and purity, one taken from a wider cultural warehouse. Its heavenly association is largely implicit and it lacks association with any specific deity. Nor, although it seems to exist in both heavenly and worldly locations, do the references to Anotatta greatly elevate its sacral associations. There is a sacred Himalayan landscape which includes Anotatta, and it can be reached only by those of magical pow- ers, not least that literary clique, the king who is endowed with those powers but who has lost his kingdom. Tribes and hunters form a prosaic background, those excluded are members of Indic society who have caste. It is a land other- wise open only to renunciates. The Pali references, like the Sanskritic, emerge from within existing under- standings, cultural knowns that allow Kelasa/Kailas to serve as a metaphor. They represent combinations of bodies of knowledge and cycles of myths from within that and neighbouring cultures. Although their authorship is elite, there are other voices expressed here which suggest many local variants and regional forms. Sagas and ballads telling of exiled kings utilise sacred geographies pri- marily as appropriate settings. Their toponyms, some meaningful, some mean- ingless, are primarily shaped by literary not geographical demands. Prosaic and embellished accounts, vaguely known locations, stock descriptions, tropes and models, anthropomorphic geographies, and encoded maps overlap with both remembered and visionary landscapes and enable us to glimpse different world-views held by many travellers. If we accept that the Buddha and/or his near contemporaries are the authors of at least parts of the Pali canon, the sources do enable us to gain insights into the world view of a particular community of north Indian renunciates around the 3rd–1st century bce. The meaning of the toponyms with which they encode the landscape may be largely lost to us, but clearly the centre of their world is Varanasi (Kasi/Benares) and the surrounding region embracing the modern Buddhist holy sites. It is actually to these, not Kelasa or Anotatta, that most frequent reference is made. For this community, the Himalayas are a peripheral and rarely mentioned frontier between the manifest and the divine worlds. On one level the Himala- yas are perceived as a land of gods, demi-gods, and renunciates, while on another level they are a land of hunters, tribes, and dangers. (Su)Meru/Neru seems entirely in the divine world, Kelasa might be identified in this world but an early buddhist kelāsa 129 it primarily exists as an other-worldly metaphor for strength and purity and is neither a World-centre nor a focal point of any enhanced sacrality. No par- ticular deity is associated with it and while there is sometimes a sacred lake below it that exists on a liminal frontier, it is the lake which is renowned, not the mountain. Even the lake is just one feature of the sacred landscape, not a major centre or a destination in itself.

Postscript: A Jain Kailas

The Jain tradition recognises 24 enlightened beings (kevalins), the tīrthankaras (‘those who have crossed over’). At least the last of these, Mahavira (c. 540– 468bce), is considered historical. The majority of tirthankaras are described in Jain sources as having attained liberation on a mountain top,51 and Jain associations with Kailas rest on the suggestion that Kailas be identified with Mount Aśtapada, where Riśabha, the legendary first tirthankara, attained lib- eration. Retiring to Astapada with 10,000 monks, Risabha fasted on the moun- taintop for six and a half days before his final liberation,52 whereupon his body was cremated and the remains entombed there in a stupa. The stupa was enclosed within a four-sided palace known as Sri Astapad, or the Ratnamay palace. The identification of Astapada with Kailas is now commonly stated as fact and is part of its wider sanctification as a multi-faith centre in popular media and tourist literature.53 But Jains do not necessarily hold to this belief, rather it is held to be a possibility, for the identification is nowhere confirmed. The Ash- tapad Research Foundation54 has recently endeavoured to identify mountain sites fitting the description of Astapada in Jain texts. They isolated ten possible Himalayan sites of which five, including Kailas, were considered most likely.55

51 42 Other Jain sacred mountains are noted by Jain (1991: 64–65); also see Loseries-Leick (1998: 149–152; although the translation “Spider” is untenable). On Jain mountain cave renunciate retreats, see Flugel (2010: 406–407). 52 Muller and Jacobi (1989: 285, citing Akaranga Sutra). 53 Johnson and Moran (1989: 21) has a photo of Jain pilgrims at Kailas, indicating their modern presence there by the late 1980s. 54 A body formed by the New York Jain centre in alliance with the Lalbhai Dalpatbhai Institute of Indology, a Jain research centre in Ahmedabad. 55 http://www.digambarjainonline.com/news/news13.htm; accessed 21 October 2010. The source of this (and identical accounts elsewhere online) is a scarcely credible report in the Ahmedabad Mirror, 30 November, 2009. This stated that the (New York?) centre had 130 chapter 5

John Bellezza then led a small team to the Himalayas in 2009 in search of evi- dence that Kailas was Astapada. Their work, “yielded absolutely no physical evidence of a Jain monumental presence at Mount Kailas in any chronologi- cal period.” But Bellezza concluded with the suggestion that, “Sri Ashtapad is Mount Kailas itself, rather than a man-made temple of epic proportions or oth- erwise.”56 In discussing a Jain Kailas history we face many of the same problems as with Buddhist history. Again, the creation of a canon came nearly a millennium after the passing of the founder of the faith. By around 466ce., when the last of four councils was held in Gujarat to formulate the canon, the Jains had already split into two main branches over the question of whether particular texts were canonical. In seeking references to Astapada (or Kailas) we find, as with the Pali sources, that relevant material is primarily located in the commentarial literature which developed from the canonical material and which thus dates to at least the latter half of the first millennium ce.57 The Jains, like the Buddhists, to a large extent took on the existing Indic understanding of cosmology. Thus they maintained the existence of a cen- tral world mountain, one they generally called Mandara rather than Meru. But there was no suggestion that Astapada was to be identified with that moun- tain.58 Separate too was Mount Vaitāḍhya or Veaḍḍha, the mountain which would survive the destruction of the world at the end of the Kali yuga (era) in which Jains, like Hindus, believe the modern world exists.59 Again, separate bodies of knowledge seem to coalesce. The Jain texts also indicate that like the Buddhists, they were involved in an encounter with a world in which Nagas were prominent. Willem Bollee cites the possibility that one reason for the popularity of the semi-legendary Tirthankara Pārśva, predecessor of Mahavira, may be that he is associated with Nagas, and interestingly Jain texts do preserve the name of the Naga king

published 19 volumes (spread over 8,522 pages!) on the religious significance and mystery of Astapada mountain. I am, however, unable to identify any such publications in the Ahmedabad centre publication lists, while the New York centre website is apparently inoperative. 56 Report available at http://www.tibetarchaeology.com/wp-content/uploads/2009/07/KM -III-Exploration-Report-JVB-reduced.pdf. Accessed 11 October 2010. Also see Cort (2010). On Sumeru in the Jain tradition, see Flugel (2010: 407–408). 57 On which see Bruhn (1998: esp. 120). 58 Muller and Jacobi (1989: 287) demonstrate the common use of mountains as metaphors in their translation of the Sûtrakritâṅga. 59 Granoff (2006: 44). an early buddhist kelāsa 131

Mucalinda, the lake toponym in the Vessantara-Jataka,60 evidence of some conceptual link to (the Sanskritic) Madros. But there is no support in the early material for Astapada, the liberation place of Risabha, being identified with Kailas and its associated lake or rivers.61 We seem to be dealing with an entirely separate body of knowledge. There is a Kailāsa mentioned in Jain commentarial literature, but it is not Astapada. It is a deity of that name, one of four aṇuvelaṁdhara deities dwelling on a mountain named Kailas which is located in the cosmological Lavana ocean, and Kailas is also the name of the deity ruling over the obscure island of Naṁdīsara. In one reference from the Uttarādhyayana-cūrṇi, a 6–8th century prose commentary, it even seems to refer to a Himalayan mountain, but there is no association with the liberation place of Risabha.62 Again we are dealing with mythical and shifting toponyms that exist in cosmologies, and which may be drawn on as metaphors or as temple models or other artistic discourse and representations. It is also notable that the Himalayas are far from the modern Jain heartland of Gujarat and south-central India. While Mahavira’s field of operations was north India, it was in south-central India that the Jains became established as a significant group. Thus, like the Hindu renunciates from the south who were so influential in construction of upper Himalayan sacred geography, their Himalayas were imagined. That imagination was demonstrated in the literature of the Jains, whose medieval poets adhered to Sanskritic court poetry conventions in which moun- tains suggested heavenly sites for renunciates and demi-gods, with the poets drawing not on reality but on established Indic mythologies and poetical prece- dents. (Phyllis Granoff even cites a 16th century poem concerning a moun- tain which has a commentary, apparently by the poet himself, referencing not nature, but Kalidasa’s Meghadūta as a source!63) By the 14th century, Jain poets were equating worldly and other-worldly mountains,64 and again it appears that it was this poetic or literary representation that dominated the under- standing of the site and eventually enabled the equation of Astapada with Kailas.

60 Bollee (2007: n. 1 [page] 1). 61 Although we might note that http://www.hinduwebsite.com/jainism/thirthankaras.asp, accessed 22 August 2010, identifies one of Risabha’s two “attendant spirits” as “Gomukha” (perhaps Gaumukh?). 62 Malvania (1970: [no page numbers] citing Jīvājīvābhigama 160, 183); also see Schubring (2000: 229). 63 Granoff (2006: 31–50, esp. 36, 38). 64 Granoff (2006: 46). chapter 6 Kailas on the Edge of Modernity

Introduction

The second millennium ce., saw much of the sub-continent come under the control of dynasties of foreign origin, although the impact of these dynasties on different regions of India varied considerably.Under the Mughals and earlier Islamic dynasties pre-modern models of statehood allowed large parts of India, notably in our case the Himalayas, local autonomy in return for acknowledge- ment of Mughal overlordship. After 1858, under the British, India was divided into territory directly ruled by the colonial Government of India and other terri- tory, commonly known as the ‘Princely states’,where again local autonomy was allowed, at least in the cultural sphere. Most of the Indian Himalayan belt came under various forms of the Princely state system and the regions with which we are concerned were, therefore, largely spaces which remained outside of direct colonial rule. The importance of such divisions is often overlooked, with the British naturally pre-occupied with the territories under their direct rule and our scholarship tending to focus on their central archives. Rule by Islamic and Christian governments had considerable impact on Indic beliefs, both in terms of doctrine and philosophy and in terms of defin- ing most of the Indian population as ‘Hindus’ (a term that originally meant only ‘the people beyond the Indus’1). But in those spaces—such as the Himalayas— where Hindu rulers retained autonomy in return for acknowledging Mughal or British supremacy, the impact of the monotheist traditions took a different form. Patronage was a key issue here. While both Mughal and British rulers patronised Hindu festivals and institutions in certain circumstances, Hindu rulers continued much more broadly based patronage systems on traditional models, and thus continued to attract Brahmans to their territory. A notable feature of this was patronage of renunciates, a class generally regarded with suspicion and often outright hostility by Mughal and British authorities. In allowing cultural autonomy under traditional kingship systems in the Himalayan states, Mughal and British government effectively created spaces in which religious continuities could operate. In these regions beyond central control, both tradition and dynamic formulations of the ‘new’ could prosper,

1 The term is a Persian variant of the Sanskrit sindhu, meaning the Indus river.

© koninklijke brill nv, leiden, 2015 | doi: 10.1163/9789004306189_008 on the edge of modernity 133 and in some senses even contest colonial domination. While the central struc- tures of religion in the dynastic centres had to respond to foreign beliefs in order to obtain recognition and patronage, internal dynamics continued to shape religion in these protected spheres. Indeed there is a dynamic history of organic processes in those spheres, a history yet to be properly written. That history would take into account such features as the trans-Himalayan develop- ment of royal chronicles (vaṃśavalīs), Brahmanical immigration into the hills, and the growth of regional sacred sites and associated pilgrimages. As discussed in the Introduction, the encounter with Islam and Christian- ity enhanced monotheist tendencies in Indic beliefs, while to a large extent modern Hinduism was an ‘invention’ of British colonial scholarship, albeit an invention heavily influenced by Brahmanical informants and Hindu mod- ernists. The result of this nexus was the privileging of early Sanskrit texts over later vernacular sources, and a reliance on Brahmanical authority concerning the actual practice of religion in India. When observable reality failed to reflect the religion of the Sanskrit texts, colonial scholarship concluded that this real- ity represented a state of decline and degeneration from an earlier ideal age.2 Developments such as those reflected in Puranic and Tantric literature were seen as evidence for this corruption of the ‘pure, original’ philosophy of the earliest Hindu (and Buddhist) texts. This doctrine was echoed in the Brahmanical belief that the modern era was the Kaliyuga, an age of decline, and many other elements of the colonial con- struction of Hinduism also benefitted the priestly class, particularly those spe- cialising in texts rather than ritualists or renunciates. As Richard King stated, the Brahman caste, “proved amenable to an ideology which placed them at the apex of a single world religious tradition.”3 That tradition, with its imposed unity and hierarchal systemisation, was now understood to date to the Vedas and Upanisads (texts whose archaic Sanskrit could be read by only a handful of Brahmans and Western scholars). The stage was thus set for those tendencies within the Indic world that looked to those early texts to emerge at the cen- tre of reform movements responding to the colonial knowledge bodies and to scientific modernity. In the context of Kailas-Manasarovar, the modern construction of Hinduism had several important consequences. Firstly, it obscured the diversity of Indic belief and practices concerning that sacred site, subsuming the formerly dis-

2 There is a considerable literature concerning this issue; see for example, King (1999a); Dodson (2007); also see, for a succinct discussion and relevant scholarship, King (1999b). 3 King (1999b: 170–171). 134 chapter 6 parate bodies of knowledge into an incongruous unity mirroring the antithetic character of a unified Hinduism. Secondly, it privileged the authority of Brah- man pandits over local religious authorities in the wider definition of that sacred site, encouraging a Sanskritised understanding of Kailas. Finally, in pro- ducing a unified belief system on the Western model, it ceded various claims to knowledge in allowing, to the extent possible, the validity of modern science. Scientific geography was acknowledged as relevant to spiritual enquiry into the nature of a sacred place, with incompatibilities often understood as conceptual problems to be solved. The new Hinduism acknowledged decline and stimulated reform in a num- ber of areas, such as a new emphasis on social service that reflected Christian missionary impulses.4 But most relevant to our enquiries was the develop- ment of a new generation of Hindu renouncers, including powerful individ- uals who—following the colonial teachings—rejected what were now seen as ‘primitive’ or ‘obscene’ aspects of the faith. They admired and accepted Euro- pean scholarship concerning their religion, and sought to reconcile scientific modernity with their fundamental beliefs. As will be seen, these individuals played a major role in the modern construction of Kailas-Manasarovar and related Himalayan sacred space. It was early in the transitional period that two texts emerged that would be fundamental to locating Kailas-Manasarovar within modern Hinduism. Both texts raise difficult issues of authenticity and traditional authority. They rep- resent a systemisation of the traditional Kailas in the modern world yet one, and probably both, were strongly influenced by the encounter with the West. Indeed they are, in critical Western understanding, examples of ‘pious fraud’,5 in that they are modern understandings framed in an ancient format that enabled them to masquerade as traditional knowledge.

The Skanda Purana

In Chapter Three we saw that in those Puranas that refer to Kailas there was no consistent account of the site, or even close association of it with such fundamental modern understandings as the mountain being the residence of Śiva. But one Purana does clearly stand out in this regard, the Skanda Purana, and it is this text to which most modern references to the Kailas-Manasarovar

4 On which see Pinch (2003). 5 The earliest use of this term may be in Gerard (1996 [1840]: 212). on the edge of modernity 135 region refer. In its earliest form it dates to the 7th or 8th centuries, with the oldest surviving version, a Nepali palm-leaf manuscript, dated to 810ce. The importance of the Skanda Purana to our enquiry is that it acquired an ‘open’ status, not only were later versions very different, but from the 12th century onwards many new texts were attributed as khaṇḍas (‘sections’) of the Skanda Purana.6 In its earliest form, this Purana was closely associated with the Pasupata sect and primarily concerned with the activities of Śiva and Durga. Its geographical focus was on the area between Garwhal and Kuruksetra, and centred on Kasi (Varanasi), which it promotes as the most sacred place on earth. The text was probably composed there, “or in a (Pāśupata) centre that had close contacts with this city.”7 The text identifies the three most sacred sites after Kasi as three Himalayan locations; Mahalaya (one of the peaks of Himavat: ie., the Himalayas; and located to the south of) Kedara (Kedarnath), and another peak named Madhyama. Kailas has no particularly exalted status in this account. It remains the “sanctuary” of Kubera, but there may be indications of conceptual change when we read that Kubera installs a golden linga on the Gandhamadana peak to worship Śiva.8 Of Manasarovar there is no clear mention. Kedarnath however, which is not mentioned as a toponym in the Mahabharata,9 rises to prominence in this text, which is one of the earliest known references to the site. Kedarnath is here associated with the release of the Ganges myth, located as the place where “Hara [i.e.; Śiva] himself released the holy water from the mass of his matted hair.”10 Later versions of the text added additional verses linking Śiva to Kailas. These stated that he dwelt there at least on occasion, for it was his favourite resort, and that he ruled his kingdom from there. Kubera was now downgraded to an acolyte, worshipping Śiva at Kailas. But the text also associated Śiva with a Mount Hemakuta and with Meru, which is here ascribed three peaks, with Brahma residing on the middle peak and Śiva and Visnu on the other two. Manasarovar again escapes attention and at one point the text identifies Badarikasrama (Badrinath?) as the holiest of all sites.11 As will be

6 Adriansen, Bakker & Isaacson (1994: 326); also see Adriansen, etc., (1998). 7 Bisschop (2006: 7, 14, 18, 22, 37–41, quotation from n. 285 214). For convenience I give page numbers rather than textual citations, not least due to different versions of the text reproduced by Bisschop. 8 Bisschop (2006: 18, 75, 177). 9 Bisschop (2006: 21); it is used as a common noun in an “interpolation” in Mhb. 17.2.1. 10 Lorenzen (1991: 109 & 173) refers to Pasupatas and Kalamukhas as visiting Kedarnath. 11 Skanda Purana 1.30.32–35, 1.37.32–33, in Bhatt/Tagare (1992: 256, 304); 1.11.37, 1.111.2. in 136 chapter 6 seen, it was only with the linking of a local mahatmya to the Skanda Purana that the text came to include an understanding of Kailas consistent with the modern. What is highly significant, however, is that in earlier versions of the Skanda Purana, as with the entire range of pre-modern literature, there is no suggestion that the Kailas region was actually a pilgrimage site. In the Mahabharata (3.83.88–89), there is a clear distinction, “between accessible (gamya) and inaccessible (agamya) tīrthas, which can be reached by thought (manasā) alone and are occupied by … [various enlightened beings].”12 Thus in the Epics Kailas was visited only by Hero-figures such as Arjuna and Rama. In the early Puranas it is still a realm of various forms of deities and deified spiritual practitioners. This division is continued in the Skanda Purana, where the Nepali manuscript opens with a description of Śiva’s sacred sites that are explicitly stated to be those “accessible to men” and the list does not include Kailas or Manasarovar.13

The Manasakhanda

The textual transformation of Kailas to a site understood to be accessible to ordinary pilgrims only emerges with early modern additions to the Skanda Purana. Other additions such as the Himavatkhanda (also known as the “Nepal Mahatmya”),14 which focuses on the sacred sites radiating out from the Kath- mandhu valley centre, testify to an expansion of pilgrimage centres and net- works into the Himalayas. The Kailas-Manasarovar region is then centralised by the addition to the Skanda Purana of the Manasakhanda, with a Kedarakhanda also emerging which centred on Kedarnath. These texts imply antiquity through their expression in Puranic literary form, but K.P. Malla has shown the Himavatkhanda to be a “pious fraud”

Bhatt/Tagare (1992: 61–80); (This version of the Skanda Purana is later than the edition of Bisschop [2006] and “has as such nothing in common with the original text beyond its name”; email communication from Peter Bisschop, February 2011.) 12 Bisschop (2006: 10). 13 Bisschop (2006: 7, 66, 75–76). But unlike humans, benefits are granted to other-worldly beings travelling there; “[o]n the Kailāsa peak … Trilocana grants the merits of all tīrthas to the Suras and Asuras who see that sanctuary of Kubera.” 14 Correspondence with Jayaraj Acharya (Harvard), 26 May 1996. A 1956 edition by Yogi Naraharinath, Varanasi; Yoga-Pracarinini, is being studied by Professor Hans Bakker and others. on the edge of modernity 137 dating to no earlier than the mid-17th century.15Similarly, Swami Pranavananda observed of the Manasakhanda that:

The author has secured a manuscript copy of Manasa-khanda … Though it claims to be a part of Skanda Purana, in fact it is not. It is not more than two or three hundred years old and is written by some pandit of Almora.16

While Pranavananda testifies to the existence of the text in an Indian language, the Manasakhanda has a curious history and the full text does not appear to have been published or even widely circulated. Pranavananda states that he planned to translate the text (presumably into English), but he is not known to have done so. An English summary of it, however, along with a précis of the Kedarakhanda, had appeared in 1884 in a work by the British-Indian scholar- official Edwin Atkinson ics.17 This summary was actually a “paraphrase” by Sir John Strachey of the Bengal Civil Service. He was a distinguished colonial official who had spent 10 years in Kumaon and Garwhal early in his career, including a visit to Gangotri.18 His brothers Henry and Richard Stachey visited Kailas-Manasarovar in 1846 and 1848 respectively, and Sir John had therefore, considerable knowledge of the region, although it is unlikely that he could translate Sanskrit. In 1850 Sir John Strachey had been appointed by the East India Company to investigate the system of grants in favour of the temples at Badrinath and Kedarnath, and his main informant was an Almora pandit, Rudrapatta Pant.19 Strachey seems to have obtained the copy of the Manasakhanda he gave to Atkinson from Pant. Badri Datt Pande, a leading ‘Freedom fighter’ and author of the standard Hindi-language history of Kumaon also précised the text in that work (Kumaun ka Itihas: 1937). There he cited both Pant and Atkinson as his sources, which suggests that Pande had not seen the original manuscript or even that it was composed by Pant. Pande concluded that the Manasakhanda,

15 Malla (1992: 145–158). 16 Pranavananda (1983: 9). 17 Atkinson (1884: 31–57 [Manasakhanda], 57–84 [Kedarakhanda]). 18 Sir John Strachey (1823–1907), Haileybury, Bengal Civil Service. 1866 Oudh Chief Commis- sioner; 1868 member Governor-General’s council; served briefly as Viceroy after assassi- nation of Lord Mayo 1872. 1874 Lt-Governor North-Western Provinces; 1876–1880 Finance Minister Government of India; 1885–1895 member Council Secretary of State for India. Author of several works including India. Its Administration and Progress, London, Macmil- lan & co, 1903; see 38–39 of that work for his experiences in Kumaon and Garwhal. 19 Atkinson (1884: 3, 85). 138 chapter 6

“appears to be a modern work composed much earlier than [i.e.; since] the time of Shankaracarya … composed in the time of the Chands or sometime before them … in Kurmanchal [Kumaon] by Kurmanchali pandits.”20 F.W. Thomas, the Tibetan specialist then serving at the India Office Library in London, also had doubts concerning the text. He stated that;

For the Mānasa-khaṇḍa … Atkinson appears to be the only authority, either English or otherwise. I have found only one reference to a Mānasa- khaṇḍa—there exists in India a Rāmacīlamāhātmya professing (but such professions are very untrustworthy) to belong to the Mānasa-khaṇḍa of the Skanda Purāṇa. But no such Mānasa-khaṇḍa appears to exist in either manuscript or print. Atkinson’s version must have been made especially for him from some unknown manuscript.21

Atkinson was aware of the status of the Manasakhanda. In a footnote he ascribed “a post-Musalman period for this composition.”22 In British colonial understanding this meant post-1757, and while in the indigenous understand- ing the Puranas are not a closed canon but allow an organic process of continu- ing revelation, British officials, with their ‘Protestant presuppositions’ regarded the creation of additional Puranas as forgeries.23 Beyond that brief footnote Atkinson however, made no comment on the antiquity or validity of the Man- asakhanda. His version is thus largely situated within the ‘traditional’,but made more accessible to the Christian-educated reader by the translation holding to a Biblical tone through the use of such terms as ‘thee’ and ‘thou’. The Manasakhanda, the source of the most-quoted references to Kailas- Manasarovar, is therefore (if indeed it exists in an original form), a compara- tively recent work dating to between the late 17th century at its earliest and more probably the late 18th or even early 19th century. It is also a text which, while it may reflect Chand agency, in its promulgated form bears a heavy colo- nial imprint.

20 Pande (1993: 144). Joshi (2009: 329) states that Pande’s work largely relies on Atkinson. 21 Correspondence cited in Hedin (1916–1922: vol. i, n. 4 13–14). 22 Atkinson (1994: n. 2 37). 23 e.g. Oakley (1905: 144) notes that, “[o]ne of the early Anglo-Oriental scholars is said to have detected his own pandit (or Sanskrit tutor) forging some new chapters to these books; and no doubt large portions of the Puranas, and especially this Skanda Purana, have but little claim to antiquity.” This probably refers to Captain Wilford (see Chapter 15), but could even refer to the Manasakhanda by Atkinson, who is cited by Oakley. on the edge of modernity 139

The Text

The Manasakhanda begins in the usual manner of a Purana with an account of the creation of the world. In this case while the account has a Vaisnavite gloss it also presents a pantheon more characteristic of Hindu modernism, with the trinity of Brahma, Śiva and Visnu located respectively on the Vindhyachal, Kailas and Himachal mountains. Again, internal consistency is lacking, Śiva is later stated to reside on Himachal but the question is clarified when the rsi Dattátreya visits Siva and Parvati at Kailas and asked;

Which is the greatest of mountains and where do you live yourself, and in the earth which is the most sacred place? … Siva answered “I dwell everywhere, but Himachal is my peculiar seat and on every one of his peaks I dwell for ever and on the mountain of Nanda dwells Vishnu, and I and Brahma also. There is no other mountain like Himachal; look upon him and receive whatever you desire.”

Siva’s association with Kailas is thus open. He dwells on Kailas, Nanda, and Himachal, the latter simultaneously a divinity and a mountain range, and yet also a specific mountain. But ultimately he is present everywhere. Later in the text however, the pre-eminent sanctity of Himachal is reiterated in verses that are the most quoted reference from ‘traditional’ sources in accounts of Kailas. These lines appear in some form in virtually every modern account of the site, and following Sherring (of whom more later), they are frequently misattributed to the Ramayana. So common is that misattribution that it seems almost deliberately intended to strengthen the authority of the quotation by rooting it in the demonstrably ancient verses of the Epic, though the Biblical tone of its expression is again prominent. The lines appear after the return of Dattatrya from Kailas to Kasi. The Kasi raja enquires of him as to the greatest of tirthas. The rsi replies that;

He who thinks on Himáchal, though he should not behold him, is greater than he who performs all worship in Kashi … In a hundred ages of the gods I could not tell thee of the glories of Himachal … where Siva lived and where Ganga falls from the foot of Vishnu like the slender thread of a lotus flower and where the Rishis worship … When the earth of Mánasa-rovara touches anyone’s body or when anyone bathes therein, he shall go to the paradise of Brahma, and he who drinks its waters shall go to the heaven of Siva and shall be released from the sins of a hundred births … There is no mountain like Himáchal, for in it are Kailás and Mána-sarovara [sic]. As 140 chapter 6

the dew is dried up by the morning sun, so are the sins of mankind dried up at the sight of Himáchal.

As we have seen, Atkinson’s text is actually a “paraphrase” by Strachey. He is not known to have been a Sanskritist although in line with the requirements for British colonial officers he would almost certainly have passed examinations in the official vernacular language of his district. Strachey’s pandit Rudrapatta Pant appears to have been a key figure in the transmission of the text, partic- ularly if it was in Sanskrit rather than a local language. But it is clear from the wording of the quotation, with its “paradise”, “heaven”, and “sins”, that the ver- sion Atkinson gives is one heavily influenced by Christian concepts (and per- haps even copied by Stachey from an oral source). But given that the text then circulated in the English language, and that the verses above are fundamental in the modern construction of Kailas-Manasarovar, the shadowy ‘original’Man- asakhanda becomes less important. It is an imagined text rather than a reality, for it was the English version of the text that became the standard edition. Other important aspects of the modern understanding of the sacred Kailas centre also emerge from Atkinson’s text. It refers to Manasarovar as “the sacred lake”, and repeats the origin myth of its being created from the mind of Brahma at the request of the rsis practicing at Kailas. It also describes the Ganges as descending from the foot of Visnu (not Śiva) to Kailas and thence Manasarovar. But it is also the earliest Indic work in which we can definitely identify the textual Kailas with the mountain in western Tibet and also understand the site as one which is open to ordinary pilgrims. For if the mountain geography is not entirely fixed in the case of the multiple identities of Himachal, and the account of the Ganges’ origins is traditional rather than scientific-modernist, the text specifies an itinerary for pilgrimage to Kailas in which its modern location is clear. The route is briefly described as being that which follows the Kali river and crosses the Himalayan watershed at the Táraka-dhura (Lipu Lekh?) After bathing at Manasarovar the pilgrim should “look on Kailas”, bath at “Ráwan- hrad” (which is identified in a footnote as Rakas Tal), then return via the Karnali route and Mala (near Badrinath). There are several notable features of this itinerary. Firstly, it implies that the pilgrim will depart from Almora, which is consistent with its creation by an Almora pandit. Secondly, it instructs the pilgrim to travel and to return via Kumaon, thus retaining the pilgrims’ Indian expenditure within British rather than Princely state territory. Finally, having detailed an itinerary that closely follows the geography of the Kailas- Manasarovar region, the text specifies that from Mala the pilgrim should visit “Jwála-tirtha” (Jvalamukhi)—in far-off Kangra. “Thus”, it states, “is the pilgrim- on the edge of modernity 141 age completed.” As we have seen in Chapter 4 this reference to Jvalamukhi as the final site on the pilgrimage itinerary seems entirely out of place here, its sanctity part of a very different chain of meaning. Strachey/Atkinson make no comment on it, and it seems to represent neither the interests of the colonial state or of the Almora pandits, and to be explicable only in terms of acknowl- edgement of a wider body of Tantric understanding to which the text (or at least this précis) makes no other reference. Again, it is important to contextualise the references to Kailas-Manasarovar in the Manasakhanda. Numerous other sites in this region of the Indian Hima- layas—Nanda Devi, Dunagiri, the Kosi river, and so on—are also described, usually but not always in geographically logical terms, and many are eulogised and related to various mythological cycles. While the Kailas-Manasarovar com- plex is central, the sanctity of the entire area is emphasised. As noted, Atkinson’s exposition of the Manasakhanda was followed by the Kedarakhanda, which he obtained in manuscript form from Srinagar and Al- mora pandits (Ganga Datta Uprethi and Dharmanand Joshi). He followed an Almora version stated to have been “copied” in 1816, presumably by the pan- dit. Atkinson omitted the bulk of the text comprising myths and legends, and thus described his version as “little more than an index to a portion of the Kedarakhanda.”24 This extract describes pilgrimage sites on and around the Mandakini, Jumna, Alakananda and Bhaigirathi rivers, Badrinath, Kedarnath, Rudraprayag, Devap- rayag, and Badrinath, along with other local sites. Mention of places such Gan- gotri or Gaumukh at the Bhaigirathi source—in Princely state territory—is notably absent (at least from this précis of the text), although Atkinson identi- fies a “Ghosheswar temple” as being at the junction of the Mana-rudta stream with the Jadh Ganga above Nilang. The identification is dubious, for no temple is otherwise recorded above Nilang,25 and the description is entirely inconsis- tent with the reality of that area. How such an apparently important sacred site as Gangotri can be omitted from even an abbreviated version of this text will be discussed later. Here it is sufficient to emphasise that these local mahatmyas, while following the tradi- tional Puranic model, are of no great antiquity, and while more geographically realistic than the phantasmagorical accounts of the region in many Puranas, are perhaps most important in indicating that the region was open to ordinary

24 Atkinson (1994: n. 5 59). 25 The course of the Jadh Ganga is within a closed military area and I have been unable to access this region. 142 chapter 6 pilgrims. The understanding of it as the home of the gods and thus a place inac- cessible to men had been fundamental since the time of the Epics. Now that prohibition was repealed. The local agents of that transformation are difficult to identify beyond rep- resentatives of the Almora religious and political establishment. That they would seek to stimulate a sacred geography embracing their territory, as a gateway if not a centre, is consistent with practice elsewhere and there is no direct evidence for colonial government agency. But pilgrimage to Kailas via the Kumaon routes (rather than through the Princely state of Tehri-Garwhal) was in British interests in that it retained resources within British territory and the Pant-Strachey-Atkinson nexus did promulgate textual support for pilgrim- age to Kailas-Manasarovar by that route. The British understanding of Kailas, particularly as it emerged in the work of Charles Sherring (see Chapter 15), was influenced by their presence in Kumaon. Indeed the region provided considerable input into the ways in which Himala- yan religion and society came to be understood. As a non-regulation district of British India, its Commissioners enjoyed considerable authority and also served long terms in office. There were just three in the first 70 years of British rule, which tended to favour greater understanding of local perspectives. The administrative centre of Almora was developed as a hill-station that, like others of its ilk, was a potential stepping-stone to Tibet. It was already a base for cross- border trade and many of the 19th century travellers who entered western Tibet did so via Almora.26 The town also provided a congenial home for imperial offi- cers such as Atkinson, who undertook a study of the temples there,27 and was thus well-known to the local Brahmanical community that was apparently the source of the Manasakhanda. There is another political context in that the area had been under Gurkha rule for a generation prior to the Anglo-Nepali war of 1815, following which the western Himalayas were restored to local rulers under British suzerainty. That local elites made efforts to restore their religious authority and to stim- ulate the economy through encouraging pilgrimage seem predictable. But the importance of British in encouraging the pilgrimage to Kailas was to become a major part of the modern history of Kailas-Manasarovar. We examine that aspect in Chapter 15, here, however, we turn to the Kailas of the modern Tantric

26 See Snelling (1990: 93, 104) re travellers; Lamb (1986: 198) on the ‘forward policy’ of the Almora dc; on 1881–1905 applications to travel in western Tibet generally, see oioc, l/p&s/10/186, report of 11 January 1906. 27 See Atkinson (1884). on the edge of modernity 143 textual tradition, which also demonstrates the existence of other, independent bodies of knowledge concerning the site.

The Mahānirvāṇa Tantra

A realistic Kailas-Manasarovar, one grounded in geographical reality and shorn of fantastic proportion if not of legend, emerges in classical textual literature only in the colonial period, with Atkinson’s version of the Manasakhanda. Similarly it is in the Mahanirvana Tantra that the picture of Kailas as the mythical paradise of Śiva is fully developed. The text opens with a paean to Kailas:

The enchanting summit of the Lord of the Mountains, resplendent with all its various jewels, clad in many a tree and many a creeper, melodious with the song of many a bird, scented with the fragrance of all the seasons’ flowers, fanned by soft, cool, and perfumed breezes, shadowed by the still shade of stately trees … It was there that Pāravati, finding Shiva, Her gracious Lord … beloved Master of all yogi, Whose coiled and matted hair is wet with the spray of the Gangā, and (of Whose naked body) ashes are the adornment only; the passionless One, Whose neck is garlanded with snakes and skulls of men, the three-eyed One, Lord of the three worlds …28

As the first translated Tantric text to be published in Europe (in 1913), the Mahanirvana Tantra was widely read, not least in India where it was popular with the English-educated middle class.29 In the introduction to the text, the sanctity of Kailas was further stressed:

Here in these lofty uplands, encircled with everlasting snows, rose the great mountain of the north, … In these mountains Munis and Rishis lived. Here also is the Kshetra of Shiva Mahadeva, where His Spouse Par- vati, the daughter of the Mountain King, was born, and where Mother Ganges also has her source. From time immemorial pilgrims have toiled through these mountains to visit the three great shrines of Gangotri, Kedarnath, and Badrinath. At Kangri, further north, the pilgrims make

28 Avalon (1913: 1–4). 29 Taylor (2001: 129, 136). 144 chapter 6

the parikrama of Mount Kailasa (Kang Rinpoche), where Shiva is said to dwell. This nobly towering peak rises to the north-west of the sacred Mansarowar Lake (Mapham Yum-tso) … The paradise of Shiva is a sum- merland of both lasting sunshine and cool shade, musical with the song of birds and bright with undying flowers. The air, scented with the sweet fra- grance of Mandara chaplets, resounds with the music and song of celestial singers and players …

This Tantra, however, which was published in English as The Tantra of Great Liberation, was not an ancient text. There is no mention of it prior to the late 18th century and although its English editor believed it to be of considerable antiquity it is now generally accepted as a late 18th or early 19th century Sanskrit work from Bengal, one reflecting the reformist tendencies of Ram Mohun Roy and the Brahmo Samaj. It ultimately held to an abstract, almost monotheistic, ultimate principle and omitted the esoteric rites characteristic of earlier Tantras, with their explicit sexuality and sanctified violence. As such it was a modernist Hindu text, one consistent with Vedantic ideals.30 As the work of Kathleen Taylor has demonstrated, the editor of the (greatly abridged) English text, Sir John Woodroffe, was not a Sanskritist. He relied on the translations of a local collaborator, Atal Bihari Ghose. Ghose was himself well-educated and Woodroffe, who used the pseudonym of Arthur Avalon in publishing this work, was a high court judge in Calcutta. He was also part of a wider intellectual circle which included Theosophists and others who were interested in the mystical traditions and spiritual wisdom of India and who sought through their enquiries to counter some of the negative images of Indian religion held in the West.31 His contacts included the Indian art historian E.B. Havell, and the aristocratic Sikkimese Buddhist monk, Kazi Dawa Samdup, translator of the Tibetan Book of the Dead and Woodroffe’s primary informant on Tibetan Tantra. Woodroffe’s Tantra was a comparatively recent work bearing the imprint of Hindu modernism, and its depiction of Kailas emerged from that worldview. So too did Woodroffe’s understanding of Kailas as expressed in his editorial commentary arise from modern constructions, for his introduction cites the famous lines from Atkinson’s text noted above: “He who thinks of Himachala [sic] …, is greater than he who performs … worship in Kashi … As the dew is dried up by the morning sun, so are the sins of mankind by the sight of

30 Taylor (2001: 137); also see Urban (1995: esp., 66–67). 31 See the excellent biography of Woodroffe by Taylor (2001). on the edge of modernity 145

Himachala.”32 By 1913, when the Tantra was published, this understanding of Kailas was common, for as we will see in Chapter 15 the region had become a site of considerable interest to Europeans during the course of the 19th century, and in 1906 it had been systemised and skilfully promoted in an influential English-language publication.

The Modern Textual Tradition

The primary Indic textual sources for the modern understanding of Kailas, the Manasakhanda and Mahanirvana Tantra present very different understand- ings of Kailas to that in earlier texts. It is in these texts that the modern Indian understanding of Kailas-Manasarovar as the abode of Śiva and a site for ordi- nary pilgrims to worship him first fully emerges. Neither can be precisely dated and the circumstances leading to their creation cannot be properly established. It is possible that they represent a late flowering of Sanskrit literature,33 and are indicative of local developments, particularly the Sanskritisation policies of the Chand dynasty (on which see Chapter 9). The hill kingdoms were rela- tively powerful courts, and the texts are consistent with a claim to space beyond Mughal control. But they represent major developments in, and systemisation of, the understanding of Kailas-Manasarovar. The clear statement as to Śiva’s residence at Kailas that was made in the Mahanirvana Tantra may simply rep- resent a development of earlier ideas and it is in Woodroffe’s introduction that these ideas are amplified. None-the-less, this does indicate that by the late 18th or early 19th century the Kailas-abode-of-Śiva model had been applied to Kailas-Manasarovar. The most likely agents of this concept were the Śaivite renunciates who ventured there, but Kumaon state interests and those of the Almora pandits played a significant role in promulgating this understanding. As will be seen, the Manasakhanda became the primary textual source for the European construction of Kailas, despite its dubious authenticity. The Mahanirvana Tantra was of greatest significance in representing a purified form of Tantra to a Western audience, but was also widely read by those who shaped the 20th century understandings of the site. By the time these texts emerged into the public domain, a systemised Hindu religion had been con- structed, and in line with that construction sacred space was also in the process

32 Avalon (1913: 1). 33 Sheldon Pollack (2001) describes the dynamic and innovative character of Sanskrit litera- ture in the two centuries prior to the mid 18th century. 146 chapter 6 of being systemised. But that Kailas-Manasarovar emerged as a supreme site was, as will be seen, the result of specific and identifiable colonial agency. We must also note the near contemporary promotion of Kedarnath and the Mandakini river as a sacred centre. Its appeal was to be subsumed under that of Kailas-Manasarovar, but its contemporary claim to sanctity seems no weaker, indicating the contested nature of regional sacred centres and the importance of British colonial agency in the construction of a modern Kailas. Similarly, it is notable how little input to this understanding derives from Tibetan sources. While Kailas-Manasarovar was in Tibet politically, it was in the Indian conception within Indic sacred space. section 2 The Kailas Mountains of India

chapter 7 Above the Naga Lakes: Kaplaś Kailas & Manimahesh Kailas

Introduction

Kailas-Manasarovar may now be considered Asia’s pre-eminent sacred moun- tain, but in the Indic world the concept of a mountain abode of Śiva known as ‘Kailas’ is not restricted to a mountain in Tibet. There are (at least1) five other such mountains in the western Himalayan regions of modern India; the Kaplaś, Manimahesh, Kinnaur, Adhi and Sri Kailas mountains. The study of these sites draws attention to the fact that in both historical discourse and popular understanding the western Himalayan region is over- whelmingly contextualised as peripheral to some greater power. Whether the cultural centre discussed is the Kushan empire, Vedic, Epic, Hindu, Mughal, or colonial India, Zhang-zhung, Guge, or Buddhist Tibet, the region is almost invariably seen through the perspective of states or cultural empires. This reflects, of course, Western academic preoccupation with centralised polities, the reading of history as a lineal progression towards increasingly centralised empires, interspersed with fragmented periods of ‘darkness’ between epochs of elite central control.2 It also reflects the perspective of modern India with administrative divisions and regulations (e.g. Inner Line permits) enshrining the peripheral, albeit spiritually hallowed status of these regions in the Indian nation.3

1 Neel Kanth, in Lahaul Spiti, is referred to as “Neel Kanth Kailas” in some modern literature; also see Madden (1848) who refers to a “Kylas” mountain apparently south of Almora, but I am unable to comprehend his geography. Tucci (1977: 27) notes a Kailas range northeast of Gilgit of which Rakaposhi is the highest peak. As this is now within Pakistan any Hinduisation processes have presumably ceased. The Indian Kailas mountains discussed here are those where the toponym Kailas is specified on maps or official publications. 2 That perspective is now being challenged, see for example Khazanov & Wink (2007); Beck- with (2009a). 3 That peripheral status does not imply neglect in modern India. The Himalayas attract numer- ous pilgrims, and strategic considerations ensure the presence of Indian Army personnel along with state activity aimed at development. The great majority of the population are from the higher castes, and economically the hill regions are among the wealthier Indian states.

© koninklijke brill nv, leiden, 2015 | doi: 10.1163/9789004306189_009 150 chapter 7

Yet the ‘isolation’ of these places was relative. Not only were there organised systems for rapid communications from political centres to their peripheries (such as the ‘arrow letters’ in the Tibetan empire), but numerous individuals and groups regularly traversed the Himalayan regions. What did not travel eas- ily across the mountain realms was external political authority. The Mughals, for example, whose cavalry forces were orientated to warfare on the plains, were unable to exercise direct authority in the mountains, and even British power in the Himalayas was largely mediated through local elites rather than imposed directly from the colonial centre. What was characteristic of the Himalayan periphery was the latent devel- opment of modern state structures and processes. State formations retained a traditional character down to the British (and in the case of Tibet, the Com- munist Chinese), colonial period, with aspects such as summer and winter capitals, distinct religious and political centres, and imprecise external bound- aries in which sovereignties merged or even overlapped. In that model of state- hood there was none-the-less a centre, the physical body or symbolic presence of the king (usually representative of or closely associated with the regional deities), and there were replica centres around lesser regional rulers. Thus the model of statehood may be conceived of in terms of overlapping or even three- dimensional mandalas of power, or as “galactic polities” on the model famously proposed by Stanley Tambiah.4 From the perspective of the periphery,a separate identity and culture existed from that of the centre, indeed ‘peripheral’ Himalayan polities themselves claimed centrality, although factors such as limited resources mitigated against the wider or lasting acceptance of these claims. Thus while analysis focussed on the perspectives of centralised empires enables a centre-periphery division to be sustained and its processes demonstrated, the frontier model of historical processes and agencies seems to better represent the perspective of the periph- ery by locating it within dialectical historical processes.5 The western Himalayas were a cultural region in themselves. Given its fron- tier character the exact extent of that region is best left imprecise, but that referred to here is from the Kali river in the east (which forms the modern Indo- Nepali border), westwards through Kumaon and Garwhal, and northwestwards in the arc that includes the former Punjab Hill States, Kangra, Chamba, Kulu, and so on, while—for our purposes at least—ending with the Kashmiri realm.

4 Tambiah (1985). 5 I have discussed the Himalayas as a historical frontier in McKay (2009); also see Lewis (1994: 25–46). kaplas and manimahesh kailas 151

The region thus has historical frontiers with—or extensions into—what are now Nepal, Tibet, Kashmir, and in the south, the Indian plains’ kingdoms. Historically, no single centre ever dominated this region6 (although the Gurkha kingdom attempted to do so), and that lack of centralisation has de- terred scholarship. But historically its internal politics and processes are broad- ly consistent with the model of the “galactic polity” in which centres rose and fell in prominence. Culturally, it was the Sanskritic model that was increasingly followed in western Himalayan royal courts. But the institutional power of the indigenous territorial deity system, which devolved considerable authority to the local level could, as we will see in the case of Bushahr/Kinnaur, act as a brake on the development of the Sanskritic model of strong central kingship. Of particular importance to us is that the ‘centre’ perspective assumes that the people of what is now the Indian Himalayas (who were/are not Buddhist, Sikh or Muslim), may be classified as ‘Hindu’. The problems involved in the def- inition of that term are well known and its use in any but the most general sense for the pre-colonial period is best avoided without some qualification. If we place the western Himalayas at the centre of our enquiry the problems of that definition are obvious. Expression of defining elements of Classical and later Hindu thought, as well as structural elements such as Brahman functionaries at Sanskritic temples, arose much later here than in the Punjab or the Gangetic valley and as was the case in Tibet with Buddhism, the World-religion only slowly permeated down from court level to wider acceptance. Throughout the region it was belief in distinct household, local, and territorial deities, and par- ticularly the Nagas, that was historically pre-eminent. These deities were only slowly absorbed into the Sanskritic pantheon.

6 Extreme claims have been made (e.g. Chandola 1987: 27) for the extent of the Khasa Malla kingdom which existed in various forms from the early 12th to the late 14th century. From its twin capitals at Dullu and Sinja (Jumla) in western Nepal, it expanded to include territory on the Tibetan plateau to the south of Manasarovar and by the 3rd quarter of the 13th century under King Aśokacalla its western extent included the Garhwal Tehri. On the Malla, see Tucci (1955, and esp., 1962) and for the Tibetan sources Vitali (1996); Jackson (1976). More recently an important contribution has been made towards our understanding of the western Mallas by Lecomte-Tilouine (2009: esp. 177–183, 266), who describes a “territory of the flames” sacralised under the Mallas, which conceptually embraced Kangra’s Jvalamukhi flame temple in addition to the natural flaming gas sites of Dullu and Mukhtinath. But despite a 13th century Dullu epic claiming Sirmaur (hp) as tributary, and numerous cultural parallels with what I have termed the Western Himlayan Cultural Complex, claims of Malla authority west of Garhwal lack supporting evidence, while the claim that the Malla kingdom be termed an empire also requires further definition. 152 chapter 7

To suggest that our region was distinct from the Hindu is historically consis- tent with indigenous religious categorisation. That the hill regions were recog- nised as holding to a distinct belief system(s) is clearly stated in for example, the 1857 proclamation attributed to the last Mughal Emperor, Bahadur Shah. In calling for revolt against the British, the proclamation stated that the revolt was a religious war for all of the inhabitants of the sub-continent, specifically including the following categories; Nepalese, Hindus, Muslims, Sikhs, “men of the eastern countries”, and “natives of the Himalaya hills.”7 Thus despite a Sanskritic overlay at court level, separate belief systems were maintained in the Himalayas down to the modern period. One implication of this was that the hills were a primary target for missionary endeavours by the Indic-derived traditions, Hinduism and Buddhism. (That understanding is echoed in Christian missionary activities, which have historically focussed more on tribal communities than the dalit, or ‘untouchable’ classes.) But the sheer weight of references to the Himalayas in the classical literature, however imprecise, indicates that Sanskritic and local traditions were in communica- tion, however occasional, from the first or even the second millennium bce. While there were discreet bodies of knowledge held by different social groups, there was also a complex process of inter-penetration typical of a frontier zone, with Sanskritic deities absorbed into local pantheons and local conceptions informing the Sanskritic.8 The more remote cultural centres were from the exposure brought by trade and pilgrimage routes the longer they survived as independent entities into the modern world. Thus sites which were exposed to Sanskritic influence from at least the mid-first millennium ce., such as Brahmaur in the Chamba valley, and those such as Ladakh which were exposed to Buddhist influence from the same era onwards, are today culturally dominated by Hindu and Buddhist understandings. In contrast, as will be seen, sites more remote from Indic and Tibetan cultural centres and less economically and strategically significant, such as Bhadrawah and Kugti, retained a primarily local cultural orientation into the late 20th century. In the modern Indo-Tibetan frontier regions there still exist distinct religious phenomena that survive from the spiritual world that existed there prior to the introduction of Hinduism, Buddhism and systemised Bön. Their survival testi-

7 Quoted in Quraishi (1997: 246). Emphasis added. 8 See for example, Moran (2013) concerning the incorporation of Ayodhya deity Raghunath into the Kulu local deity system. We may consider that in addition to aspects passing to and from systems, deities might arise separately in both local and Sanskritic traditions and only later be (partly) reconciled; see this work (n. 47 211) re Gangama. kaplas and manimahesh kailas 153 fies to their continuing relevance to the local people, who consult the deities on day-to-day as well as important issues, and to the extent to which local beliefs are interwoven with local identities, social system, and relationship to the sur- rounding environment.9 Those local beliefs, which are conceptually distinct and diverse in worship and ritual forms from those of the World-religions, are simplistically classified in elite discourse by such terms as ‘village Hinduism’, ‘local Buddhism’, or ‘early Bön’. Yet their ritual and conceptual core is indepen- dent of those systems.10 We can, therefore, speak of a distinct regional cultural system, the Western Himalayan Cultural Complex. The term is used here to define those elements of the regional culture that predate, fall outside, or have survived the impact of World-religions. This is not to deny that certain cultural elements are specific to either Indic or Tibetan worlds; sacred bathing for example, is characteristically Indic. As process, however, the most significant narrative was one in which systemised World-religions, supported by the resources of strong centralised states, over- laid the Western Himalayan Cultural Complex through a wide variety of sophis- ticated missionary strategies. In particular, their command of ideological repro- ductive technologies and more-highly developed economic and political struc- tures enabled them to absorb Western Himalayan culture into their own sys- tems. This spread of World-religions into the Western Himalayas involved two overlapping stages. Firstly came the initial frontier encounter, which was char- acteristically through the agency of renunciates operating at both elite and popular levels. Second came the Sanskritisation or Buddhacisation processes. This saw text and mythology production by the new priestly classes with royal patronage of structural developments such as temples, along with the intro- duction of specific elements of the World-religion within those structures (i.e.; caste status and purity concepts in Sanskritisation). This latter phase also saw the semi-formal incorporation of indigenous elements into local manifesta- tions of the World-religion. In these chapters the individual histories and cultures of each Kailas moun- tain will be discussed both as a centre and as a frontier or periphery depending on context, with the understanding that the modern Hindu, Buddhist and Bön religions are not only imported into this world but are distinct from their earlier manifestations, a distinctness partly arising from their on-going dialogue and

9 On which see Moran (2013); Berti (2009); and Berti (2009a). 10 Berti (2009: 15) makes the important distinction that deities introduced into the (Kulu) region do not enjoy a specific relationship to territory (hār), and cannot be accessed through mediums (but see n. 8 above). 154 chapter 7 encounter with the local traditions of the Western Himalayan Cultural Com- plex.

Kaplaś Kailas

The 14,241 foot (4,341 mtrs) Kaplaś Kailas peak is part of a mountain chain above Bhadrawah, a town 120 miles (200 kilometres) east of Jammu in Jammu and Kashmir state.11 Bhadrawah is also easily accessible from Chamba (Himachal Pradesh) via the Padari pass. Although no significant rivers originate there, Kaplas kund (Paharī: ‘lake’), is located at around 10,000 feet (3,050 metres) on the mountain side. The history of pre-colonial Bhadrawah is obscure. Tibetan sources record their 8th century monarch Trisong Detsen sending troops to recover a statue from what may have been Bhadrawah,12 but otherwise the oldest dateable reference is an 11th century Chamba copper plate inscription. In the mid-15th century a local dynasty was founded, with the first historically verifiable king, Nag-Pal,13 mentioned in an inscription from around 1584. During the late 18th century Bhadrawah became part of Chamba, but under the 1846 Treaty of Amritsar (which ended the first Anglo-Sikh war), it was gifted by the British to the Raja of Jammu in return for his support in that war and subsequently became part of Kashmir.14 During the colonial period Bhadrawah attracted few European visitors (the artist, G.T. Vigne in 1839 being the first). With peaks too

11 At the time of research (2004), Bhadrawah was effectively a war-zone due to the Kashmiri separatist insurgency. On visiting Bhadrawah we were strongly advised against undertak- ing the yatra route due to the presence of Pakistani militants in the forest zone between the town and the mountain, although we were given a heavily armed police escort to visit the lower sections of the route. Research was thus restricted, but information was obtained from the pilgrimage organising committee, local religious functionaries and local people. We were also given a copy of a video taken on the pilgrimage (2003?). 12 See note 37. 13 The founding of an earlier dynasty by one Jobnath is mentioned in royal chronicles, with the latter part of the name suggesting Nath influence, but the text links that dynasty to the (legendary) time of the Pandavas; Hutchinson & Vogel (2000: vol. ii; 615–619). 14 Rose & Hutchinson (1996: 104, 106); also see Hutchinson & Vogel (2000: vol. i, 104–108). Bhadrawah has remained under Jammu since that time, but is still predominantly Hindu, and its main cultural and economic links were with the Chamba valley until the Kashmir troubles started in the 1980s. There was even a jeep road across the mountains, but that is now abandoned and Bhadrawah is cut off. The Gaddis can still cross the mountains, but at the risk of the militants ‘taxing’ their flocks for food. kaplas and manimahesh kailas 155 low to excite mountaineers and situated far from Kashmir’s political centres or major trade routes, Bhadrawah remained isolated. Its history and culture are almost entirely unstudied and in recent decades its involvement in the Kashmir separatist conflict has accentuated that isolation. While Nag Pal’s court was Sanskritised, his name is a reminder that the ear- liest ascertainable local beliefs centred on the worship of Naga (colloq. Nag). This faith predates the presence of World-religions, not only in the Himalayas but throughout most of mainland Asia. Early conversion myths in both Hin- duism and Buddhism commonly involve status shifts in the role of the Nagas, shifts that are not necessarily acknowledged at local level. In this case the local royal Chronicle (Skrt: vamśāvalī) traces the origins of dynastic rule before Nag Pal to Vasuki-nag, (or Basukinag; hereafter, Vasukinag). Vasukinag is in some accounts not only the first raja of Bhadrawah,15 but is also the presiding ter- ritorial god, the patron-deity of the state.16 (This overlap between Nagas and territorial deities is common across the region; Nagas are sometimes territorial deities, but not all territorial deities are Nagas.) Despite Vasukinag’s status he is not a native of Bhadrawah. Widely known in some form throughout the western Himalayas, often as the king of the Nagas or their underworld realm, he is said in this case to have come from Kashmir hotly pursued by the Nagas’ traditional enemy, a Garuda.17 Accounts of subsequent events vary slightly.According to the royal Chronicles Vasukinag’s ascendency is attributed to the Sanskritic goddess, Bhadrakali/Kali, identified there as Vasukinag’s sister. She took pity on Vasukinag and granted him her own earlier power over Bhadrawah when he took up residence in a lake on the summit of Kaplas.18 Another version of the tale has Vasukinag taking refuge in the lake, only for the Garuda to damage the lake shore so that the water began to drain out. Vasukinag then prayed to another Sanskritic deity, the goddess Sarasvati who resided in a higher lake on the mountain. She allowed the waters of her lake to replenish his.19 Both versions of the myth and particularly the Vamsavali’s account of an earlier period of rule by Kali seem designed to link Vasukinag to the Sanskritic Hindu tradition,20 but less weighted and perhaps earlier versions also exist. In one prominent account Vasukinag had a companion, Jitman-ji, who allowed

15 Hutchinson & Vogel (2000: vol. ii, 616). 16 Handa (2004: 102). 17 On Nagas and garudas, see Krishna (1980: 72–78). 18 Handa (2004: 102). 19 Ganhar (1975: 65–66). 20 On which see Sharma (2008), for an example from Kashmir. 156 chapter 7 the Garuda to feed on his flesh to enable his friend to escape. Thus Jitman-ji is depicted standing alongside Vasukinag in the central altar position at Ghata, the main Vasukinag temple in Bhadrawah today.21 Alternatively, in a local version of a common Himalayan myth, the Garuda is said to have tried to drink the lake dry then vomited it up when told by the gods that the water-supply was endless.22 The account of Vasukinag’s flight from Kashmir has been interpreted as reflecting the expansion of Vaisnavism in Kashmir at the expense of Naga worship,23 and the movement of Naga cult followers into Bhadrawah. But while Garudas consume serpents in numerous myths and in artistic representations across India, and the Garuda is the vehicle of Visnu, it is also significant as the Zhang-zhung totem. Thus the myth may equally reflect a first millennium ce., conflict between the Naga worlds and the Tibetan, or more precisely Zhang-zhung, tribes to the east. Accounts of the later royal lineage also embrace Vasukinag. The Vamsavali describes how Nag-pal was born when the royal line was threatened with extinction after the death of the previous king. The widowed rani became preg- nant after spending a night in Vasukinag’s temple, doubts over the authenticity of the new heir being assuaged by his birth with a snake’s hood issuing from his back.24 Here we seem to deal with an interruption in the lineage or the establishment of a new royal line seeking to retain the status of an established lineage. Mythologically therefore, Vasukinag is central to the authority of the former ruling dynasty whether within a Sanskritic or local context. Deeper links to the Sanskritic tradition are however, difficult to identify and most importantly in the context of our interests it appears that the appellation ‘Kailas’ may be very recent indeed. Hutchinson & Vogel’s 1933 History of the Pun- jab Hill States makes no mention of ‘Kailas’ in referring to Kaplas and its annual

21 Jerath (2001: 156). Interestingly this account is attributed to a Vasuki Purana—“where pious site in the region are mentioned along with the pilgrimage of Vasuki Naaga to Shivadham [sic].” No details of the text, its language, date, etc., are given and I have seen no other reference to it; further investigation is needed. Perhaps it is not a local Bhadrawah text? This temple contradicts the claim of Daraganj to be the only Vāsuki temple in India; see Irwin (1983: 259). 22 This legend was told to me by local informants Police Superintendent Ravinder Kumar (to whom I am greatly indebted for his provision of a security escort), and Dr. Ashok Datta. I am also greatful to the pujari of Gatha temple (whose name I have unfortunately mislaid), for his assistance. Garudas often vomit up jewels in other western Himalayan myths. 23 Hutchinson & Vogel (2000: vol. ii, 616); also see Handa (2004: 104). 24 Hutchinson & Vogel (2000: vol. ii, 620–621). kaplas and manimahesh kailas 157 worship ritual. It identifies the mountain as “Kund Kaplas or Kamalas (perhaps Skrt. Kamalasaya = ‘lotus seat’).”25 But the first title is clearly applicable to the lake (kund), rather than the mountain while the second, a title not otherwise recorded, is speculative. The mountain is sacred only in the sense of it being the site of the lake and there may be no reference to ‘Kailas’ in the context of Kaplas mountain which is earlier than modern tourist literature! Even the three day annual community ritual on the mountainside is referred to locally not as a ‘Kailas yatra’ but as the ‘Vasukinag yatra’ and the central rituals of deity possession of low-caste chelas (mediums of the deity) concern not the mountain or Śiva, but the Naga deity. The ritual, organised by a local committee, follows the Hindu calendar—commencing on the 15th day after Janamashtami—and begins at the Gatha temple outside Bhadrawah, where the pujari is a Brahman. But this temple, traditionally attributed to King Nag- Pal, is the centre for local worship of the Naga deity rather than the Hindu pantheon, with its central image being of Vasukinag and his companion Jitman- ji.26 Sanskritic elements are obviously manifest in the actual ritual in only two ways. Firstly in that the two kilometre circumference lake (but not the moun- tain), is circumambulated by the worshippers and secondly in that, according to local informants, sacrifices there have been eliminated. But while images of Vasukinag and Jitman-ji are erected there during the festival,27 there is no tem- ple at the lake, only a Naga stone of uncertain antiquity and there is no local mahatmya. Modern Hindu norms may, however, also now permeate into the ritual due to its accompaniment by an Indian Police security escort.

Conclusion (Kaplas)

The addition of the toponym ‘Kailas’ to Kaplas is recent. The sacred centre is not the mountain but a Naga lake that hosts an annual community ritual in honour of the Bhadrawah deity, Vasukinag (with the presence of local Muslims on the yatra evidence of that predominantly localised identity). The Naga lake on the side of a mountain, the classic up-down journey to the ritual site rather than the Sanskritic circumambulation of the mountain (there is a feasible route, but it is not taken as part of this ritual), rites of possession by low-caste participants,

25 Hutchinson & Vogel (2000: vol. ii, 618; also see 614). 26 Hutchinson & Vogel (2000: vol. ii, 621); personal observation. 27 See Jerath (2001: between 54–55) for a photo of these images on site. 158 chapter 7 earlier sacrificial rituals, and the absence of temples at the destination are all indicative of the common Western Himalayan annual communal rituals of worship (for which there are close parallels with Tibetan mountain worship ritual patterns28). As is the case generally in Indic culture there is no evidence of mountain worship here, with the Kaplas mountain largely incidental to the site. The central deity to whom the ritual is directed is not a deified form of the mountain in any sense but a Naga, and an immigrant Naga at that! This suggests an earlier layer of belief and practice onto which Vasukinag was grafted, or that an immigrant group brought their deity with them and became the dominant culture of the region. While the lack of sources makes it difficult to uncover its origins, the pil- grimage today is clearly subject to a process of Hinduisation. Sacrifices have (reportedly) been eliminated, circumambulation of the lake introduced, the mountain has been named in association with the Kailas-abode-of-Śiva trope, and the fundamental shift from the local to the Sanskritic deity as the centre of worship is thus underway. As Śaivite Hinduism absorbs the local Naga belief system the yatra is increasingly presented in local media as a Śaivite ritual. The process is explained mythologically by an account of Śiva (rather than Kali) residing at Kaplas Kund before giving his home to the refugee Vasukinag and moving to a mountain in Chamba district. In return, the grateful Vasukinag gifted Śiva his favourite jewel (maṇi), hence Śiva became known as Maṇima- hesh (Maṇi + Mahādev [Śiva]) and his Chamba mountain as Manimahesh Kailas.29 While this mythological association of Kaplas with Śiva is weak, there is now a growing popular belief that Siva (still) resides on Kaplas mountain. None-the-less, this remains a local rather than a regional or Pan-Indian pilgrim- age. The recent addition of the name Kailas to this mountain may have been conceived by local tourism interests, although the current political situation makes that unlikely. It may be the result of religious interests, a Hindu mission- ary statement of incorporation towards a local ritual from Western Himalayan Cultural Complex beliefs. It is even possible that designating the site as ‘Kailas’ may be envisaged as a marker against expansionist Islam. But it may also be a reflection, a borrowing, from the mountain that is the topic of the next section.

28 On which see for example, Buffetrille (1998). 29 Sharma (2001: 117). K.P. Sharma is a (retired) medical officer whose duties took him throughout Chamba state; during that time he collected local lore concerning Manima- hesh Kailas. The resulting monograph is an invaluable source, conscientiously compiled and stimulatingly analytical, while not primarily intended as a critical work. kaplas and manimahesh kailas 159

The Sanskritisation process that may be seen here in outline is more clearly apparent in regard to that subject, Chamba’s Manimahesh, a mountain which, as the myth noted above suggests, is linked in a number of ways to Kaplas Kailas.

Manimahesh Kailas

Maṇimaheśa Kailas is an 18,564 feet (5,658 mtr) peak located around 40 miles (65 kilometres) east of Chamba town (see map 5). No major rivers originate there but a small lake, Gaurikund, situated on the northwestern side of the mountain at around 13,500 feet is the site of an increasingly popular annual ritual commonly termed the Manimahesh yatra. This begins at the hamlet of Harsar (Harḍar) some 10 miles past Brahmaur, the site of the famous Chaurāsi temple complex with its shrines to the 84 Mahasiddhas (to whom we shall return).30 While there are other claimants, Brahmaur was probably that centre known in classical sources as Brahmapurā31 and from around the sixth century ce., it was the capital of a regionally significant kingdom. Brahmapura as a toponym serves multiple interests through multiple meanings, a common device in the Himalayas. The Sanskrit equates to ‘town of Brahma’,but in local understanding it refers to Brahmāṇi/Bharmāṇi Devi, the patron-goddess of that early king- dom.32 Brahmapura emerges into proper historical record in the early 8th century with the reign of Meruvarman,33 who held to Sanskritic values. The earliest inscriptions indicate Śaivite/Kashmir influence,34 with the outstanding mani- festation of this being a florescence of temple building on vāstuśāstra princi- ples at Brahmaur. Craftsmen were attracted from Kangra, Kashmir, Kulu and

30 Although by no means all of the Mahasiddhas are actually represented at Chaurasi; see Sharma (2001: 72–84) for a discussion of the identities represented by the existing shrines. 31 See the discussion in Goetz (1955: 14–18). 32 Sharma (2001: 30–31); Goetz (1955: 5–28). The toponym Chamba is traced to “Byams-pa (Brahmā)” by Sharma (2009: 114), thus linking it to Brahmapura. Given the tendency for toponyms to appeal to different interest groups (as discussed in the Conclusions) we might look at other references in this toponym. An origin in the Tibetan term Cham-pa (‘religious dancers’), might be constructed, but the term is not generally known, although Cham-dpon (‘dance master’) is used; email communication, Dr. Mona Schrempf, 21 April 2009. Cham-pa (Tib: ‘cold’ or ‘catarrh’) seems unlikely. But another possible derivation is the Kunāwari cham = wool; Takeuchi (2011: 135–136). 33 On the kingdom’s origins, see Thapar (2000: 797–806); also see Sharma (2004: 395–396). 34 Sharma (2004: 395). 160 chapter 7 beyond to build an earthly recreation of a Sanskritic cosmos.35 That imagined realm centred on Mount Meru and central to this Chaurasi temple complex was what is now known as the Manimahesh temple, built on the Meru model. Its inscription does not however, refer to Manimahesh, but states that, “[t]he King himself erected a prāsāda [temple/palace], like Mount Meru on the summit of the Himavant.”36 The kingdom collapsed after the original wooden temple complex was burned during the late 8th century invasion by the Kiras. While their identity is debated, the recorded invasion of Bhadrawah around this time and the pres- ence of numerous Tibetan inscriptions in the region suggests they originated in Tibet.37 Brahmaur was certainly accessible from western Tibet, with a major trade route running via Lahaul and the Kugti pass to Brahmaur, Chamba, and the lower hills to the west. The kingdom re-emerged in the 10th century, with its centre relocated from Brahmaur to Chamba.38 The new king Sāhilavarman claimed continuity with the Brahmaur dynasty and the rebuilding of the Chaurasi temple complex, including the central—now stone—‘Manimahesh’ temple is attributed to him.39 His Varman dynasty ruled Chamba state until Indian independence around a millennium later, retaining their position over the centuries by offer- ing ritual acknowledgement of Kashmir, Mughal, Jammu, Sikh, and British overlordship as necessary. Following the loss of Bhadarwah in 1846, Chamba and the rest of its territory was included in the new Punjab Hill States. But it

35 The influx reflects both the attraction of a new centre and the acquisition of craftsmen within the process of military expansion and conquest. 36 Sharma (2001: 51). As Sharma observes, ‘Mount Meru’ signifies both a mountain and a type of temple. 37 The Rājatarangiṇī refers to Kīras around this time; Rājatarangiṇī, viii: 2767, cited in Sharma (2004: n. 49 404); also see Sharma (2009: 34–35); Goetz (1955: 26–27). The later Tibetan/Guge presence in the 11th or 12th century is indicated by the well-known inscrip- tion referring to the “august younger prince of the Garuda Lords”; see Francke (1977: vol. 1, 253–255). 38 Chamba may have already been the summer capital, or perhaps the Kira threat dictated the shift to a more secure location. Ecological restrictions on agricultural growth around Brahmaur is suggested as a factor by Jha (2000: 199), but Mahesh Sharma (email corre- spondence, May 2012), locates the shift in the context of the expansion of the Brahmaur kingdom. 39 Commonly attributed to Sāhilavarman, it is dated by the pre-eminent regional Buddhist specialist, Laxman Thakur as more probably the product of his successor’s reign in the 11th century; see Thakur (1989: 155–160). kaplas and manimahesh kailas 161 was important enough to be allotted a British Resident from 1863 onwards,40 and the dynasty survived under his ‘guidance’ until after 1947. The longevity of the Varman dynasty can be seen in addition to their obvious diplomatic skills, to have owed much to their skilful manipulation of religious symbols and belief systems, not least those surrounding Manimahesh. While a Sanskritic tradition was apparent at court level from at least the 8th century ce.,41 a range of spiritual beliefs and practices indicate that the Chamba valley was part of the Western Himalayan Cultural Complex, with its diverse range of understandings of the sacred. In addition to worship of Nagas and creatures such as cows, fish and monkeys, as well as trees and plants (e.g.; banyan and tulsi), these included belief in a range of denizens of the spirit world. That world included local guardian spirits whose domains extended over the lakes, rivers, and other natural features they were believed to inhabit. Human settlements were overseen by localised village deities (deo- tas) associated with a defined territory (generally explicable through their origin myths), over which their power extended. Again characteristic of the Western Himalayan region, this diverse assembly of sacred beings and enti- ties were not systemised and any hierarchies were rudimentary. Particular spirits were acknowledged or worshipped when relevant, at natural shrines rather than temples, in some cases by individuals, in others by communities or groups.42 It is in this context that we locate the Manimahesh yatra, which is closely related to the Kaplas yatra. Thus the annual ritual journey does not tradition- ally centre on worship of the mountain in some form or even of Śiva (although that aspect is rapidly developing in popular understanding). The central rite is again the propitiation not of a local mountain or Sanskritic deity, but a Naga deity, in this case Kelangnaga (or Kailunganaga). This is explicit in offi- cial instructions for the ritual cycle given as recently as 1995, which make no mention of Śiva.43

40 Most of the Hill States did not have a separate British Resident. 41 The earliest recorded inscriptions in Brahmaur date to the 7th century ce.; Sharma (2001: 5). 42 See for example, Sharma (2001: 10–14, 26, 51–53). 43 Sharma cites a Document (iii. 22.5) concerning “Expenditure Incurred by the Lakṣmī- Nārāyaṇa Temple on the First Day of Maṇimaheśa Jātrā in Chambā in 1995; On the auspi- cious occasion, sharp at 4:00am the banner (chaḍī) will be immersed in water and the holy dip will begin with the shamans (celās) of Kailung Nāga jumping into the water. Thereafter shall all the devotees take bath.” Sharma (2009: 122). On the pilgrimage process see Sharma (2009: 122–123). 162 chapter 7

As at Kaplas there are no temples on Manimahesh, only a single image at Gaurikund—a small white marble chaumukhalingam (‘Four-faced [Śiva] lingam’), of uncertain antiquity.44 And as with the Kaplash yatra a major part in the ritual is played by spirit-mediums, the chelas through whose trances the deity manifests and responds to pilgrim’s questions. But who is that deity? Modern literature identifies it as Śiva. In earlier sources however, as well as in official edicts, the mediums are chelas of Kelangnaga just as the chelas at Bhadrawah are of Vasukinag. Kelangnaga is actually the protective deity not of Brahmaur or Chamba, but of Kugti village, a settlement some hours walk east of Harsar (see map 5). Kugti is a frontier town, strategically and ritually significant as the last village in the Chamba valley before the pass to Lahaul and a temple to Kelangnaga marks the farthest extent of the village boundary on that route. Below the temple is a fractured rock where the Naga is said to reside and which sits beside the trail through the empty frontier with Lahaul to Triloknath, control of which was historically disputed between Kulu, Ladakh and Chamba. Thus Kelangnaga can be seen as guarding the frontier with Lahul,45 and his worship at Manimahesh situated in that protective context. Yet a simple model of ancient ritual cannot be easily supported here, for the antiquity of worship of Kelangnaga at Manimahesh does not necessarily predate a Śaivite presence in the region! As with Vasukinag and many other territorial deities, Kelangnaga is an immigrant. The Chamba Gazetteer of 1904 dates his arrival from Lahaul to some 15 or 16 generations earlier.46 That would suggest (allowing for the age of the source), the 16th century.47 But at the hamlet

44 The statue is commonly ascribed to the 7th century ce; Ohri (1991: 78). Ohri (1991: 19–24), on the grounds of the statue style, makes an interesting case for the reference to Kailas in Banbhatta’s Harshacharita (7th century ce.) being to Manimahesh Kailas. A later date is probable, however, for there was something of a regional fashion for white marble statues in the early second millennium. See Vitali (1996: n. 491 317–318); also see, on diverse regional white marble statues, Buffetrille (2000: 57); Cousens (2008) and (2010: 54–56); Malhotra (1983: 49); Sharma (2004: 420–421, 427), Sharma (2009: 59); Thakur (2001: 40). A comprehensive study is called for. 45 “Until you have seen the Manimahesh Kailas from the 5040 metre top of Kugti Pass you cannot appreciate what an incredible mountain it really is.” Dr. Peter Raine (University of Waikato), email, 6 January 2005. 46 The Gazetteer records that the Naga came, “because disease was prevalent among the cattle of the state”,and that Vasukinag had come to Chamba from Bhadrawah for the same reason “100 years ago”; Rose & Hutchinson (1996: 188). For Lahaul and Chamba beliefs concerning Kelangnaga, see Handa (2004: 208–213). 47 Here I follow the conclusion that a generation equates to 25 years; see Spitzer (1973: 353–385). kaplas and manimahesh kailas 163 of Harsar where the ritual ascent of the mountain begins and which is the site of the local cremation ground, there is a temple with inscriptions that enable us to date it as Śaivite by 1582.48 Perhaps the mythology which associates Kelangnaga’s migration with curing disease among livestock relates to the fact that Chamba, enriched by the dis- covery of copper mines, conquered western Lahaul during the 1558–1582 reign of Pratap Singh.49 Chamba’s expansion across the Kugti pass seems to have generated a mythological corpus to reinforce the new realities, which appar- ently included the movement of a population group from Lahaul to Kugti. In a device that may be associated with the need to unite the frontier populations,50 Kelangnaga, deity of a conquered peoples or refugee group, was cosmologi- cally elevated to the role of ‘Prime Minister’ (wazīr) of Manimahesh,51 and a shrine was dedicated to him at the Chaurasi temple complex. Kelangnaga, now protector of the frontier, thus took a central role in the ritual complex around Manimahesh. As we have seen, the Sanskritised Kaplas Kailas mythology indicates that the term ‘Manimahesh’ derives from the combination of mani (‘Jewel’) and Mahesh (‘Mahadev’ = Śiva), in what seems “a forced etymology, a conscious design to propel the local to universal.”52 Yet while Maheshwar is one of the names by which Śiva is known, and of a sect of his followers, it is not exclusive to Śiva; such titles can serve different interest groups. There is, for example, a clan of local deities in Kinnaur whose identity has been Sanskritised under the name Maheshwar,53 and it seems plausible that those deotas were referred to in the earliest use of the term, with a possible medium being refugees from Kinnaur who settled at Brahmaur in the second quarter of the 8th century.54 It seems significant that—outside of Śaivite associations—there is no myth- ological structure around or iconographical representation of a divine entity

48 Sharma (2001: 86–88). The central image at Harsar temple today is of Śivshankar. 49 Goetz (1955: 108); Rose & Hutchinson (1996: 84–86). 50 The elevated stature given to Kelangnaga and the myth of disease-curing might suggest a population group from Lahaul who preferred residence in the Chamba kingdom rather than a conquered group taken as slaves. Settling an immigrant group on the frontier (e.g.: Kugti) would be consistent with Indic, indeed Asian, tradition. 51 Sharma (2009: 115). 52 Mahesh Sharma, email communication, March 2012; also see, Sharma (2009: 106–121). 53 Sanan and Swadi (1998: 111, 118). I am suggesting here that the clan name precedes Mahesh- war as one of the names of Śiva, and that its adoption by the Śaivites was a part of the process of their absorption of local deities. 54 Goetz (1955: 22–26), whose sources include both the Rajatarangini and numismatic evi- dence. 164 chapter 7 called Manimahesh, which would be expected of a territorial deity.55 Nor does the Manimahesh temple at Brahmaur help us here, there is nothing to indicate any relationship to the mountain beyond the name,56 there are no early refer- ences to it by that name, and we do not know when it acquired the name it shares with the mountain. Even the central image in the temple is not a local deity but a lingam indicating a Śaivite focus. Here we turn to the work of Mahesh Sharma, who has published a num- ber of important critical studies of Chamba history from epigraphic sources, focussing particularly on the ruling elites’ religious policy strategies. Sharma locates the earliest reference to Manimahesh in a grant dated to though per- haps later than, 1440. The grant begins, “homage to Śrī Maṇimaheśa”,57 a Man- imahesh who does not appear to be a Sanskritic entity.58 Sharma has found no evidence that a Manimahesh was worshipped in Chamba at that time and even today a deity Manimahesh is not known to be worshipped outside a sin- gle small shrine at the Charpaṭnāth temple in Chamba, in the portals of a courtyard restricted to the renunciate followers of Charpaṭ (to whom we shall return). A non-Sanskritic Manimahesh (or simply ‘Mahesh/war’), was most probably a local territorial deity in the Brahmaur region. Indeed (in accordance with Sharma’s wider findings), the 1440 inscription might suggest the first (ritually and status-wise highly significant), royal acknowledgment of a popular cult too powerful and influential to ignore, and which was thus soon Sanskritised. Certainly the status of Manimahesh had increased and been significantly transformed through Śaivite association when it next appears in the sources. In 1618, there is reference to “Mahā-Rudra-Maṇimaheśa”, and land grants to four merchants to establish an alms house for pilgrims (sadāvrata) at “Datta Hilsar” (Harsar?) in honour of that deity. The grant also refers to the now Śaivite-associated bathing ritual at the lake indicating, Sharma concludes, that “[t]hese grants formulate the extent of acculturation, whereby local symbols of

55 This point requires further elaboration through the study of local (Pahari language) oral history. I was advised by Chamba pandit Harish Chander Sharma (Interview, Chamba, 20 June 2003), that there is no mahatmya for Manimahesh. 56 The central tower (shikhara = ‘mountain peak’) above the main shrine of major Hindu temples commonly takes the name of a great mountain such as Kailas, Meru etc., depend- ing on style; Eck: (2012: 77). Here therefore it has simply taken the name of the most prominent local mountain. 57 Sharma (2009: 44–45, 107). 58 Sharma notes that in contrast to grants made to Pan-Indian Brahmans, grants to popular local deities such as Manimahesh were in Pahari; Sharma (2004: 410). kaplas and manimahesh kailas 165 faith were universalised, Sanskritised and reordered in a definitive Sanskritic sectarian hierarchy.”59 In a faint echo of the mythology of Kailas-Manasarovar, and perhaps an attempt to link the sacred site to the wider Sanskritic cosmos, the Vedic deity Kubera also appears on the sacred landscape around this time. Yet the under- standing of Kubera here might also be linked to the Nagas. Kubera was a loka- pala, the guardian of the northern direction, while the Nagas with which he was associated guarded underground treasure.60 A shrine dedicated to Kubera (with central lingam61), was created to the left of the Manimahesh temple at Chaurasi. This was mythologically explained through the Epic-influenced account of Kubera fleeing to Manimahesh after being expelled from Lanka by Ravana. At Manimahesh, his friend Śiva offered him a place to settle at Dhanchho, the usual overnight halting place on the mountain section of the yatra.62 Other Sanskritised myths emerged. Śiva is said to have first revealed his true identity to a Gaddi shepherd he employed to carry salt after himself manifest- ing as a shepherd near Gaurikund lake.63 The prominent Puranic account of grief-stricken Śiva carrying the disintegrating body of Umā (with the falling pieces giving rise to the Sakta-pitha sites), was adapted to end with Śiva’s return to Dhanchho where he became lost in meditation until the gods persuaded him to return to his usual role.64 The cosmological world of Manimahesh was thus increasingly peopled by Śaivite-centred deities and filled with Śaivite sacred sites, and even the presence of a Naga was typical of the Śaivite frontier world in which local deities of the Western Himalayan Cultural Complex were reimag- ined as aspects or manifestations of Śiva. Through Sanskritisation, Manima- hesh was becoming Śivabhumi—Śiva’s (sacred) land. (There exist, however, different ideas of Śiva’s presence, whether it be seasonal, with Śiva resident there in summer, or as another account indicates, resident there for 364 days of the year, with the other day spent at Kailas-Manasarovar.65) The question of agency may be located in the religious structures of Chamba state. That Manimahesh was initially Sanskritised through the medium of Rudra suggests the influence in the region of Śaivite renunciates who followed

59 Sharma (2004: 410–411). 60 I am indebted to Mahesh Sharma for this point. 61 The lingam of course, is usually representative of Śiva. 62 Sharma (2001: 38, 88). 63 Sharma (2001: 88–89). 64 Sharma (2001: 91–93). 65 Bhawa (1998: 37). 166 chapter 7

Rudra-Śiva. Supporting this is the legend of the Brahmaur protective deity Brahmani Devi allowing the 84 Mahasiddhas to stay overnight on her terri- tory when they came on pilgrimage to Manimahesh. In the morning she was angered to find Śiva lingams scattered through her realm and to appease her Śiva ordained that pilgrims would henceforth bath at her pool on the mountain (Gaurikund). As Sharma observes, “[w]hile this myth creates a sacred space, it also acculturates and accommodates the existing symbol, as well as lay basis for the pilgrimage.”66 The 84 Mahasiddhas play an important role in the foundational accounts of Chamba state. A number of traditions link them with Brahmaur, includ- ing the samadhi shrines and lingams of Chaurasi temple that are attributed to them. Tradition also associates Sahillavarman’s 10th century relocation of the capital to Chamba with the 84 Mahasiddhas, in particular with Charpaṭnāth, a Kānphaṭa/Goraknāth/Nāth67 renuciate who became the king’s rajaguru.68 But their mythological prominence, like the Pandavas or Padmasambhava, makes them notoriously problematic for Western historical analysis and Charpatnath is no exception. Reputed to have been a Rasa (alchemist), who came to the region in search of herbs or minerals, he became the subject of a cult sub- sequently eclipsed by Laxmi-Narayan, at whose temple in Chamba he was honoured with a shrine.69 But although Charpatnath remains closely associ- ated with the 10th century formation of Chamba state, there are no known references to him before the 12th century, and the earliest reference to him in Chamba was in the 16th century Chamba Vamsavali.70 Perhaps his real signifi- cance is as a symbol of the regional presence of renunciate movements such as the Naths. That Śiva is said to have first revealed himself to a Gaddi shepherd is signif- icant. The Gaddis, native to Brahmaur,71 are a migratory tribe who winter in the Punjab and graze their sheep and goats in Himachal during the summers. They provide an explicit link between the Kaplas and Manimahesh pilgrim- ages, bringing sheep from Bhadrawah for sacrifice during the Manimahesh fes- tival and they may be the earliest group to have determined the site as sacred.72

66 Sharma (2001: 158). 67 The terms are interchangeable; on this movement, see Briggs (2001). 68 Goetz (1955: 29–35). On the Mahasiddhas, see Dowman (1985); also see White (1996: 78–93). 69 Sharma (1996: 75); also see Sharma (2001: 16–17). 70 Sharma (1996); also see Sharma (2009, esp., 75–136). 71 Goetz (1955: 25) dates the arrival of the Gaddi in Brahmaur “not before 1000”. 72 Sharma (2001: 117–121). Sheep are also brought from Triloknath after the annual Pori kaplas and manimahesh kailas 167

Their ritual journey begins at Bheja’s Vasukinag temple at the foot of the Padari pass,73 and after Manimahesh they return to take part in the Kaplas yatra. One narrative to explain this is that their journey is in homage to Śiva for allowing refuge to Vasukinag.74 Thus while Bhadrawah passed to the political control of Jammu more than 150 years ago, the Gaddis continue to share a cultural world and symbols of local identity that transcend the modern political boundary. The Gaddis have attracted some ethnographic and anthropological atten- tion, most prominently Molly Kushal’s work, which firmly locates the Gaddis within a Śaivite world.75 She recounts their belief that Śiva created the Gaddis from a speck of dirt on his body and (in an echo of Tantric sartorial emula- tion of the deity), that their characteristic dress—cap, woollen gown and black woollen rope around the waist—is precisely that worn by Śiva. Importantly, they define the land in which they graze their flocks as Śivabhumi.76 Their cosmology, Kushal states, envisages Chaurasi temple complex as a key landmark—“the original celestial kingdom”, but;

[T]he most outstanding and symbolically relevant landmark of the land- scape is Mount Mani Mahesh. Believed to be the abode of Lord Shiva, its

festival there in August. That also involves a ritual carried out at Saptadhara, the site of seven springs on the side of a mountain outside Triloknath, on which see Cousens (2010: 60–66). Despite this link between Kaplas, Manimahesh, and Triloknath, the latter does not seem to be identified as a Kailas-abode-of-Śiva site, again suggesting the predominantly local origins of these sites. 73 Hutchinson & Vogel (2000: vol. 2, 617). 74 Sharma (2001: 120); also see Rose & Hutchinson (1996: 188). The belief that Śiva gave Kaplas to Vasukinag is prominent in Bhadrawah rather than Chamba. 75 Kushal (2000). Yet the Gaddi world even today is not monotheistically Śaivite. In addition to a cult around the Śaivite siddha-related popular deity Bābā Koṭ (“an entirely mythic synthesis of real siddhas centred on Mandi kingdom”: Davidson 2002: 191–199) they main- tain beliefs in deities such as Sankri Devta, Sankhpal Devta, Sopor Devta and Sendu Bir (who has a small body and long beard, whistles as he moves around, and bestows power over animals). The Gaddis thus maintain, and assist in the integration, of differing layers of belief; see Jerath (2001: 165); also see, for a 1904 list of Gaddi religious practices, Rose & Hutchinson (1996: 150–151); also see Shashi (1979: 92–98). For a socio-economic study of the Gaddi, see Sharma (2012). 76 See oioc, Emerson, Treatise, chapter xvii, 6: “The Brahmur [sic] pargana of the Chamba State is known as Shiva bhumi, or the land of Shiva, for the god is believed to dwell on Kailas which overlooks it. This is not the Kailas of Bushahr which again is but a block, removed through the austerities of a pious ascetic, of the true Kailas … situated … in Western Tibet.” 168 chapter 7

presence makes the entire region sacred and divine in origin and essence. It transforms the physicality of the landscape into a cultural text, which is constitutive of [a] large discourse through which people contemplate and speak about themselves, the society and the divine.77

The Gaddi pilgrimage to Manimahesh is thus, Kushal states, a “cultural text” that links the physical landscape with its imagined reflection at the macrocos- mic level. But it is also an indication that in both local and Pan-Indic bodies of knowledge it is the presence of Śiva that sacralises a mountain. They were and are not sacred to a local deity or seen as sanctified other than as Śiva’s abode.

The Chamba Centre

While the Chamba royal court had been Sanskritised much earlier, it was only in the 16th century that a Śaivite sacred geography was imposed onto the exist- ing landscape sacralities. In their earlier form both Kaplas and Manimahesh were sacred as the location of Naga lakes, not as sacred mountains. Indeed the actual sacrality of either mountain exists only in the modern Śaivite mytholog- ical overlay, just as the sacrality of Lake Manasarovar predates that of Mount Kailas. But in contrast to the Naga lake deity of Manasarovar, Vasukinag and Kelangnaga are both immigrants to the lakes they inhabit. Thus implying an ancient relationship between these three lakes, or situating them all within the sacred geography of the Western Himalayan Cultural Complex requires an understanding of that culture as conceptually fluid and dynamic in its under- standing of the shifting realms of deities. But there is considerable evidence that deities, including—in the Tibetan realm—sacred mountains, moved with the peoples who worshipped them and thus specific Naga lake rulers need not be conceived of as ‘timeless’ or ‘ancient’, but rather situated within historical processes of discovery, migration, conquest, identity and suchlike. It is certainly consistent with the practices of renunciates that they should be attracted to such sites of local sacrality, however temporary or shifting those sacralities may have been. In their conception these were places, indeed sources, of primeval sacred power. It was these primeval powers that renun- ciates sought to access, and World-religions sought to overcome, and the ac- counts of the conversion of Nagas to Hinduism or to Buddhism, or the Bön claim to original authority over them, represent the suppression of the belief

77 Kushal (2000–2001: 34). kaplas and manimahesh kailas 169 system(s), and thus probably the political system(s), of the peoples of the West- ern Himalayan Cultural Complex. The initial agents of that take-over—in an environment where armies could not prosper—were the renunciates. For the World-religions this was a field of contestation for the assimilation of local beliefs, and even—as in the identity of the 84 Mahasiddhas—a field where those identities might blend together. The history of these regions is inexorably linked to the presence of renunci- ates, both Indic and Tibetan. The Chamba valley was for them an alchemical centre en route to Jvalamukhi, the place of the eternal flame. As noted, Jvala- mukhi was a source of sulphur, important in the context of Tantric alchemy, and it was a major gathering place for renunciate followers of esoteric traditions, a site where the distinctions between Hinduism and Buddhism were clearly blurred. We know that Tibetan ascetics visited the Chamba valley from at least the early 13th century (see Chapter 12) and there were conceptual models and ritual technologies that these renunciates understood and deployed as appro- priate, not least that of a mountain-dwelling deity at the centre of a sacred realm. Thus Mahesh Sharma identifies the making of the Manimahesh sym- bol as influenced by these Tibetan visitors,78 and as will be seen in the next chapter, a similar influence is probable in regard to the Kinnaur Kailas. The Tibetans do seem to have established an on-going link to Manimahesh, albeit a minor one. (A link that has been re-established post-1959, with Tibetans from Dharamsala joining the yatra today.) The earliest mention of the moun- tain in European sources is the account of Alexander Gerard, who travelled in Kinnaur in 1821. Discussing the Buddhists of Bushahr he refers to “Mumma- hez in Chumba” [sic] as among several sacred places “frequented by the Lamas …”,79 although Jvalamukhi and Rewalsar seem to have been more esteemed by them.80 With its temple complex at Brahmaur on the route from Lahaul to Chamba and the Chamba valley’s easy accessibility to Jvalamukhi, Chamba state had a lasting attraction to renunciates of all types. Accounts of the presence of great renouncers accentuated the sanctity of the valley and could incorporate a variety of traditions within a wider sacred narrative. In the 16th century for example, Sri Chand, a renowned Hindu ascetic who was the eldest son of the Sikh Guru Nanak (1469–1539), is believed to have come to Manimahesh

78 Sharma (2009: 120–121). 79 Gerard (1996: 211–213); Gerard was aware of Kailas-Manasarovar, which he refers to here. 80 Rewalsar (Tib: mTsho Padma) is the site of a lake regarded by Tibetan Buddhists as the residence of a Klu, transformed by the powers of Padmasambhava into a Buddhist sacred site; on which see Emerson, Treatise, Chapter 7; 14; also see Cantwell (1995: 3–9). 170 chapter 7 on pilgrimage. He reportedly demonstrated miraculous powers in Chamba, where there is a temple dedicated to him, before ‘disappearing’—i.e., attaining samadhī—at Manimahesh.81 Chamba certainly provided a favourable haven for Śaivite renunciates, lo- cally known by the generic term ‘Jogis’, and the nature and extent of their influ- ence on medieval Chamba is testified to in numerous inscriptions. The Jogis’ dwelling places were usually in the peripheral areas of the state, at localised shrines, caves, springs, cremation grounds and so on. Such residences implied transaction with the local spirit forces. Cults formed around the more charis- matic Jogis who demonstrated their mastery over the local spirit world and the collective body of renunciates acted as a force for Sanskritisation through the reinterpretation of local meanings in a Sanskritised context; the local under- standing of monkeys as sacred, for example, was easily reoriented to the San- skritic deity Hanuman.82 Other elements of the process included the transfor- mation of popular myths and the introduction of Sanskritised deities, rituals, and festival calendar. But this renunciate activity did not equate to an immediate transition to the Brahmanical ritual and purity norms. The Jogis commonly used alcohol or cannabis, maintained consorts, and otherwise violated those norms. Their relationship with temple Brahmans was therefore often problematic. But the renunciate community required patronage to prosper. That patronage might initially be found from the local community attracted by the charismatic pow- ers of an individual, but the local power developed by successful Jogis could, even while serving its broader interests in terms of the Sanskritisation process, be seen as threatening to the ideological control of the established elites. That class—following the model articulated in the Arthasastra—therefore used patronage to assert its authority over the renunciate community. Patronage was already used as a means to establish and control institutions such as urban tem- ples with Brahman pujaris. Extending that patronage to renunciate shrines was a logical development of that strategy.83

81 Interview with pandit Harish Chander Sharma, Chamba, 20 June 2003; the temple is located near the museum. I have located no mention of this story in the basic sources regarding the Sikh gurus. See however, Clarke (2006: 44), who records that the tutelary deity of the Udāsin and Nirmal akhāṛās is Candra Bhagvān = Śrī Chand, the eldest of two sons of Guru Nanak; also see Clarke (2006: n. 9 55–56). Guru Nanak is also said to have visited Kailas-Manasarovar; see Berar (1999: 132–133). 82 Sharma (2001: 10–14). 83 See for an example of a similar process in Nepal, Bouillier (1991: 151–170). kaplas and manimahesh kailas 171

In a process that will be seen to be common throughout the Western Himala- yas, Chamba royalty’s patronage of the renunciate community was just one aspect of their wider Sanskritisation policies. They also adopted Sanskritic titles, patronised Sanskritic temples and learning, issued land grants in which the preliminary royal eulogies defined kingship on the Sanskritic model, and undertook pilgrimages to Sanskritic shrines such as Kurukeshetra and Prayag to legitimise their status on the Pan-Indic model. At such sites they acquired social capital through generous offerings the scale of which would be known at Pan- Indic level, and they would invite religious specialists (both temple Brahmans and renunciates), to visit Chamba. Those who took up the invitation would be given land grants and became part of the legitimising apparatus of the state and its Sanskritisation programme (which might include retrospective genealogi- cal attributions of high Sanskritic status, to the royal family, i.e., descent from deities or semi-divine figures of legend).84 Sanskritic norms thus became institutionalised in state structures, and pro- vided an elite model for emulation by other social groups whose interests might be best served by association of their cultural manifestations with those norms. Such re-interpretations as a local monkey deity being a manifestation of Hanu- man brought both validity and potential patronage to their realm. From the royal perspective, of course, that patronage of Sanskritic culture was the duty of kings. As in Kashmir and much of the western Himalayas, the royal families ini- tially tried to use Laxmi-Narayan rather than a Śaivite emanation as their cen- tral deity and principal integrating symbol. In Chamba the new cult was cen- tred around the pre-eminent Laxmi-Narayan temple, patronised through land grants and serving as a symbol of royal power.85 Yetwhile stylistically and by the Vamsavali dated to the 10th century, the temple may be many centuries later.86 The chronicle also indicates that the protective deity of Chamba was actually Bhadrakali, worshipped “particularly to gain victory in war.” That association with (impure) violence meant that Jogis rather than temple Brahmans (with their purity concerns), acted as the ritualists of Bhadrakali at centres such as the Laxmi-Narayan temple.87

84 Sharma (2004: 387–432). 85 Bhawa (1998: 51, 160). On Laxmi-Narayan as an integrative agency in the context of diverse Śaivite and Vaisnavite tendencies in the region, see Sharma (2009: 135–137). 86 The first epigraphic reference to the temple is not until 1481 and the first grant recorded dates to 1582; Sharma (2009: 58–59). 87 Sharma (2009: 180–181). 172 chapter 7

Under this balancing and integrating (or dividing and ruling!) policy, royal patronage was extended to a variety of renunciate orders, who were not neces- sarily complementary. By at least the late-17th century, Jogis of the Charpatnath tradition began to be challenged by the Dasanamis, with the Girī order charac- teristically prominent in expansion into new Himalayan territory. A land grant to “Balahhadra Giri and Bihari Giri, disciples of Ramesvara”, is recorded in 1688, and the Giris were soon granted a symbolically significant role in the Manima- hesh yatra.88 Manimahesh’s 16th century rise to prominence seems locatable in this con- text; that of the Chamba ruling dynasty using Sanskritic symbols to manufac- ture “consent to rule” and to link them into the Indic religious world. This was achieved by such means as the recasting of the central deity as Śiva, and by the construction of a cosmology appropriated from the Sanskritic. This saw local features such as western Himalayan rivers being linked to the major rivers of India, and Manimahesh to Kailas-Manasarovar, “thereby creating a parallel sacred territory by elevating the local to the universal.”89 Such correspondences were not unique to Chamba. A similar process is described by Eva Neumaier in Zanskar, with the creation of a Kailas centre moving Zanskar from the periph- ery in a Lhasa-centered perspective to a more central position. She makes the important distinction—clearly also applicable in the Chamba Manimahesh process—that in contrast to;

the documented galactic polity of Buddhist states in South East Asia, where the center inscribes its own organizational paradigm on to the epi-centres, in the Tibetan situation the periphery or frontier writes itself into the symbolic center, thereby displacing the center from its position of power.90

Neumaier describes this process as it occurs in traditional texts. While there is no mahatmya for Manimahesh Kailas, we may locate the process textually as occurring today in tourist literature, newspapers, and popular books and magazines. These outlets, in various languages, retell the mythology, promote the benefits of and encourage participation in the Manimahesh yatra, and

88 Sharma (2004: 415–416). For a description of the ritual, see Yadav (1988: 5) [This work is obtainable in Chamba]. 89 Sharma (2004: 407–418; the quotation is from 407). 90 Neumaier (2002: 327–341; quotation from 329); also see, for a study of this process in Nepal, Michaels (1990); and in the construction of an “eternal Ayodhyā … at the pinnacle of a cosmic and soteriological hierachy”, Paramsivan (2009: 101–115, quotation from 110). kaplas and manimahesh kailas 173 reinforce the perception of it as a Pan-Indian pilgrimage. These modern texts, however, describe a Śaivite ritual cosmos that is far from fully integrated with the reality of the Manimahesh yatra, which is that the ritual performance remains centred on Kelangnaga. The layering of sacrality allows an imaginative space for competing narratives, with ritual participants understanding the yatra within their own conceptual world. The mountains thus serve a variety of interests, local, regional, and Pan-Indian. The transformation of the land of local territorially-based sacralities and ritual worship of lake-dwelling Nagas into Śiva bhumi, the realm around the central Kailas mountain abode of Śiva, is thus on-going, with a similar process in its infancy at Kaplas. In both cases a local lake-centred sacred geography is reimagined. The ideologically dominant sacred centre becomes the Kailas mountain abode of Śiva, with the sacred geography, in a move from the local to the Pan-Indian, becoming Śaivite. Kaplas and Manimahesh are not ‘sacred mountains’ in themselves; they have no local sacred associations and lack the rivers essential to the ideal model of a cosmological mountain. They are sacred only as the residence of Śiva, whose presence displaces the local centrality of the Naga deities. chapter 8 The Kinnaur and the Adhi Kailas

Kinnaur (‘Kinner’) Kailas

The 19,789ft (6032 metre) Kinnaur1 Kailas peak is situated northeast of Simla on the Sutlej river route east. It is most clearly viewed from Kalpa (formerly Chini2), an ancient town above the modern administrative centre of Rekong Peo. In the colonial period this region was part of Bushahr, the largest of the Punjab Hill States and the peak’s proximity to Simla, the former British imperial summer capital, made it the best known of the Indian Kailas mountains.3 But whereas the others are distinct peaks of striking appearance, this Kailas differs from its namesakes because its identity is not immediately obvious. It is one of a number of peaks on a massif or range of mountains (sometimes collectively referred to as Kailas), and many earlier travellers and even recent guidebooks have confused it with the higher Jorkanden peak to the southeast (6473mtrs), or a third peak to the southwest named Raldang (5499mtrs).4 The peaks may not have been specifically distinguished in pre-modern thought. The surveyor J.D. Herbert, who travelled through this region with Lt. Patrick Gerard in 1819, observed that Indians called, “… every high place by the term Kailas”, and the map accompanying his published account of the journey showed a Raldang, but not a Kailas, peak.5 Herbert also quoted a local Buddhist monk who;

admitted that the snowy peaks were objects of great reverence; in fact he seemed to believe in a genius of the Himalaya whom he considered as

1 Singh (1989: n. 5 1) gives 14 examples of different spellings of Kinnaur, which is now the official English transcription of the vernacular. While I reluctantly retain this spelling as most familiar to foreign readers, I am advised by Dr. Rafal Beszterda (University of Warsaw; email communication, May 2012) that the spelling Kinner is invariably used locally in regard to the mountain and that this form has a distinct identifying function. 2 On which see Tucci (1971a). 3 Writing for a Christian weekly in 1917, the Christian renunciate, Sadhu Sundar Singh, thought it necessary to clarify that the Mount Kailas he sought to travel to was in Tibet, rather than “near Chini where there is a hospital maintained by the Salvation Army”; Singh, quoted in Thompson (2002: 110). 4 Kapadia (1988–1989: 32); also see Kumar (1981: 18). 5 Herbert (1825: quotation from 351).

© koninklijke brill nv, leiden, 2015 | doi: 10.1163/9789004306189_010 kinnaur and adhi kailas 175

entitled to worship. He called these peaks Kailas that rise immediately from the village, and which constitute the Raldang cluster.6

Lt. Gerard’s brother, Captain Alexander Gerard, refers to “the alpine group of Ruldung [sic], where there is one peak of 21,000 feet surrounded by many others of inferior elevation” and also to “the lofty Kylas or Ruldung peaks”.7 Captain Madden, who travelled there in 1845, also refers to the “Ruldang group”, adding the otherwise unknown claim that “Ruldung” was Kinnauri for “Śiva”.8 Some 85 years later the forest officer Maurice Glover, being with his wife the first Europeans to circumambulate the mountain, had a different idea. He stated that;

The topmost peak, Raldang, 21,250ft, shows only its head above the lesser peaks which hide it from the bungalow [Kalpa government rest-house]; to the right lies “Castle Rock,” while the name “Kailas” is given to a pillar of rock which lies away to the left.9

Glover’s “pillar of rock” may be significant here. It seems to be that now known as ‘Śivling’, or by the local name cited by M.R. Thakur. He records the belief that while Śiva brought the sacred Kailas to Raldang the deity actually lives on that “pillar of rock” known as “Khaskar.”10 Elsewhere Thakur explains that;

For the old people [of Kinnaur], the Kailāśa was not the abode of Śiva. It was believed to be the home of the souls of the dead.11 Instead it is Khakhsar [sic], another moni [sic: maṇi?] in front of Kalpa and Rekong Peo … which has a sacred character. It changes its colour thrice a days, white at sun-rise, red at mid-day and green at sun-set—and people wor- shipped it from time immemorial.12

Clearly there are different bodies of knowledge competing here, but for our purposes we may accept that historically the present Kailas massif was more commonly known as Raldang, that it was associated with the dead ancestors of

6 Herbert (1825: 361–362). 7 Gerard (1993/1841: 10); also see Gerard (1996/1840: 93). 8 Madden (1846: 112). 9 Glover (1930: 81). 10 Thakur (1997: 181). 11 That it was the home of the dead was also noted by Glover (1930: 82). 12 Thakur (2003: 321). 176 chapter 8 the community, and that the term ‘Kailas’ was, at least in part a reference to the snow-capped peaks in general rather than specifically to one peak. The Khaskar outcrop was also sacred space in local understanding, not least due to the natural colour phenomena it manifests. In the modern period, the three major peaks of the massif acquired individual toponyms, which only became widely established when the mountaineering and scientific communities recorded them. Not only does Kinnaur Kailas lack the distinct identity of the other Kailas mountains, but it is also atypical (and highly unusual for a Hindu ritual site), in that there is no lake or river system associated with the sacred centre. In addition, although the predominant character of the modern pilgrimage is promoted as Śaivite and it follows the Hindu ritual calendar, it is unique among Indian Kailas’s for its explicit Buddhist component. It is also the only Indian Kailas in which the annual ritual is centred on a parikrama. The tradi- tional route—over 200 kilometres in total—proceeds via a Buddhist temple at Powari,13 on via the Tedong/Tirung valley over the Charang pass and down into the Baspa valley on the return to Powari.14 That Buddhist character is a clue to its origins.

The Political Background

As with most of the western Himalayas, Kinnaur’s history is difficult to isolate from legend, particularly prior to the second millennium ce. Early references in classical Indic sources to a Kinner/Kinnara tribe cannot for example, be equated to a precise geographical location or identified with any certainty as related to a specific ethnic group, although a connection to modern Kinnaur is generally assumed. But much of Kinnaur was probably tributary to the primarily Śaivite Katyuri dynasty prior to Katyuri decline in the 11th century.15

13 There are differing accounts of the starting point for the pilgrimage; Thakur (2001: 7) provides the fullest description, stating that the “pilgrimage begins from Kupa. The pil- grims after visiting Harang peak would pass through Mebur, Barang, Tangling, Powari, Purbani, Ribba, Rispa, Morang, Thangi, Lambara, Tsarang, Lalanti (uninhabited), Chitkul, Takcham, and finally arrive at their journey’s end at Sangla.” 14 Today however, the route is largely travelled by motorable roads. Instead of a journey of around two weeks it comprises a three day walk from Thangi (past Morang in the Tedong valley) to Chitkul in the Baspa valley. Either the Charang (5266m) or the Mangsu (5367m) passes are crossed to reach Chitkul. See, for a rather eclectic account of the pilgrimage, Jerath (2001: 135–137). 15 Joshi (1988). kinnaur and adhi kailas 177

Administrator and chronicler Jogishwar Singh, noting the absence of any mention of Bushahr in key regional sources such as the account by 7th cen- tury Chinese pilgrim Hiuen Tsang, the 12th century Kashmiri chronicle the Rājātārangini, or accounts by either the Jesuit or Mughal scribes, argues that the region was simply not important enough to attract outside attention! Cer- tainly the processes of state formation, including the development of a written script for the Kinnauri language and the emergence of symbols of state identity, were absent or apparent only from the late medieval period.16 The Bushahr royal chronicles claim an unbroken dynastic rule of over 120 generations back to a son or grandson of the god Krishna.17 But inscriptions, language, culture, and the chain of temples attributed to Tibetan Buddhist pioneer Rinchen Zangpo (see Chapter 11) indicate that the rule of western Tibetan dynasties extended across the modern border to the vicinity of the Kinnaur Kailas until the collapse of the Guge kingdom in the mid-late 17th century.18 That upper Kinnaur territory then briefly passed to the control of the Ladakhi kingdom which conquered Guge in 1630. By the late 17th cen- tury new political movements to the south of Kinnaur, particularly among the Sikh kingdoms began to challenge Mughal authority. The emergence of Bushahr as a distinct principality in upper Kinnaur dates only from this unset- tled period. The key to Bushahr’s rise was cross-border trade, which could be enormously profitable.19 In 1639 Ladakh closed its Kashmir frontier and Kashmir-Tibet trade then shifted to the Sutlej route,20 which Bushahr controlled along with other routes traversing Kinnaur. These included trails north to Kulu or Spiti-Lahaul and others southeast via Chitkul to the upper Ganges (see map 6), notably a difficult route to Nilang on the Jadh Ganga which was “practicable for loaded sheep.”21 Kamru, in the Baspa valley became the centre of the new trade. It was

16 Singh (1989: 57–60, 67). 17 Singh (1989: 50–51); but as Singh (1989: 59) concludes; “[t]hat Bushahr’s claim to such an exalted existence since antiquity was so widely accepted was certainly a remarkable piece of legitimation engineering; available facts point to it having emerged to be taken seriously enough as a state only from the last quarter of the 17th century onwards.” 18 Singh (1989: 69). 19 Mittal (1985: 70) gives the example of an Indian trader returning from Tibet with a maund (approx: 80lbs/ 35kgs) of borax that cost one rupee and sold for 16–20 rupees in Bageshwar. 20 Singh (1989: 57–58). 21 Gerard (1993: 48–49); also see Herbert (1822: 355). Channa (2013) emphasises the social links between Kinnauris and the Jads of Uttarkashi and in particular Nilang. 178 chapter 8 from the Kamru Thakurs or land-owning class that the Bushahr ruling dynasty emerged, but for reasons probably including defensive distance from Tibet and the elites’ growing identification with Sanskritic India, Rampur became the political capital of Bushahr with nearby Sarahan the religious centre. The growth in Bushahr’s power and influence was indicated by the Bushahr- Tibet Trade Treaty of 1685, which followed the Tibet-Ladakh-Mughal war of 1681–1683 in which Bushahr assisted the Tibetans.22 The Treaty gave Bushahr the right of free trade with Tibet and rule over the former Guge territory of upper Kinnaur which had been under Ladakhi rule.23 Bushahr thus faced the need to incorporate a substantial culturally-Tibetan area into its kingdom, which must have been a significant factor in religio-political policies there. Its rulers certainly played off their neighbouring empires, Tibet and Mughal India to maintain their position. Singh gives the example of their claiming a Mughal title while also promoting a tale that the Bushahr rajas incarnated as the Dalai Lamas!24 Legitimacy among the local religious groups was advanced by similar methods. In a letter of 1869 the raja refers to the presiding devta of Kamru, his clan deity Badrinath, as Buddha incarnate while an 1875 letter refers to Badrinath as Krishna incarnate, thus appealing to both Buddhists and Hin- dus.25 In the late 18th century Bushahr also resisted a new power active in the Himalayan regions; the Gurkhas. Accustomed to mountain warfare, the new Kathmandhu valley-based Hindu kingdom rapidly expanded across the Hima- layas to the east and west. But Bushahr survived, more likely than popular accounts of their gaining military victory is that the invaders reached Bushahr but, with their supply lines over-extended, withdrew on the promise of an annual tribute mission.26 The East India Company’s victory over the Gurkhas in 1815 then saw Bushahr come under British rule. It became part of the Pun- jab Hill States but down to 1947 it remained isolated from the plains, and was almost entirely remote from the Independence struggle. The British explored means of access to Tibet via Bushahr, and in 1850 began construction of a Hindustan-Tibet road beside the Sutlej, but it got no further than Chini in the 19th century. Rampur did become the centre of a lucrative trade in shawl-wool, which in the late 1830s “became the most

22 On which see, in particular, Halkias (2009); also see Petech (1947: 292); Ahmad (1968: 340–361). 23 Singh (1989: 17–19, 58). 24 Singh (1989: 88–89). 25 Singh (1989: 87). 26 Singh (1989: 14). kinnaur and adhi kailas 179 important element in the economy of the hill states along the Sutlej”, with the takeover of that trade by the Dogra Sikhs a contributory factor in the 1845 Anglo-Sikh war.27 But the shawl-wool trade then declined, and by the end of 19th century the British search for approaches to Lhasa had shifted to the eastern Himalayas. Bushahr returned to its former peripheral status in the perspective of its neighbouring empires.

The Religious World of Kinnaur

Historical studies of Kinnauri religion are a major lacuna,28 and only in regard to its Buddhist art history are we well-informed. Much of this literature how- ever, situates the region as peripheral to either Tibet or Kashmir and tends to ignore the cultural division between the upper, Tibetanised regions and the wider Indic culture(s) dominant elsewhere in Kinnaur.29 As we have seen, this cultural frontier was also a political divide down to the late 17th cen- tury. Being within the Tibetan world at that time, the Buddhist areas of Kinnaur (which in the most general sense are those to the north and east of the Kinnaur Kailas, and indicated by toponyms in Tibeto-Burman languages), were subject to an early phase of Buddhist propagation in the first millennium ce. This phase is associated with the legendary Buddhist magician-missionary Padmasamb- hava and represented by Ribba temple on the Kinnaur Kailas circumambula- tion route. While locally attributed to that seminal figure in western Tibetan Buddhist history Rinchen Zangpo, and probably visited by him, Ribba is actu- ally earlier, being founded around the 9th century.30 Rinchen Zangpo represents the later phase of missionary endeavour which in western Tibet (including upper Kinnaur), succeeded in firmly establishing Buddhism at a non-elite level. Interestingly, the section of the Kinnaur Kailas parikrama from Ribba to Charang replicates that route that Rinchen Zangpo

27 Lamb (1986: 47–65, quotation from 48). 28 While history is not his primary concern, Singh (1989) provides the best critical history of Kinnaur. Also see the recent (primarily anthropological) studies by Sutherland (1998); Moran (2007:); Berti (2009); and Das (2012). 29 See however, the works of Laxman Thakur. 30 Thakur (2002: 29–44); also see Klimburg-Salter (2002: 1–28), which contains a useful discussion of the history of Buddhism in Kinnaur and the dating of Ribba temple, but again proceeds from the etic position that “Kinnaur existed on the periphery” (2002: 4) a position which tends to obscure or devalue local agency and perspectives. 180 chapter 8 and others followed from Guge to Uddiyana/Swat,31 and a number of Buddhist temples on this route are attributed to him.32 Laxman Thakur, who has carried out extensive fieldwork in the region, concludes on the basis of archaeolog- ical findings that the Kinnaur Kailas parikrama route was in use by the 10th century, and that the establishment of these Buddhist centres was an attempt by Rinchen Zangpo to make use of the ritual, “in much the same way as per- haps he witnessed in the Kailash-Mansrowar pilgrimage in Tibet”.33 Thakur has also pointed out that, “… the two frontier-most villages of Kinnaur—Kunu and Tsarang—in the Tidang valley were on the itinerary of the annual Kailāśa parikrama.”34 Thus, as at Manimahesh, the ritual centre has some associations with territorial definition; the mountain as frontier. Tibetan Buddhist influence remains strong in Bushahr including culturally specific elements of Tibetan beliefs such as that in beyuls35 (‘Hidden Lands’, popularly known in the West via the Shangri la legend). According to one recent guidebook, the local goddess of Chitkul, Mathi Devi (consort of the Kamru Naga deity, which bears the Sanskritised name of Narayan), is the only non-Buddhist deity acknowledged in the parikrama process,36 although the designation of the range as a Kailas mountain is clearly a Śaivite statement within the Sanskritisation process. Given the possible origins in or at least early links with, the Kailas massif as the imagined home of the souls of the dead, it is of interest that for the Buddhists of Charang Yama, the Lord of Death in Hindu and Buddhist tradition, is the presiding deity there.37 The 1971 census determined Kinnaur as 85% Hindu and 14% Buddhist,38 with the Buddhist population still largely inhabiting the regions historically influenced by Tibetan Buddhism and by close relationships between the elite classes on both sides of the modern border.39 The daughter of the Guge ruler, for example, was given in marriage to the Bushahr king after the 1685 Bushahr- Tibet Treaty, and her palace in Sapni (the Guge Rani Ka Mahal: ‘Palace of the

31 Klimburg-Salter (2002: 7). 32 Thakur (2001: 7); also see, Singh (1989: 54). 33 Thakur (2011: 212). 34 Thakur (2001: 7). 35 Das (2012: 33). 36 Sanan and Swadi (1998: 3). 37 Information from Dr. Laxman S. Thakur, Simla, October 2001. 38 Kumar (1981: 12). 39 Genetic tests showed high caste Kalpa, Kothi and Chini people had more Tibetan blood than the lower castes; Singh (1989: 69). This would suggest marriage ties rather than major power imbalances (again casting doubt on the peripheral nature of Kinnaur). kinnaur and adhi kailas 181

Guge Queen’), still survives.40 Close religious links between Kinnaur and Tibet continued until recently. Before the 1962 closing of the Sino-Indian border, Kin- nauri Buddhist monks were primarily orientated not to Ladakh but to Tholing (the largest monastery in western Tibet), and to Lhasa where they travelled for advanced training.41 There were also regular visits to Kinnaur by representa- tives of the various Tibetan sects, while Gerard reports that British rule in India stimulated visits by another group of Tibetans, the Khampas. While specifi- cally referring to an origin in Kham (eastern Tibet), the term tended to be used by colonial officials in a generic sense to describe itinerant Tibetans,42 whose numbers included both pilgrims and religious specialists along with those (an overlapping category), who made a living as entertainers, as traders, as beggars, and as thieves.43

There is a sect of wandering Tartars called Khampa, who are in some respects similar to the Jogees of Hindoostan. They visit the sacred places and many of them subsist wholly by begging. Some are very humorous fellows [who sing and dance] … Since the British Government have got possession of the Hills, Khampas come down in crowds to visit the holy places to the westward.44

Those “holy places” included Uddiyana, Jvalamukhi, Triloknath and, in more recent times, Rewalsar.45 But the Kinnaur Kailas, despite its local Buddhist component has not been sanctified within Tibetan understandings of the holy places of India. To an extent this may be explained by the medieval Tibetan visitors to the western Himalayas (see Chapter 12) being followers of the Kargyu school, who deliberately sought to follow in the footsteps of their pioneer Orgyenpa, not Rinchen Zangpo. Even in more recent times, the famous Tibetan

40 Handa (2004: 174) states that the Guge Rani was killed at the instigation of the Bushahr king’s first wife. 41 Gerard (1993: 49) states that Tholing was “resorted to by all the Lamas of Koonawurv … [sic].” While Ladakhi monasteries predominantly belong to the Kargyu sect, Tholing is Geluk (the leading Tibetan Buddhist sect: see Chapters 11 & 12) and a ‘daughter monastery’ of Lhasa’s largest such institution, Drepung, thus Kinnauri Buddhism primarily followed the Geluk tradition. 42 On whom see Sharma & Sankhyan (1996: 314–320). 43 For a discussion of the term in the context of its use in Nepal, see Rauber (1980). 44 Gerard (1993: 117). 45 The earliest visits to this site by Tibetan pilgrims appear to be mid-19th century; Huber (2000: n. 50 118). 182 chapter 8 intellectual Gendun Chöphel (1903–1951), from a Nyingma family but ordained as a Geluk monk, knew of the travels of Orgyenpa and his successors and followed many of the routes they took. But he makes no mention of Kinnaur Kailas in his Guide to India.46 There was however, one Tibetan who travelled extensively in Kinnaur in the early 20th century; a renowned Bön monk, Khyung-sprul Jigme Namkhai Dorje (1897–1955). In 1924 he first reached Kinnaur from Rewalsar during his travels to the sacred sites of India and, after circumambulating Kailas-Manasarovar in 1930 he returned and stayed in Kinnaur for four years. He was able to attract wealthy and influential patrons, not least through his successful rituals for the Bushahr queen, who sought a son. Given that as Per Kværne observes, “the sources for the history of the religion of this region seem to be particularly meagre”, Khyung-sprul is of most relevance to us here in regard to;

his main activity … [which] was to suppress the custom of worshipping the local mountain gods (yul-lha) with animal sacrifices … In all he bound sixteen yul-lha of Kinnaur in an oath not to demand animal sacrifices and made them promise to be protectors of Bon.47

Kværne’s study of Khyung-sprul confirms that mountain gods are an aspect of the Tibetanised Kinnauri world and that while cultural meaning is not attached to the Kinnaur Kailas pilgrimage in wider Tibetan Buddhism it is within a local Buddhist understanding. Khyung-sprul had mastered the rituals understood to bind the mountain gods to the service of a new faith, transforming them from independent spiritual powers (in external understanding, hostile and malev- olent), into protectors of a faith that claimed relevance beyond any particular territory, a World-religion, in this case, Bön.48 In the Buddhist world, the use of this ritual power was most commonly associated with Padmasambhava, who is understood to have performed such rites not only in Tibet, but throughout the Himalayas. But this was also a power possessed by Bön practitioners, for it arose from a specific ritual performance rather than from a specific belief system.

46 See Huber (2000). 47 Kværne (1998: 71–84, quotations from 72, 78); also see Blezer (2007). 48 Systemised Bön, the form with which we are primarily concerned, is defined here as a World-religion on the grounds of its Salvationist doctrines being articulated as applicable and accessible to all, in contrast to the geographical restrictions of local territorial deities and deities of place. Modern Bön thus contains missionary elements absent from local belief systems, which make no claim to their deities’ power over or even accessibility to foreigners in distant lands. kinnaur and adhi kailas 183

Khyung-sprul’s activities raise the question of whether early western Tibetan influence in upper Kinnaur included Bön. Khyung-sprul understood Kinnaur as having a connection with Zhang-zhung, which the Bönpo saw as their early homeland (an issue we return to in Chapter 13). As a Bön monk he sought to promote that tradition, and understood himself to be reclaiming a lost heritage rather than acting as a missionary in a new land. Yet there is no evidence for the existence of Bön in Kinnaur in any systemised monastic form prior to the establishment of Menri monastery at Dolanji (near Solan), in the late 1960s. (It is of interest that the annual ceremonies there include a commemoration ritual for Khyung-sprul.49) Khyung-sprul’s activities seem to be within a culturally Tibetan world-view rather than a specific tradition. Tibetan sacred sites were understood to be divinely established and thus revealed or rediscovered rather than created or discovered. While “profoundly committed to the Bon religion and tradi- tion”,50 Khyung-sprul himself seems to represent the late 19th century Rimed movement, a manifestation of that tendency within Tibetan thought that de- emphasised sectarian divisions. His openness, however, should be seen in the context of his field of operations in Kinnaur, which was within a primarily Tibetan Buddhist world. Khyung-sprul did not confront Buddhism. Following the wider pattern of World-religion expansion strategies he confronted local powers, local deities. He did so however, only within the Buddhist areas of Kinnaur. In the rest of Kinnaur local territorial deities seem to have held sway over villages and settlements rather than the uninhabited mountain peaks, with belief in yullha mountains apparently restricted to the Buddhist population.51 The characteristically Tibetan annual worship of yullhas by a one-day moun- tainside ritual52 reportedly still survives in at least one Tibetanised village, Pilo/Spilo/Spello, beyond Morang in the direction of the Tibetan frontier. S.C. Bajpai describes a one-day celebration there rather than a parikrama;

49 Ramble and Kind (2003: 736–738). 50 Kværne (1998: 83). 51 But see Sharma (2003: 357). Sharma describes a Himachal Pradesh system of knowledge preserved in manuscripts held by local families. These are in local scripts and interpreted by those families. They include myths and legends, many of which are Sanskritised and support his contention that they date to the medieval period and have links to the Naths. He briefly notes that they also include “invocation mantras of the mountain goddesses (Yoginis or Joginis) running into several pages … called ‘jugnao.”’ Further investigation is needed. 52 On which see for example, Buffetrille (1998). 184 chapter 8

The Kailas-zalmo, or “visit to the Kailas mountain”, is celebrated at Pilo or Spilo, in Sjuwa pargana, on any auspicious day in Har fixed at the will of the zamindars, and lasts one day. Worship of the Kailas mountain is performed.53

The Local World

Throughout Kinnaur, Western Himalayan Cultural Complex traditions of the type discussed in regard to Kaplas and Manimahesh Kailas’s remain promi- nent. Territorial deities are the most-remarked upon feature of the regional belief system. Each village or group of villages has its deity (or deities), often related and variously termed devtā or deota. They speak through a possessed human medium on specific occasions and in establishing their heroic history the deities, like a royal eulogy or a mahatmya, begin by establishing the lin- eage that is their authority.54 These lineages indicate however, that just as many of the local deities of Chamba and Bhadrawah are not in the historical sense local, so too are the territorial deities of Kinnaur largely immigrants. The ori- gin myths of many of the deities describe them as native to places in Tibet. Thus Sri Badrinath narrates his origins in Guge as does his wife Mathi, the pre- siding deity at Chitkul. At Rakchham in the Baspa valley, the deity Sounige claims origins at Kailas-Manasarovar (where no trace of him now seems to remain).55 In contrast to the territorial deities around Chamba, those of Kinnaur are loosely ranked in a hierarchy headed by Bhimakali of Sarahan, Sri Badrinath, and the Maheshwar brothers of Sungra, Katgaon and Chagaon.56 This may, however, represent an elite perspective, or ordering principle, rather than a local understanding. While the presiding deity in the various Himalayan states is usually the ruler’s clan deity,in Bushahr the deity Bhimakali was supreme and the raja considered her earthly agent, although the raja’s personal clan deity was Sri Badrinath (of Kamru). This suggests that the presiding deities represent

53 Bajpai (1991: 211): The account might suggest some accommodation given that the title ‘Kailas’ is clearly Indic, but we are reminded that Herbert’s Buddhist monk apparently used the term. 54 As Lecomte-Tilouine (2009: 173, and see 190–195) states, “the social meaning of these texts … [is] more important than their obscure narrative meaning.” 55 oioc, Emerson (Treatise: chapters 1, 2); Singh (1989: 48–49); Jerath (2001: 136). 56 Singh (1990: 250). On the preceding page is a list of the 32 principal devis and devtas of Kinnaur. kinnaur and adhi kailas 185 a later phase of development, a marker of political processes in Bushahr where, “the village goddess Bhima Kali still controls only the deities of those portions of the State which were conquered at an early date.”57 Given that the Bushahr raja was Bhimakali’s earthly representative, opposi- tion to the raja meant opposition to the presiding regional deity and vice-versa, so this local deity system was closely linked to existing social structures. Con- trol over the local deities was exercised though patronage of their functionaries and represented by the deities’ being taken on regular tours of outlying districts. The deities themselves certainly served elite interests, acting as landowners and moneylenders through earthly agents. But as Jogishwar Singh notes, in the long-term this acted against outside money-lenders taking local land and wealth, while the local power enjoyed by territorial gods of small communities had a democratic aspect; “[i]n a sense the village godlings represented the collective will of the people against unchecked absolutism by the Raja.”58 There was therefore, a complex religio-political balance and while it may only have been ordered in this form in the 17th century it represents the systemisation of earlier tendencies within Western Himalayan Culture.59 While more academic attention has been devoted to the Kinnauri system of territorial deities,60 the Naga cult is also strong there. Bushahr was named after Bushahru Naga, whose temple is near Rampur, and villages in the Kinnaur Kailas area such as Chini, Kalpa, Chitkul, and Kamru all have their local Naga deity, for in Kinnaur Nagas are village specific. The Kamru Naga deity Badri- nath is considered the eldest of the region, the name indicating the influence of the cult of that deity in Garwhal/Tibet.61 That they have been partially absorbed into the Brahmanical system is seen in that many of the Naga deities are known here by the Vaisnavite appellation Narayan/Naraynus. The Sarahan deity Pir Naga is another partially absorbed into modern Hinduism, for in certain rituals he shares the central role at

57 oioc, Emerson (Treatise: chapter 3; 38–39). 58 Singh (1990: 248); also see Singh (1989: 86–90); Sharma (1990: 131–140); Berti (2009). 59 Although local belief in the power of the Kalis, female “fairy-like mountain spirits” who control the weather seems more indicative of links to the Persian Peri than Tibetan female spirits such as dakini; on Peri and dakini see Templeman (2002). On the Kalis, see Das (2009: 11, 20). 60 A definitive critical account is a lacuna; for a recent anthropological enquiry see Raza (c. 2008: 117–145). 61 The suffix nath would suggest a date in the second millennium, but the name may be a later Sanskritised version of the original. 186 chapter 8 the Bhimakali temple, the only one in the region with Brahman pujaris.62 In the absence of a lake associated with the pilgrimage, however, Nagas are not central to the Kinnaur Kailas yatra as they are to the Kaplash and Manimahesh rituals. Kinnaur Nagas, in contrast, are believed to worship at their ancestral home, Duling lake, on the Buran pass south of the Baspa valley, a place not obviously connected to the Kinnaur sacred centre but one more reflective of the older non-Sanskritic concept of placing deities as protectors on the frontiers of the settled territory.63 Situating the Kinnaur Kailas within this local world is problematic. The absence of a lake or of any obvious territorial deity or Naga associated with the site or its pilgrimage separates it in form from the Kaplas and Manima- hesh Kailas pilgrimages. It is to a mountain, but no single peak is uniquely distinctive. The toponym Raldang seems historically prominent and may have referred to such a presiding deity, but the lengthy parikrama route and the his- torical background indicate that the ritual is best contextualised in terms of territorial inclusion, with a focus on the frontiers of the sacred space rather than its centre. Given the need to include diverse population groups and belief systems into Bushahr, the apparently imprecise nature of the centre may have been sufficiently fluid to allow each group to worship that centre within its own conceptions, and in the wider sense to provide a sacred space and a protective mountain on the frontier.

A Śaivite Mountain

While historically Kinnaur Kailas pilgrimage is more closely associated with Buddhism than Hinduism,64 its modern promotion is primarily in a Śaivite context. The basic Śaivite myth concerns its sacred origins;

[L]ocal legend says that the people of this part of the Himalaya were finding it difficult to go to Kailash every time for pilgrimage. So an ancient

62 We have noted that there is an opposite case at the Bhadrakālī temple in Chamba, where Jogis acted as pujaris to avoid the impurity of the sacrificial rites accruing to the Brahmans. 63 Handa (2004: 55, 66, 107, 144–149, 166–170; for a list of the Kinnaur Naga deities, see 344–345). 64 There is no mythology linking places on the Kinnaur pilgrimage with the classical Indic world aside from Chitkul’s claim to a Pandava fort. Interestingly, Raza (c. 2008: 130) detects a modern drift to Buddhism in Morang district attributable to the activities and profile of the current Dalai Lama. kinnaur and adhi kailas 187

king and a devotee of Shiva got [Śiva’s agreement] and brought down the Kailash to Raldang in Kinnaur and made it the home of all the deotas …65

Gerard’s account of the myth hinted at a humble status for the Kinnaur Kailas;

[It] must not be confused with the other Kylas near Munsurowur [i.e.: Manasarovar]. The people say this last is by far the highest, and the Reedung [i.e., Raldang] Kylas is only a piece of it, which was removed by the gods to please a very pious devotee, who lived opposite to Reedung on the right bank of the Sutlej.66

Captain Madden also noted this myth, adding that, “it is considered meritori- ous to perambulate the mountain, keeping it always to the right hand … the sublimity and immaculate purity of the Ch’hota Kylas [sic!] render it no mean emblem of ‘the high and holy one that inhabiteth eternity …”’67 Today, Śiva is said to reside there in the winter months enjoying the local cannabis before travelling to Kailas-Manasarovar for the summer. In a related myth encompassing, acknowledging, but subjugating Kinnaur’s local deities, they are said to assemble annually in the month of Magha (Jan/Feb) for an annual conference with Śiva.68 Yet not only is there no recognised bathing place, but despite the Śaivite mythology overlaid onto the site, there are no associated images or structures and no wider complementary Śaivite sacred geography, no associated chain of sacred places, or shrines radiating out from the centre. This lack of development may be due to the impact of renunciates being considerably less pronounced in Kinnaur than in Chamba. In the early 19th century J.B. Fraser recorded them as common in Rampur, but that was proba- bly due to the unsettled conditions of the time, for renunciates were, “the only people who seem to have escaped the desolation” attributed to Gurkha depre- dations.69 They are not mentioned by other travellers and Bushahr lacks Nath shrines, pithas, or other natural gathering places for renouncers, who are still not prominent there today.70

65 Thakur (1997: 181). 66 Gerard (1993: 13). 67 Madden (1846: 112). 68 Bharti (2003: 117). 69 Fraser (1820: 258). 70 One significant account of renunciate agency does exist, however, in the form of a legend attributing the naming of Bhimakali to a Giri renunciate, one Bhimagiri. Returning from 188 chapter 8

The strong, cohesive Kinnauri identity and its distinct dialect may have made it a less congenial environment for renouncers than Chamba. But the presence of Brahman pujaris at only one Bushahr temple suggests that royal patronage of Sanskritic agencies was less generous overall, with more resources devoted to the territorial deity system. Bushahr royalty must have been aware of the considerable financial benefits to be derived from pilgrimage sites on their territory. In the early 19th century, for example, offerings at nearby Jvalamukhi, which were part of the Kangra raja’s revenue, amounted to 25–30,000 pounds (sterling) annually.71 And (not least to raise their status at a Pan-Indian level), Bushahr rulers had made pilgrimages to Sanskritic sites such as Badrinath and Varanasi, while also maintaining sanctity in Buddhist perception through visits to Tholing.72 The limited patronage they offered to Śaivite renunciates and institutions may thus reflect policies designed to balance the interests of the majority Western Himalayan Cultural Complex with a Buddhist population on their frontier with Buddhist Tibet. Kinnaur Kailas was used as a unifying symbol to incorporate differing communities within their realm, a policy consistent with that followed elsewhere in the Indian Himalayas. Certainly the standard ritual journey is inclusive of both predominantly Buddhist and predominantly Hindu areas of Kinnaur. Thus it can be seen as embracing a cultural frontier and the different groups resident there. While we might speculate that the names Raldang (or Jorkhanden) derive from a deity now forgotten, a local deity associated with Kinnaur Kailas in either the Indic or Tibetan traditions is not apparent.73 Nagas—deprived of their traditional watery home—and renunciates—lacking patronage except in occasional cases of highly placed individuals such as Khyung-sprul, also have little or no role in the Kinnaur Kailas history. There is no mahatmya of the site,74 and while it is articulated as a Śaivite pilgrimage, there is no

Kailas, he stayed the night at Sarahan and the next morning found his staff (daṇḍa), which he had dedicated to Durga, too heavy to lift. Recognising this as a sign that he should remain there, Bhimagiri established the worship of the goddess at Sarahan and she became known after him, as Bhimakali; Bharti (2003: 108). 71 Gerard (1993: 130). Sharma points out that the Sanskritised shrines developed a whole economic world around them—image makers and sellers, water carriers, musicians, etc; Sharma (1996: 87). 72 Bharti (2003: 15, 84–85). 73 Inquires might, however, be made with monks at Charang monastery. 74 Nor, when I undertook the parikrama myself in 2001 did I obtain any significant new information about the pilgrimage (other than that Charang monastery has a resident kinnaur and adhi kailas 189 particular mythological association with classical Hindu concepts beyond an incidental acknowledgement of Kailas-Manasarovar as the original abode of Śiva. The mountain seems neither Buddhacised nor fully Hinduised in the sense of any strong claim to Pan-Indian relevance. The lack of an associated lake is particularly problematic, for bathing is the classic Hindu rite at pilgrimage sites. We enter the realms of speculation in attempting to date the sacred origins of the site. The earliest sacralities may attach to the Khaskar outcrop rather than to the higher peaks behind, suggesting a close association with the Chini populace. The mountains’ role as the place of the souls of the dead seems similar to that of Manimahesh (although in Kinnaur this element does not seem to have attracted the cemetery-dwelling ascetic traditions). It is certainly possible that the pilgrimage originated with or was systemised by Rinchen Zangpo, and that he transformed an annual ritual of worship on the side of the mountain directed to a (presumably named?) yullha into a Buddhist ritual involving circumambulation. With the Bushahr take-over of the culturally Tibetan area after 1684, that Buddhist ritual would then have been acknowledged by the Bushahr rulers as a socially incorporating device enhanced by a Śaivite gloss to appeal to a southern constituency.We might then date the introduction of Śaivite mythology to that late 17th century period. It is certainly difficult to locate any connection between the pilgrimage and the historically predominant local territorial deity system in Bushahr. Places of worship on the full parikrama route are predominantly Buddhist, while the site is now primarily interpreted from the Śaivite perspective. Its socially incorpo- rative function and World-religion status thus seems predominant. We might still identify Kinnaur Kailas in a separate context as a traditional frontier pro- tector marking the cultural frontier between the Western Himalayan Cultural Complex and the Tibetan world, and acting as a symbolic defensive marker in the manner of Manimahesh. If this is the case, however, it remains unclear whether it was intended by Tibetans to mark the extent of their world, or by Bushahris to mark their frontier with the Tibetans.

order of Buddhist nuns). While the Tourist office in Rekong Peo advise that Inner Line permits are not required on the circuit, access to the route is at the whim of the Indian Army, who refused access to a Kinnauri-speaking academic seeking to stay some months there, while allowing that same day, myself and two companions (with two porters) to undertake the trek, as they have other trekkers. Being indebted to the military and civil officials who have allowed me access to these regions, I did not feel it was appropriate to undertake extensive enquiries along the route. 190 chapter 8

The Adhi/Baba/Chhota Kailash

The 20,311ft (6191 metres) Adhi Kailas—that is, ‘Half’ Kailas—also known as the Chhota (Hindi: ‘Little’), or (particularly in mountaineering circles), ‘Baba’ Kailas—is situated in India’s Kumaon Himalaya (comprising the Pithoragarh, Almora and Nainital districts of ). It is commonly reached from Almora and Pittoragarh via the trekking route up the western bank of the Kali river, which is that taken by the official Indian pilgrimage parties visiting Kailas-Manasarovar today. Some 20 kilometres short of the Lipu-lekh pass to Tibet, the route to Adhi Kailas turns west via Kuti (Kuthi/Kunti) village towards three other passes leading to Tibet (the Mangshang, Lampiya Dhura, and Darma). The Adhi Kailas is part of a range separating the Kuti valley from the Darma valley (reached by the Sinla pass). It has an associated ‘Parvati’ lake situated above Jyolingkong, a patch of level ground below the mountain (see map 7). Pilgrims carry out a parikrama of the lake and a small temple dedicated to Śiva and Parvati was constructed there around the year 2001. During the summer season the Indian Army now provides food and lodging for pilgrims in large fibre huts at Jyolingkong. Until the Gurkha invasion of Tibet in 1788, Kuti and surrounding villages were part of Tibet.75 The British then took the area from the Gurkhas after 1815 and later included it in British Garwal.76 Being close to the Tibetan border, access to the region was restricted until recently and historically it was, like Bhadrawah, seldom-visited even by Indian officials. Colonial officer Henry Strachey, who passed through the area en route to Kailas-Manasarovar in 1846, seems to be the first European to mention it.77 He noted in regard to a “small pool” at Jyolingkong, that;

[T]here are traditions that some Raja of Byáns [Kumaon] in days of yore indulged his fancy by calling the puddle Mántaláw, and one of the neighbouring snowy peaks (of no remarkable figure), Kailás, after the great originals of these names in Húndés [Tibet].78

75 Sarcar (1931: 127) citing the “Introduction” to The Narratives of the mission of George Bogle to Tibet and of the journey of Thomas Manning to Lhasa, 1879. Sir Clarence Markham (ed.) 1879. London: Trubner; lxxvi. 76 Atkinson (1884: 73–74). He notes that after their defeat in 1815 Nepal requested that Kuti remain their territory, but in 1817 the British disallowed their claim. 77 Richardson (1998: 462) notes that the Armenian merchant Hovhannes Joughayetsi trav- elled to Lhasa via Kuti in 1686. 78 Strachey (1848: 31). kinnaur and adhi kailas 191

Strachey’s account does indicate that the toponym ‘Kailas’ is of some antiq- uity, but the mountain was referred to by various names. His brother John Strachey, in his précis of the Manasakhanda, calls the mountain by the oth- erwise unknown toponym ‘Puloman.’79 But the mountain was also known as Jyolanka after the plain below it. The premier Hindi-language regional history refers to Jyolanka as a “bordering mountain”,80 again suggesting what seems to be an ancient understanding of the role of particular mountains as frontier protectors. In 1908 the similarity of the mountain and lake at Jyolingkong to Kailas- Manasarovar was again noted by the well-educated Marathi renouncer Bhag- wan Sri Hamsa, although he does not name them.81 But in his 1949 guidebook Kailas-Manasarovar, Swami Pranavananda, who travelled extensively in these regions in the 1920s–1940s, footnotes them under the title of “the Chhota Kailas and Manasarovar [lake]”, adding that they, “are worth visiting.”82 The Swiss geologists Gansser and Heim based themselves in Kuti for some weeks in 1936. Augusto Gansser accompanied two Kuti villagers making what was apparently a routine 12 day trek to Kailas-Manasarovar to buy sheep and goats. The animals were considered of sacred status due their being raised in a sacred homeland, and they were required (presumably, although not stated, for sacrifice), at what the Swiss called Kuti’s ‘festival of the dead’. Gansser also refers to the mountain under which he sheltered at Jyolingkong; “[s]o closely did this resemble Kailas in shape that our Kuti men called it Kailas Baba.”83 Yet there is no suggestion in any of these sources that this mountain was sacred, and it is not sacred today to the people of the Kuti or Darma valleys. There is no annual festival associated with the mountain, no mahatmya, and no deeply rooted mythology. Kuti, like the villages of the Darma valley, seems to follow the traditions of the Western Himalayan Cultural Complex. Prominent are the tall shrine poles (darcho), single or in clusters, bedecked with pieces

79 It is identified as a “peak in the dividing range between Darma and Byáns, at the foot of which is a small lake known as Mán taláo or Byankshiti between the Jhúling and the Rárub Yankti.” Strachey’s précis is contained within Atkinson (1884: n. 14 44). 80 The standard history of the area, Pande (1993: 50–51), confirms that Chhota Kailas is the mountain named Jyolanka: “Between Vyans and Darma lies the bordering mountain Jylolanka.” It does not mention this place as sacred. Also see Moran (2003: 82–92); and see the related expedition report (Report: 12) at the Alpine Club library in London (copy marked accession no. 28186). 81 Hamsa (1998: 197–198). 82 Pranavananda (1983: n. 2 119). 83 Heim and Gansser (2000: 77–90, 112–118, also see photos, 109, 137–139). 192 chapter 8 of cloth of varying colours. The most of prominent of these today is around 35–40 feet (10 metres) high (see photo 11). Kuti has its own local deity, Mapang Gulach, worshipped by an annual festival which includes sacrifices. Depicted as a fearsome six-armed, snout-nosed deity, wearing a garland of skulls (see photo 13), the distinctive figure of Mapang Gulach is served by Rajāju, said to ride around the village at night to ensure its security.84 There is however, a local Sanskritic tradition concerning Kuti. The name of the village is attributed to the goddess Kuti/Kunti, regarded as the mother of the Pandavas, whose actual ‘historical’ presence is testified to in several ways.85 They are said to have made a hut here—which provides an alternative explanation for the village toponym (Skrt: kuti = hut)—and there is another local mountain deity known as Pandu Parvat. In addition, on the approach to Kuti is a promontory known as Pandakila (Skrt: kila = fort; i.e., ‘Panda(va) fort’). This Sanskritic tradition co-exists however, with the contrary origin myths of the people, who state that they migrated here from Tibet 4–500 years ago. There are currently no Buddhists living in the area or undertaking pilgrim- ages to the Adhi Kailas, but it is just 16 kilometres from the Tibetan frontier and there were close links with the Kailas-Manasarovar region as late as the mid-20th century. The language and culture of the Darma valley have histori- cal links with Tibetan culture,86 as does the belief system among the people of

84 Oral sources; Kuti village, 2003. Due allowance should be made for my approximate transcriptions from the local dialect, inadequate translation, and limited number of local informants consulted. Despite the deity in a representation within one shrine being depicted as having six arms and a garland of skulls Mapang Gulach does not appear to me Sanskritic, but if not entirely local, is more likely Tibetan-influenced (though without apparent connection to Mapang, the Klu of Manasarovar). I am uncertain if this deity is related to the Jad deity Me Parang, on which see Channa (2013: esp. 150–151, n. 16 176), who notes the similarity of Me Parang to the Vedic deity Rudra. Above the doors of many local houses are carved wooden images of splay-legged naked female figures displaying their sexual organs; the position known as mukhagaruḍa (Grey 2007: n. 11 198). I have consulted a number of Himalayan art specialists in regard to both this figure and that of Mapang Gulach, but they could suggest no close parallels to these images, which may thus be uniquely local; see, however, for early European parallels, Dexter & Mair (2010). I am preparing an article on these figures. 85 On the Pandavas continuing ‘presence’ among the Jad peoples, see Channa (2013: 105). 86 Yael Bentor, email communication, 8 May 2006. Charles Sherring (1974/1906: 204–205) noted in 1906 that the harsh regime of the Taklakot Dzongpons (Governors) had driven many to emigrate across the mountains from Tibet to the Darma pargana; “the population of pargana Darma has almost doubled in the last thirty years, whereas in Taklakot there has been a decrease.” Darma is a Tibeto-Burman language; on which see Krishan (2001: 347–400). kinnaur and adhi kailas 193 the smaller village of Gunji (east of Kuti), who also claim origins in Tibet, albeit apparently from a different region. Their annual Jumlee festival in early August (a week before the Kuti festival), seeks to repel a demon from Tibet said to have once killed and robbed local people. Amidst these beliefs an earlier layer of culture is indicated by the presence in the Kuti valley of standing stones of the type common to the western Tibetan plateau,87 and to numerous sites in Europe and southern Russia. These are, however, objects that are outside of local cultural mythology, nameless, and explained only as tests of strength, formerly lifted by strong men of the village.88 Their original meaning has been lost and any link to earlier cultures and migrations is largely speculative.

Sanskritisation

We may conclude that the association of the Adhi Kailas with Kailas-Manasaro- var derives solely from its perceived similarity of appearance! Although that identification is of some antiquity, the understanding that it is a sacred site is very recent, and not within local beliefs. The agents behind this process include the State tourist board as well as the Indian Army (which maintains a small camp at Kuti). The Adhi Kailas is promoted as an attractive pilgrimage and trekking destination, with army medical assistance available if necessary. Sanskritic Hinduism is largely a recent phenomena here, but its rapid pro- gress can be clearly observed. Local shrines are being transformed into small rock or concrete temples, the most prominent of which still features an image of Mapang Gulach.89 But the local deity is surrounded by Sanskritic symbols of worship—bells, tridents, and foremost a Sanskrit mantra on white marble contests centrality with the deity (see photo 13). In addition, the lake, referred to by Pranavananda as ‘Manasarovar’90 has now acquired the name ‘Parvati’, and on its shores the new temple features modern images of Śiva and Parvati. (It is thus the only Kailas mountain that has a temple at the sacred complex.)

87 The most prominent is 7ft (2.1mtrs) high, and 18×24 inches wide. Free of markings (other than recent Hindu graffiti), and well set into the ground, this pillar is part of a complex in the upper valley with several smaller stones on an eastwest/northsouth axis also surviving. 88 Channa (2013: n. 1 133–134) notes similar stones in Kinnaur that are associated with Bima, the strongman of the Pandavas, but here no such association has developed. 89 The transcription is approximate. 90 While given this name on local maps, the Indian Government map, n.h. 44, dated 1977, Mānasarowar, 1: 1,000,000 scale, identifies “Kuthi” and “Joylingkong”, but not (Adhi) Kailas or Parvati lake. I am indebted to David Templeman for this map. 194 chapter 8

Śaivite mythology has developed two associations, one in which Śiva bathes there while visiting Kailas-Manasarovar, the other that Śiva is sometimes at Adhi Kailas and sometimes at Kailas-Manasarovar. No doubt the mythologies will increase in scope, but to date there seems no explicit association or absorp- tion of Mapang Gulach or other local deities by Śiva or any other Sanskritic deity. Yet Sanskritisation is not entirely new to the region, an earlier phase in the first half of the 20th century may be identified. At that time a number of prominent renunciates and religious leaders travelled to western Tibet via Gunji and the Lipu Lekh. They exerted a Sankritising influence on the people, if not immediately on the place, for ideology precedes the transformation of sacred landscape. This Sankritising role was most explicit in the case of Swami Sri 1008 Narayan (to whom Pranavananda’s classic Kailas-Manasarovar guide-book was dedicated). After returning from his first visit there in 1935, the Swami established the Narayan ashram above Tawaghat (apparently on the site of a moribund Ra- makrishna mission, which may represent a preliminary stage of the process).91 The Swami, originally from Karnataka, died aged just 46, but is credited with using the ashram to transform the area, establishing a number of schools, col- leges, and dispensaries, and promoting Bhotia cultural identity. But as one commentator writes; “I think the main thing he did was to bring awareness of Hinduism in this area and taught people to say Om Namah Shivaya …”92 Even in the more isolated Kuti and Gunji villages the people there are increasingly learning to identify themselves as Hindu and consequently to interpret the landscape around them from a Hindu perspective. Their local pro- tective deity retains immediate relevance, but their modes of worship expand to embrace Hindu ritual practices and they accept the process by which Śiva and his multiple associations are located within their physical and conceptual world. In that their own localised identity and culture is not (obviously and directly) threatened, and that the Śaivite world links them to a wider modern India, the emergence of Adhi Kailas as a pilgrimage place complements their economic and social uplift. It is thus an attractive and sought-after social devel- opment, which offers direct economic benefits (though retail trade, and the provision of services). In this instance, therefore, there is evidence to support

91 The head of the Ashram is now a woman, Sri Gangotridevi, from the local (now-abandoned due to subsidence) village of Garbyang; Puranik (1999: 34). 92 Berar (1999: 47); emphasis added. Om Namah Śivaya is perhaps the most common Śaivite mantra. kinnaur and adhi kailas 195

Srinivas’s model of Sanskritisation; the adoption by low status groups of the Sanskritic modes and values as a means of status uplift. At present, however, the valley lacks resident Brahman priests or renouncers able to link their traditions to those of wider Śaivite Hinduism through ritual, myth and legend. The Adhi Kailas may thus be seen as a model of the primary stage of the process by which sacred mountain pilgrimage sites are constructed within the Indic world-religions, but it is also the case that Śaivite renunciate agency in introducing this model is seemingly absent here. It may however, also be a site in which temple Brahmans taking advantage of the ease of modern travel are able to move directly into a sacred mountain site without the precedent of ascetic intermediaries.

Some Conclusions

We have seen that the term ‘Kailas’ was used in two contexts in the Indic Himalayas. It could simply mean ‘the snowy mountains’ and in that broad sense could even be used by a Kinnauri Buddhist, as Herbert recounts. But it was specifically applied to a mountain that was home—for at least part of the year—to Śiva. Because Śiva could live anywhere, bestowing the title ‘Kailas’ onto another mountain did not contest the loosely systemised set of meanings and complex of ideas, symbols, and mythological constructions associated with Kailas-Manasarovar or other sites. It was a separate statement. It was also a key Śaivite missionary strategy associated with a lineage deriving from vratya and Atharvavedic roots, which manifested firstly within Nath and later in Giri orders of renunciates; charismatic individuals who were ritually equipped to transform what they saw as the impure world of the Western Himalayan Cultural Complex into the pure world of the Śaivite Hindu. These Kailas mountains were not established as substitutes for the ‘real thing’. Even in the case of the Adhi Kailas, substitution (the use of a local site due to the inconvenience or impossibility of visiting the ‘real’ site), does not seem a primary factor in its identification as a Kailas mountain. Promotion of the Adhi Kailas only began a decade or more after the 1962–1982 period when the closing of the Sino-Indian border prevented all Indians from visiting Kailas-Manasarovar. In its recent promotion of the site the local government tourist department does implicitly seek economic and social benefit to the area, and there is a sense that those unable to cross into Tibet might visit the Adhi Kailas instead. But the missionary impulse is the driving force in the Hinduisation of the area and the primary aim of creating a sacred Adhi Kailas seems to be to make it a Sanskritic centre in its own right, in order to convert the 196 chapter 8 local population to the Hindu worldview of the landscape, rather than seeing it through the perspective of the local belief system. It is therefore, primarily a culturally unifying device imposed by the elite classes. Another common etic explanation for these sites, retaining resources within the local area, also seems problematic. In pre-modern, particularly pre-railway India, journeys to Pan-Indian rather than regional sacred sites were predom- inantly undertaken by religious practitioners or, on occasion by royalty. The economically productive social classes were largely restricted to undertaking local or regional pilgrimages and in the case of the western Himalayas the strong local culture mitigated against the undertaking of Sanskritic pilgrimages by large sections of the population. While the major sub-continental sacred sites certainly attracted considerable resources from Pan-Indian pilgrims, this does not necessarily imply that the economic benefits of pilgrimage were one- way, particularly given the tendency for pilgrims to engage in trade (including obtaining items at sacred sites, such as Ganges water, that were far more valu- able in their local area). The economic benefits of the creation of new pilgrimage places are obvi- ously welcome at the local level93 (although that benefit may be compromised by large numbers of pilgrims subsisting on charity). But we lack studies of the relevant economic processes that would enable us to properly evaluate the extent to which a new site actually enables the retention of local resources. In any case the cultural significance of enhanced social status arising from the presence of a major sacred site in the vicinity, or the compensatory benefits to pilgrims in terms of status gain on their return, complicates any economic determinist conclusion. It is also difficult to attribute the articulation of these mountains as Kailas to the desire of the Himalayan elites for enhanced Pan-Indian status. Mahesh Sharma demonstrates that the Chamba rulers used the Sanskritisation process in their claim to Pan-Indian status for Manimahesh—and such strategies were a common device by the western Himalayan kingdoms in their relations with plains India. But we cannot detect any actual impact on Pan-Indian under- standings of Kailas (Manasarovar or Manimahesh) as a result of Chamba’s ini- tiatives. It was not textually acknowledged in wider India for example, nor did it attract the sacred journeys of royalty from other regions. Only in the later—and very recent phase—of Hindu modernity have the various western Himalayan claims to a Kailas mountain actually enjoyed a resonance beyond the local. They are not included in any authoritative Pan-Indian texts and out-

93 This was the case with persons I questioned on this issue in the Adhi Kailas region. kinnaur and adhi kailas 197 siders (Tibetan religious practitioners or Mughal scribes, for example), made no mention of these sites. It was only with the coming of the British that these Kailas mountains began to be modestly noted at a national level and only in recent decades have two of them been Hinduised in the context of a modern Vedanta pilgrimage site. The introduction or the sighting of a Kailas mountain within a particular territory can be located within a wider process in which the primary intention is the missionary impulse; gaining new believers through the sacralisation of their territory. That process has a long history and seems to be an elite strategy involving an alliance of royal and religious authorities. We have seen this process operating in earlier forms with the advance of Indo-European culture in the Vedic period, when new territory was ritually transformed from the wild and impure unknown into the tamed, pure and known world of the Vedic peoples. This process has continued in outline in the Himalayas down to today. The process of sacralisation does not require the use of the Kailas-abode-of-Śiva model, it may be achieved through various means, notably temple-building, but it is a Śaivite speciality. What seems the critical factor in the deployment of the Kailas model is the need to incorporate disparate population elements and cultures into a sin- gle kingdom or state. Ritual activities that develop around Kailas mountains assign a place in the ritual to every group or level of society, as we have seen for example, in the Gaddi’s prominence at Kaplas and Manimahesh. The annual pilgrimage rituals were, as anthropological studies of the region have shown, sites of social contestation, arenas of performance in which community and even individual status was negotiated.94 These rituals allowed space for the incorporation of local or immigrant deities such as Kelangnaga and the gen- eration of new understandings based on Sanskritic norms to be overlaid onto the site. The idea of a local Kailas was therefore, primarily aimed at an internal audience. It was a device by which the ruling elites of a territory could incor- porate diverse population groups under their ultimate authority, order society hierarchically, and unite its people around a cultural symbol. The efficacy of the model may have partly relied on the fact that the location of the Epic Kailas was not properly established in real terms, but the ultimate acceptance of the western Tibetan Kailas as the Epic Kailas did/does not dis- credit the local Kailas’s. Mythological constructions which drew on the domi- nant model were devised, such as placing Śiva on the local mountain for some part of the year. These myths viewed from a particular centre were sufficient

94 Sax (2002: 157–162); Sutherland (2004: 80–118). 198 chapter 8 to validate the local consensus. Notably, however, any contestation of identity seems to have been restricted to the local mountain vis-à-vis the Tibetan Kailas rather than with other western Himalayan Kailas mountains. Only the Kaplas and Manimahesh pilgrimages acknowledge each other. But any contradiction in multiple Kailas’s and hence multiple Śivas can be conceptually resolved in the ultimate reality that all creation is Śiva. Myth and consensus were produced by the regional intellectual and ruling classes. Visionary or charismatic revelation at the local level may have occurred, but its promotion at state level required the support of the royal court and its Brahmans, both temple and renunciate. Those renunciates may be broadly identified as Śaivite. While a medieval Laxmi-Narayan Vaisnavite impulse is apparent, it was ultimately unsuccessful for reasons not entirely clear. But the bestowing of the title ‘Kailas’ onto a local mountain was a Śaivite missionary schema, because it is primarily promulgated in terms of the presence of Śiva rather than as a cosmological transmission. With no clear evidence for any of these Indic mountains having an earlier resident deity, Śiva may not actually have been directly competing with a local sacrality, but taking an empty space on high from which he overlaid local sacralities such as the lake-dwelling Nagas. What is indicated by the title ‘Kailas’ in this device is not the set of myths, leg- ends, histories and rituals that are associated with Kailas-Manasarovar in Tibet, but the concept of the mountain abode of Śiva. Rivers and a great sacred lake are not necessary to this discourse. Where Śiva lives is Kailas, thus if a local mountain is called Kailas then Śiva lives there. The myths, legends, histories and rituals of that local Kailas are almost entirely distinct from the under- standings of Kailas-Manasarovar. Thus replication is not, other than inciden- tally, of Kailas-Manasarovar and its cosmos, but replication of the presence of Śiva, who thus manifests in new and ultimately in all regions and territo- ries. This conceptual device served the alliance of interests between western Himalayan ruling dynasties, Brahmanical temples and Śaivite missionaries. For the latter the land became, or in their understanding was acknowledged as, Śivabhumi. But the model of Kailas and the associated ritual and conceptual technology of the Śaivites became part of the political armoury of western Himalayan kings. It was promoted by the royal and intellectual classes for internal consumption, with the intention of unifying that community under the existing power structure through the use of such assimilating symbols of sacred geography. As the Kinnaur Kailas pilgrimage was used by the Bushahr rulers to unite Indic and Tibetanised populations, for example, so too did the Kaplash and Manimahesh pilgrimages incorporate the Gaddis, nomads kinnaur and adhi kailas 199 whose animal husbandry was of considerable significance in the economic and nutritional processes of the region. In that the Gaddis primarily grazed their herds in areas outside of set- tled agriculture they were not in competition with village societies, and the relationship between the two groups could thus be symbiotic, a relationship socially sanctified by their part in the pilgrimage ritual(s). In the modern period this impulse is also apparent in the ongoing incorporation of the ‘Bhutia’ communities of the Adhi Kailas region into the modern Indian Hindu world through the symbols of that mountain. The construction of the Kailas mountains of India should therefore be seen primarily in the context of local interests and discourses within western Hima- layan kingdoms and states, as part of religio-political processes integrated into the local socio-economic world. They are not substitutes for the ‘real’ Kailas nor do they appropriate Pan-Indian cosmological mythology. They are distinct entities in their own right. In the following chapter we will discuss these con- clusions in the context of Sri Kailas, a mountain located close to the Gaumukh source of the Ganges, and demonstrate how the modern understanding of that region was again shaped by late 19th and 20th century Śaivite renunciates, particularly of the Giri order, which process had a major impact on the under- standing of Kailas-Manasarovar. chapter 9 Sri Kailas: The Mountain at the Source

Introduction

Sri Kailas is a 23,083 feet (6932 metre) high peak located to the northeast of Gaumukh in the ridge of mountains between the sources of the Alakananda, Jadh and Bhaigirathi feeders of the Ganges. The Raktavarna, Sri Kailas, and Lambigad glaciers separate the peak from these rivers (see maps 8, 9). Major K.S. Dhami, who led an expedition that climbed Sri Kailas in 2001, describes it as “one of the dominating peaks of … [it] has a unique pyramid shape which makes it totally different from the other peaks in the region.”1 That distinctness had been noted earlier by German mountaineers whose leader, Professor Rudolph Schwarzgruber, recorded that when his party climbed peak in 1938;

[W]e were particularly impressed by what we saw to the north. There a pyramid-shaped, snow-covered mountain rose from the glacier below … To the north of this mountain the brown, desert plains of Tibet appeared.2

The German team subsequently trekked via the Raktavan Bamak (‘glacier’) on the first ascent of, “Sri Kailas, that mountain which had impressed us so much.”3 While this Kailas has attracted mountaineers it is the most remote of India’s Kailas mountains, visible only when traversing difficult terrain above the Bhai- girathi Ganga source at Gaumukh or the now-largely deserted Jadh Ganga course.4 There are no pastures to attract tribes to graze their flocks around

1 Dhami (2001: 171–173). Before ascending, Major Dhami’s party acknowledged the mountain’s sacred nature with prayers and puja. 2 Schwarzgruber (1939: 145). 3 Schwarzgruber (1939: 145); also see Schwarzgruber (1939a). Schwarzgruber gave the peak’s height as 22,742 feet; he encountered Major Gordon Osmaston who was then surveying the region. It was presumably Osmaston who established the height accepted by The Alpine Club (London). 4 The Jad(h), an ethnic group, closely linked to, if not synonymous with that known as the Bhutia in Kumaon (Lecomte-Tilouine 2009: 176–177), are associated with the course of the Jadh Ganga, which was a trading route for them with western Tibet.

© koninklijke brill nv, leiden, 2015 | doi: 10.1163/9789004306189_011 the mountain at the source 201 the mountain, there are no permanent settlements or temples closer to it than Gangotri, nor is there any evidence that there ever have been. Sri Kailas is not within local ritual or mythological worlds, nor is there any history of its use as a symbol of the region or its power structures. Unlike the other Kailas moun- tain that is not locally sacred (Adhi Kailas), this is not even a site promoted by the tourist industry. While it appears on maps, no pilgrims or even renunciates seem to visit it today. The designation ‘Kailas’ thus serves no obvious interests, for although Gangotri is now a major Pan-Indian sacred site the mountain is effectively inaccessible to popular pilgrimage. While located on a cultural frontier, this Kailas mountain clearly does not fit into the Indic Kailas models and processes discussed in the previous chapters. Yet a study of this peak, the history of its naming, and its location within a sacred region suggests it could be the model Kailas mountain in the Indic tradition.

The Kingdoms of Garwhal and Kumaon

During the first millennium ce., the Kuninda kingdom and its successor dynas- ty the Katyuris may have extended to the upper Ganges region. But these king- doms were centred around the regions of present-day Josimath and Almora and we cannot be certain that their writ extended as high into the mountains as Badrinath. Following the decline of the Katyuris in the late 11th century, regional dom- inance passed to the Khas Malla kings of the Jumla region (western Nepal), whose territory included parts of Ngari (western Tibet). In 1191 they invaded Kumaon and Garwhal and became the dominant power there during the 13th century.5 Originally Buddhist but later increasingly Hinduised,6 the Malla7 roy-

5 On a 1223 inscription in the Baleswar temple complex (Champawat; Kumaon) confirming this, see Vitali (1996: 448). 6 Tucci (1956: 103–116). 7 On which see Lecomte-Tilouine (2009: esp., 11, 254–276). She argues for the Mallas ruling “a defined religious territory” embracing a triumvirate of flame sites, with Jvalamukhi (h.p.) the third such centre. Yet there is no evidence that Malla authority extended that far west, suggesting that the construction was an ideal. But perhaps the flame site at Panwal outside Kedarnath on the route to Gangotri may have been intended? (Atkinson [1884: 19] states that Kedarnath was conquered by Aneka Malla in 1191.) Lecomte-Tilouine (2009: n. 18 257) notes an inscription from 1299 that may refer to a Nath presence in the polity and Kedarnath was, along with these other flame sites, a Nath centre. Chronologies, however, are unclear. 202 chapter 9 als, like other rulers of local kingdoms, visited Lake Manasarovar8 as part of their legitimising strategy. During the Malla period the growing power of the Islamic forces in the Gangetic plains was inducing population movements into the more secure Himalayan regions, particularly by the higher castes. Thus the Dehradun Gazet- teer records an influx of twelve clans of Brahmans into the region in the 1000– 1300 period (and 69 clans of Rajputs during the 1300–1700 period).9 These migrations presumably strengthened Sanskritisation, bringing the traditions of various Brahmanical gotras into at least the lower hill regions. The upper hills however, were subject to the cultural and political influ- ence of the dynamic western Tibetan kingdoms. Tibetan sources record that after 1265 parts of Garwhal and Kumaon (again?) came under western Tibetan rule and in the early 14th century the Purang king is stated to have conquered Gangotri (Gon go phra).10 As will be seen, Garwhal and Kumaon rulers both launched expeditions against Tibetan forces during the 17th century and while Badrinath was under Hindu kings, Gangotri apparently remained at least trib- utary to Tibetan powers down to the modern period. While briefly held by the Gurkhas in the early 19th century, there seems no evidence of Indic rule there in the medieval period, and on the Jadh Ganga course the earliest European explorers as will be seen, found Tibetan authority extending at least to Nilang on the Jadh Ganga and even to below the Jadh-Bhaigirathi confluence that con- trols access to Gangotri. The post-13th century decline of the Mallas saw the emergence of the Ku- maon and Garwhal kingdoms.11 In Kumaon the Chand dynasty initially ruled from their capital at Champhawat (south of Pithoragarh).12 But presumably seeking a centre more secure against further attacks from the east the Chands gradually extended their power westwards, establishing Almora as their cap- ital in 1562. By the late 17th century the Chands, following traditional models of kingship, had expanded their writ to the north. King Baz Bahadur Chand reportedly invaded western Tibet as far as Taklakot in 1670 after reports of attacks on Indian pilgrims there, and in 1673 he issued a land grant for the sup- port of travellers to Manasarovar.13

8 Lecomte-Tilouine (2009: 206, 212–213). 9 Walton (1910: 220–221). 10 Vitali (2003: 79). 11 Lecomte-Tilouine (2009: 181) dates the advent of Kumaon rule to “around 1374.”; Joshi (2009: 327) however, states that Kumaon was under Malla rule only until 1223. 12 On the dating of the emergence of the Chands, see Goetz (1969: 175, 180); Vitali (1996: n. 750 448.); also see Joshi (1988: 73–86, esp., 80); Joshi (2009: esp., 327–328). 13 Walton (1911: 12–13) credits him with having conquered Taklakot and (1911: 178) states that the mountain at the source 203

West of Chand territory in what is now Garwhal, a separate kingdom emerged with its capital at Srinagar, a town on the Alakananda river between Devaprayag and Rudraprayag (and not to be confused with the Kashmir capi- tal of that name). The kingdom reportedly originated in the late 14th century, when Ajay Pal, a scion of the Parmar principality, united 52 chiefdoms under his rule.14 His Parmar dynasty survived until the death of the last king, Pradyuma Shah, in battle against the Gurkhas in 1804. There were many similarities between the neighbouring Kumaon and Gar- whal medieval kingdoms. In both cases the ruling families claimed Rajput ori- gin and emerged from among the rulers of petty principalities that were at the heart of the kingdoms they developed, suggesting roots in the Rajput influx. Centred in the lower Himalayan foothills, both the Chand and Parmar king- doms enjoyed good access to the plains, meaning that control over north-south trade would have offered income to supplement the agricultural wealth of the fertile river valleys at the heart of their kingdoms.15 Until vanquished by the Gurkhas in the early 19th century, the two kingdoms existed side-by-side but their actual spheres’ of control varied according to their respective strength at any one time, with frequent, almost ritual, warfare between the two. Both Kumaon and Garwhal controlled routes to Kailas-Manasarovar. Ku- maon controlled the Niti la (Tib: la = pass) and passes to the east via the Kali river, while Garwhal controlled the Alakananda route to Badrinath and the Mana la and Thag la. Even earlier than the Kumaon campaign in 1670, Garwhal had intervened in the upper hills. The Portugese Jesuit missionary Antonio de Andrade (1580–1634), who travelled from Badrinath to Tsaprang over the Mana- la in 1624–1625, noted that a large Garwhali force under king Manipat Shah,

around 1670 the King set aside the revenue of five villages near the passes for providing pilgrims with food, clothing and lodging. In the wider sense this was part of the Chand Sanskritisation policy, also seen by Pande (1993: 252, 259, 265) who notes that in 1692 the Chand king invited an Atharvavedic Brahman and gave him land near Almora; also see Brown (1984: 9) who credits Baz Chand with bringing the Bhutia peoples into the Hindu world. But given that Indic texts clearly stated the Kailas region was only open to renunciates, Baz Chand’s patronage raises the question of whether those patronised were renunciates or householder pilgrims? (This is not stated.) It may indicate the beginning of popular pilgrimage to Kailas on this route. 14 For a list of the 52 Chiefdoms, which we note does not include the Gangotri region, see Rawat (2002: 52–53). A proper critical history of the region remains to be written. But see Rawat (2002); also see Saklani (1987). 15 Raper (1812: 459) notes the existence of a slave trade, with “hundreds” of slaves from “the hills” (whose identity is not recorded), traded annually. This aspect of the region’s economy remains unstudied. 204 chapter 9 taking advantage of a rebellion by “three rajahs, tributaries of the Tibetan king”, had invaded Tibet at three points in 1624. They were forced to retreat, however, when a snowstorm broke their lines of communication.16 The event is not men- tioned in Tibetan sources however,17 although (as will be seen) this was a period of turmoil in western Tibet. The modern era in the upper Ganges region may be dated to the British vic- tory over the Gurkhas in 1815, when both kingdoms passed to British authority. They were then divided into (a) British Garwhal to the east of the Alakananda (and including Badrinath); and (b) the Princely state of Tehri-Garwhal compris- ing the region to the west of the Alakananda excepting Dehradun (and with its capital at Narendranaga, 8 miles [12 kms] above Rishikesh). Tehri-Garwhal was gifted to the son of Pradyuma Shah, while British Gar- whal was initially placed under the District Commissioner George Traill.18 The status of Rawain, the district including the Jadh and Bhaigirathi sources of the Ganges and the Jumuna headwaters, was subject to some discussion. The exis- tence of pilgrims there was reported in the context of unsettled conditions in which they were frequently robbed, but their destination is not stated. The district was eventually given to Tehri-Garwhal, indicating that the area and its pilgrimage were seen as of little strategic or economic importance by the British.19 Commissioner George Traill held his post for twenty years,20 and was the key figure shaping the regional encounter with the British colonial state. Traill’s primary duty—as with other such officers—was to turn his district into a revenue-producing one. Traill thus carried out a land revenue settlement as

16 Wessels (1924: 67). Apparently citing local chronicles, Chandola (1987: 27–31) describes several cross-border expeditions by Pal dynasty forces during the 17th century. These were made in response to attacks on the Ganges sources regions by Ngari raiders. Pal forces are stated to have conquered and extracted annual tribute from Daba (the secular regional centre near Tholing) and Chandola implies that it continued into the 18th century. 17 Information courtesy of Roberto Vitali. 18 A Bengal Civil Service officer, the Orcadian George William Traill (1792–1847) was initially under the Hon J. Gardner, who was appointed Commissioner. But Gardner remained at the Residency in Kathmandhu and Traill soon inherited full responsibility. He served as District Commissioner Kumaon until 1835. 19 Walton (1910: 72–73) notes the economic importance of the Kedarnath and Badrinath pilgrimages which, “must on the most modest computation be worth not less than five lakhs a year to the inhabitants.” On Rawain, see Walton (1910: 130); also see Rawat (2002: 132). 20 Around 1819, he was assisted by the later pioneer of Himalayan studies, Brian Henry Hodgson; Waterhouse (2004: 4). the mountain at the source 205 a basis for taxation, began forestry development, and so on. But he was also aware of the economic benefits of pilgrimage, in that pilgrims brought revenue into a district and created employment opportunities. He recorded that in 1820, Badrinath had attracted 27,000 pilgrims, with many more turning back because of a cholera outbreak. He also confirms that Kedarnath was a significant site, attracting around a third of those who had visited Badrinath. But Traill, the first European to visit Kedarnath, found it almost inaccessible, and from the late 1820s the roads from Haridwar to Kedarnath and Badrinath were subject to continual improvement on his orders, despite accusations that in facilitating the pilgrimage he was encouraging “heathen idolatory.”21 At the beginning of the British period Badrinath was therefore, a significant pilgrimage site, and Kedarnath also of some importance. Gangotri, however, was rarely mentioned. But as we shall see, there were contradictory and distinct bodies of knowledge concerning these sites, which were only beginning to coalesce. British surveyors had finally resolved the question of the Ganges’ source(s), and it was no longer possible to sustain the belief that Manasarovar and the Ganges were somehow connected (unless by the rather far-fetched idea of underground tunnels). Scientific modernity had demolished that Sanskritic fusion of ideal and real landscapes. Badrinath would seem the obvious ben- eficiary of this, but an alternate sacred geography emerged with the rise to prominence of Gangotri. The source of the Ganges was a place of great significance. The Indus, Sutlej, Brahmaputra, and Karnali rivers were of little or no spiritual importance to the heartlands of Indic culture on the Gangetic plains. It was the Ganges river that was central to their world-view, and sacred Himalayan geography had to embrace that river’s sources, while if the heavenly model was to be replicated on earth a sacred lake and/or mountain had to exist where that river originated. In this and the following chapter, therefore, we discuss alternative sacred geographies which will be seen to be of considerable relevance to the history of Kailas-Manasarovar. Leaving aside the issue of Kedarnath (which seems to have been developed as a medieval Śaivite alchemical centre22), we

21 Rawat (2002: 73, 239); Tolia (1994: 90–91). 22 Kedarnath, the source of the Mandakini river that joins the Ganges at Rudraprayag, also seems to be within the realm of the local deity Bhairon but is a major Śaivite centre (with Bhairon understood by Śaivites as Bhairava). Among its water sources is a red-coloured stream with alchemical associations, and its development seems associated with medieval Naths. Lorenzen (1972: 102–105) notes inscriptions record the Kālāmukha’s association with Kedarnath from the 1100s, so the site was known from at least that time. 206 chapter 9 focus on Badrinath and particularly Gangotri, the development of which seems closely linked to the modern construction of Kailas.

Badrinath

When we look into the history of the region around the Ganges sources, Badri- (nath) is by a considerable margin the most prominent toponym. Gangotri by contrast, largely escapes mention until the modern period.23 ‘Badri’ is referred to in the Mahabharata, which reference is commonly assumed to refer to present-day Badrinath. Yet in a sea of shifting toponyms this identification remains conditional. One problem with the site’s history is that according to traditional Indic accounts, Badrinath temple was established by Sankaracharya in the 8th century. But modern studies demonstrating the largely mythical sta- tus of this traditionally fundamental figure24 mean we must consider alterna- tive histories. While we cannot be entirely certain that the early toponym Badri equates to modern Badrinath, the Epic references describe Alaka, the palace of Kubera (hence the Alakananda river toponym), as located near Badri and there re- mains an image of Kubera in Badrinath temple.25 But the site was contested and at times under Tibetan influence or control. There is also a local world, with the deity Mani Badri still acknowledged at Badrinath,26 and presumably the source of the toponym,27 with the suffix nath indicating the influence of that renunciate sect from around the 13th century.We have seen that in Kinnaur the Kamru deity Devta Badrinath, the clan deity of the Bushahr Raja, is consid- ered a migrant from Guge, suggesting that family were among (or claimed to be) followers of Mani Badri who moved to Kinnaur at a time after the addition of the nath suffix. That Badrinath was a contested site is also indicated by the complex rela- tionship of its temple to western Tibet’s Tholing (Buddhist) monastery, not least in a legend of the central image at Badrinath temple moving between

23 Malhotra (1983: 147) states that the first published accounts (in Hindi) of Gangotri pilgrim- age date to the early 1960s. 24 Clarke (2006: 148–176). 25 Tapovan (1984: 10). 26 Brown (1984: 13). 27 Both local guides and Wikipedia state that Badri may, however, also denote ‘wild berries’, ‘plums’. (Pahari?) the mountain at the source 207 the two places.28 B.R. Chatterji, one of a party of Amritsar academics who trav- elled to Tholing in 1922 with a renunciate guide of the Naga sect, reports that the renunciates considered Tholing to be the original Badrinath.29 Given that Tholing monastery was founded in 996 there are several possible histories here. Badrinath may have came under Tibetan influence in the ensuing expansionist phase of the Guge-Purang kingdoms, if not earlier, while it is also possible that Tholing was built over a centre for Mani Bhadri worship as part of Buddhist sacred landscape transformation. Badrinath may only have been a Sanskritic centre since the late 14th century, the dating given by Ananda Bharati for the founding of the Badrinath temple, which he attributes to Pamar dynasty founder Ajay Pal’s desire to strengthen his northern frontier.30 Pal’s dynasty characteristically claim highly improbable origins in the distant past, but their suggested descent from immigrants in the Mount Abu region of Rajastan is tenable. Toponym transference is characteris- tic of the Sanskritisation process and it is notable that Garwahl toponyms such as Mandakini, Koteshwar, and Gaumukh (each respectively designating bod- ies of water, Śaivite temples and sacred sites), also occur around Mount Abu, where a Parmar dynasty ruled in the 9th century. Abu itself has sacred associ- ations, and elements of a transference of the local sacred geography from that site to Pan-Indic prominence in the Garwhal Himalayas are perfectly possible in the wake of upper-caste movement into the hills consequent on the Islamic conquests.31 In that case, it again suggests that it was under the early Parmar dynasty that the Alakananda river route to Badrinath was finally established as Indic territory, and Sanskritisation applied to what was previously a local or contested cultural zone. Ajay Pal’s northward expansion is associated with Nath initiatives, for he was reportedly aided by a Siddha Satyanath,32 and his Srinagar palace was con- structed at the geomantically symbolic site of a Nath shrine. Local historians

28 Walton (1911: 177) provides a parallel. Prior to Baz Chand’s attacking Tibet in 1670 he had invaded a Garwahl border fort and taken its image of the goddess Nanda back to a temple in Almora. 29 Chatterji (1940: 30–34). 30 Bharati (1963: 164); also see Nand and Kumar (1989: 127) which states that the Dalai Lama sent offerings to Badrinath temple of tea, yak’s tails, etc, and they in return sent musk, sweets, cloth., etc as prasad. [No date is given and the endnotes referenced in this work are actually missing from the publication.] 31 Rawat (2002: 22–31). As we have seen, however, the toponym Mandakini is used much earlier in reference to the Himalayas. 32 Nand and Kumar (1989: 125); see n. 30. 208 chapter 9 assign him an important role in this sect (which has remained strong in this area33), and some sources even eulogise him as one of the 84 Mahasiddhas.34 Later Nath patronage by his dynasty is confirmed by copper plate inscriptions for the period 1669–1715, and perhaps by the local system of sadāvrata, the endowment of land revenue from assigned villages to provide food for pil- grims to Badrinath and Kedarnath.35 If Naths were established there around the 13th century, they seem likely candidates to have penetrated into the upper Himalayan regions, as they did in the Chamba and Kangra kingdoms. The toponyms of Kedarnāth and Badrināth certainly imply the influence of this Saivite sect.36

33 See for example, Briggs (2001: 78–81). 34 See for example, Rawat (2002: 34) who attributes to George Briggs study of Goraknath the statement that “Raja Ajay Panth was the founder of one of the ten sects of Gorakh Panths.” Briggs (2001: 74) actually lists the name Ajāipāl, without comment, as one of the sects that developed from Goraknath’s order. Rawat cites the Sanwari Granth in support of his place among the Mahasiddhas, but the name is not found in lists from other regions; see, for example, the list given in Sharma (2001: 74–80). 35 Rawat (2002: 55, 110). 36 The Tibetan polymath Taranatha met a number of Indian renunciates who appear to have been Naths; i.e.; “the ācaryā Tīrthanātha who had come … with his entourage of many yogins.” But this is not to suggest these Naths all visited the Kailas region; Taranatha’s biography of Buddhaguptanatha (c. 1514–1610) indicates that he travelled in Iran, Java, Burma, Sri Lanka, and Tibet but does not mention his visiting Kailas; see Templeman (2008: 131, 250); also see Huber (2008: 205–207). For an account of Indo-Tibetan sacred geography developments and renunciate initia- tives in the 18–19th century, see Huber (2008: 193–231). On the many renunciates visiting Lhasa in the 17th century under the patronage of the 5th Dalai Lama, see Schaeffer (2011). On the 18th century, see Petech (1988). As Huber observes, the relationship between Chait Singh, raja of Benares 1770–1781 and the Panchen Lama is particularly significant here. That renunciates from Benares travelled to Kailas-Manasarovar in this later period is suggested by the existence there of the Sumeru pitha math. The text (Maṭhāmnāya) detailing this institute’s lineage states the lineage tirtha is Manasarovar and its kṣetra is Kailasa, although whether this was an ideal or an actuality is difficult to ascertain. The math was patronised by the Maharaja of Benares and from 1758 the Mahants were all Bengali. While active in the 18th and 19th centuries, the math was later moribund; Clarke (2006: 144–147). Concerning what may be an earlier tradition, actual or ideal, see Clarke (2006: 119–121). This concerns the northern Dasnami math at Josimath. The Maṭhāmnāya describes its lineage tirthas as Alakandandā and Badarīkāśrama, its deities as Badarīkā (Nārāyaṇa) and (female) Pūrṇagirī; and its Veda as the Atharvaveda (although its stated gotra, Bhṛgu, is Rg Vedic). This is consistent with an earlier sacred geography centred south of the Himalayan watershed, with later construction of a Kailas-Manasarovar centre. the mountain at the source 209

But Ajay Pal installed his clan deity Raj Rajeshwari in his palace and his name appears in local Tantric manuscripts,37 so like other kings he doubt- less supported a range of religious initiatives. Here the establishment of the Badrinath Laxmi-Narayan shrine38 seems consistent with contemporary west- ern Himalayan Sanskritisation initiatives, as we saw with the similarly Nath- influenced Chamba kingdom, where such a shrine was apparently an impor- tant part of the regional Sanskritisation process. While renunciates probably reached Badrinath at an earlier date, founding the temple created the condi- tions necessary to open the site to ordinary pilgrims. Andrade confirms that Badrinath was a Pan-Indian pilgrimage site by 1624 and also informs us of something of the contemporary nature of Himalayan pilgrimage. In Delhi the Jesuit joined a pilgrim party heading for Badrinath, indicating that pilgrims banded together for such journeys.39 But accounts of Gangotri seem absent from contemporary Indic sources and there is no evidence that it was visited at that time. It seems that Gangotri remained under Tibetan overlordship or perhaps remained the preserve of local communities.

On Gangotri

Like Badrinath, Gangotri is now regarded as an ancient centre of the Hindu faith, but while its antiquity is part of its modern status, there is surprisingly little evidence to support it. It is difficult to locate in the classical sources. The principal myth of Gangotri, the penance of king Bhaigiratha to obtain the release of the Ganges waters withheld by Śiva, is said to have occurred at Gorkarṇa, which is thus equated to Gangotri. But while the legendary king has given his name to that branch of the river that derives from Gaumukh, the identifications are problematic. Not only are other sites, notably in south India, identified as Gorkarna and that toponym given to two sites in the earliest recension of the Skandapurana (8th century ce.), but it also occurs in that text

37 Rawat (2002: 34). 38 I have not located a proper critical study of Badrinath or even its art and architecture; see however Kumar (1990); Watson (1961). Badrinath is primarily considered a Vaisnavite site today, and the central deities at the temple complex include Laxmi, Narayan, and the local deity Badri, Sanskritised as a form of Visnu but said by tourist literature to be considered by Buddhists to be a representation of the Buddha. Kubera and other deities also find a place there. As a frontier site, we might expect such variety, representing encounter, contestation and other religious processes. 39 Wessels (1924: 46–48). 210 chapter 9 as the name of a mountain and of a deity dwelling on Mount Mahendra.40 Gorkarna is thus another free-floating, rather than fixed, toponym. Neither is Bhaigiratha’s penance only attributed to one site. The Matsyapu- rana, for example, locates it as occurring at Lake Bindusara (as noted, said to be north of Kailas).41 The Manasakhanda states that it occurred at Manasarovar,42 and today it is associated in various tourist literature with a rock overlooking the waterfall in Gangotri, or with Tapovan (above Gaumukh). Mythological events, names, and places can be applied to many landscapes and what seems the earliest version of this common myth is actually a Vedic account of the descent of the heavenly river Sarasvati in the Kurukṣetra area, which the Kuru tribe conceived as the centre of the world.43 This was later applied to wherever the author(s) wished to locate the Ganges source and/or centre of the world. In any case, Gangotri as a historical site cannot be conclusively identified earlier than in the 14th century Tibetan chronicle mentioned above. Notably, neither the comparatively recent Manasakhanda or the Kedarakhanda (or at least their English-language precis), refer to Gangotri, while a temple there is only known from around 1806 when it was constructed on the orders of the Gurkha General Amar Singh Thapa, who had conquered the region two or three years earlier.44 According to a local tourist guidebook, the General appointed to the Gan- gotri temple, pujaris from Gangotri and Mukhwa (at which place, near Harsil, the image of the river goddess Gangama is taken every year to spend the win- ter). He also allocated land for their support. But the reference to Gangotri pujaris is problematic. The same source states that prior to 1806, “the Rajputs of Taknor” (Harsil district) acted as pujaris.45 This suggests the earlier existence of

40 Bisschop (2006: 202, 242); a popular work, Chaturvedi (n.d.: 55), identifies it as “a holy spot near Goa”. On Gokarṇa as dwelling on Mt. Mahendra, see Sanderson (2003–2004: 417). 41 Bindusara is clearly differentiated from Manasarovar in, for example, the Matsyapurana 121.23–41 (Agrawala: 1963; 202–203). 42 Atkinson (1884: 43). 43 Witzel (1997: 15–16). 44 Atkinson (1884: 89) records royal grants from the Chand dynasty to Badrinath and Kedar- nath temples in 1744 and 1745 respectively, but does not mention Gangotri. Oakley (1905: 152) notes the destruction caused by a great earthquake in 1803. The existing temple is not, in any case, the original Nepali construction, but one built by the Jaipur Raja, Sawai Madho Singhji, “about seventy years ago”; Sundaranand (2001: 66). The Gurkhas also built a tem- ple at Kedarnath; Barron (1990: 51). Malhotra (1983: 134) states that prior to the wooden temple (presumably that erected by the Gurkhas), there was a cave temple there. 45 Agarwala (2003: 112); the source is, however, eclectic at best. But given that the region above Harsil was under Tibetan suzerainty this may have been the nearest Brahmanical the mountain at the source 211 a local shrine, with local officiants (of the kind described in Chamba, etc.,) dedi- cated to local deities such as the river deity Gangama, to which General Thapa’s temple with Brahman pujaris was the first Sanskritic overlay.46 Gangotri and its surrounds were thus, until General Thapa’s forces arrived, seemingly outside the Sanskritic world. The construction of the temple at Gangotri did not immediately lead to it being a pilgrimage place. Indeed, as we shall see there was local sentiment that the temple was inappropriate, and its construction was imposed by an invader rather than initiated by a traditional leader, which must have limited its immediate regional appeal. Significantly the appeal it developed was to be Pan-Indian, in contrast to most major pilgrimage sites in Indic culture that are firstly of local, then regional significance and only later linked to Pan-Indian schema. During the 20th century Gangotri became fully Sanskritised as—along with Badrinath, Kedarnath, and Yamunotri (at the source of the Yamuna)—one of the ‘char dam’ (‘Four Abodes’ [of the Deity]), a modern sacred geography schema that echoes the earlier all-India char dam comprising Badrinath, Puri, Dwarka and Rameswaram. The reports of early 19th century European travellers suggest Gangotri’s pre- modern insignificance. They described the sanctity of Badrinath and Kedar- nath but largely ignored Gangotri (although the surveyor J.A. Hodgson, who sought informants among the Brahmans there, recorded the small stone tem- ple erected by the Gurkhas was dedicated to the deities Gangama and Bhai- girathi.47) This neglect reflects the pilgrimage patterns that they observed for they report that pilgrims coming up from Haridwar made a circuit embracing Devaprayag, Rudraprayag, Kedarnath and Badrinath, returning to Haridwar via

community. Channa (2013: 145) accepts that the Gangotri/Mukhwa (Mukhpa) priests are Brahmans long-settled in the area but we may suspect some status enhancement has occurred; further investigation is needed. We may note that the officiants at Badrinath and Kedarnath are South Indian Brahmans, but that Gangotri was absent from that formulation of authority. 46 While there is little manifestation of local beliefs at Gangotri today, they survive at nearby Harsil and other local villages; i.e. the Naga temple outside Mukhwa. 47 Hodgson (1822: 97–98). It was thus the erection of the temple that aroused resistance and which represented Sanskritisation. Moving local deities from an outdoor shrine to a temple seems to indicate the move from local deity as protector of the frontier of a particular territory to a geographically central figure of worship and is clearly a preliminary step in their incorporation into the Sanskritic pantheon. In this case the status of the Ganges meant that its presiding deity Gangama was already a Sanskritic deity, but this may have been the introduction of the Pan-Indic Gangama to the local Gangama! 212 chapter 9

Karnaprayag.48 A sacred Gangotri is not mentioned although this could reflect the interests of the colonial sources, with British economic interests being in Badrinath but not in the Princely state centre of Gangotri. Yet the absence of its mention even in regard to these pilgrimage routes is notable, as is the appar- ent absence of renunciates there. European travellers tended to record the presence of renunciates, not least because as colourful figures on the landscape they were simply noteworthy.49 But for early 19th century officials renunciates were significant in other ways which ensured they were mentioned if relevant. They were important sources of information on routes and passes, and government remained interested in them because the East India Company’s battles against the sannyasis in the late 18th century were still fresh in the memory. Thus the first British travellers to reach Gangotri in the early 19th century noted the presence at Rampur (u.p.) of many renunciates and recorded that the majority of the 45–50,000 pilgrims annually visiting Badrinath were “fakirs”.50But they did not report the presence of any renunciates at Gangotri. By 1828, however, when Thomas Skinner travelled to Gangotri, he reported the presence of pilgrims from the plains, including many poor Brahmans who took home water from the Ganges. There were also “ash-covered” renunci- ates and he recorded that the “natives esteem the faquirs highly; and many are learned and perfectly sincere.”51 While the evidence is limited, it suggests that pilgrimage to Gangotri only began after the expulsion of the Nepali (and Tibetans?) from the area. The more secure conditions following the British victory over Nepal in 1815 allowed for the development of pilgrimage to the Bhaigirathi Ganga source, while the temple erected by the Nepalis provided a ritual focus for ordinary pilgrims visiting there. Yet according to the doyen of local renunciates, Swami Sundaranand, Gan- gotri did not become a major pilgrimage site or an important site for renunci- ates until the latter part of the 20th century. He states that 50 years ago only around 5000–6000 pilgrims a year visited Gangotri and that “only four or five saddhus remained in winter. They lived alone in their caves and occasionally went to meet one another.”52 When he first visited Gangotri in 1948 there were

48 Raper (1812: 486, 540). 49 See for example, Lohner (1949: 20–22) on a silent, “extremely courteous” yogi at Gangotri who lived naked even during the winter. 50 Fraser (1820: 258, 372). 51 Skinner (1832: vol. 2; 29, 55–56, quotation from 64). 52 Sundaranand (2001: 64). Clarke (2006, n. 103 129) notes references to a Gangotri math in Edwin Arnold’s Light of Asia (1894: 331) and J.M. Ghosh’s Sannyasi and Fakir Raiders in Ben- the mountain at the source 213 no renunciates residing above Gangotri other than Swami Tapovan (to whom we shall return), and it “was only after 1962 [when motorable roads to Gangotri and Badrinath were completed] that one or two travellers could be seen going above [Gaumukh].”53 So the number of pilgrims or renunciates travelling to Gangotri in the 19th and early 20th century was much less than the number that visited Badrinath and Kedarnath. Gangotri was not neglected simply due to its inaccessibility; the journey to Kedarnath was famously difficult but that site had attracted renunciates from at least the medieval period. What seems to have deterred visitors to Gangotri was its historical lack of a Sanskritic identity or location within Hindu sacred geography or Tantric worlds. While the evidence is lim- ited, it does appear this was due in part to Gangotri having been under some form of Tibetan rather than Indic authority for most of the period since the 14th century. Given Gangotri’s former obscurity, the ‘ageless sanctity’ of the place where the Bhaigirathi emerges from the Gaumukh (‘Cows Mouth’) glacier is also hard to sustain. We might quote a mahatmya describing the Bhaigirathi source:-

Sri Gomukha [sic]—the place where the sacred Ganga actually emerges into the world. Of all the places on earth, it is the holiest. It is celebrated in innumerable songs by the great rishis … It stands close to heaven and is occupied by the gods. In that beautiful place, from a huge formation of snow shaped like the mouth of a cow, the Ganga issues forth to purify the world and wash away the sins of the dwellers of the earth.54

But these verses are from the Sri Gangotri Kshetra Mahatmyam, which was only written in the 1930s, and there seems to be no earlier version. Similarly the myth of the Ganges’ release from Śiva’s matted locks makes no mention of a cow’s mouth, and is part of an entirely different cycle of myths, one centring on the relationship between king, earth and celestial worlds.55

gal (1930: 12). He concludes this is a now derelict structure at Ukhimatha (“near Gangotri”) probably granted to the Kedarnath order of Sankaracharya “presumably” for military ser- vice to the Gangotri kings [?]. Given that neither Sundaranand nor Swami Tapovan refer to any such math, and that there is no mention of it in accounts by early British explorers, I am inclined to conclude it was a 19th century initiative of no lasting consequence. 53 Sundaranand (2001: 105, 107). 54 Tapovan (2003: 264), quoting from his Sri Gangothri Kshetra Mahatmyam. On more recent local mahatmyas of Gangotri, see Pinkney (2013). 55 On which, see the discussion in Sutherland (1992: 46–48). 214 chapter 9

We have noted that the toponym Gaumukh could derive from the sacred place of that name in Rajastan,56 and the actual location of the source of the Ganges and its identification as the ‘Cow’s Mouth’ was certainly historically fluid. The Persian author of a life of Timur records his (15th century) visit to “Durreh Cowela”, which was “evidently” Haridwar. In his account there was a rock in the river there in the form of a “cow’s mouth.”57 A century later the rock had become a mountain. Sven Hedin states that;

[T]he expedition which the great Emperor Akbar sent out at the end of the sixteenth century to search for the source of the Ganges … saw the water of the river gush out in great abundance in a ravine under a mountain which resembled a cow’s head.58

That this gorge was not in the modern Gaumukh region is confirmed by Jesuit Joseph Tieffenthaler’s reports, which drew on the knowledge of Akbar’s geog- raphers. The Jesuit asserted that the Ganges source would never be discov- ered because the way beyond the gorge of the “cow’s mouth” was “impass- able.”59 In 1616, the traveller Edward Terry stated that Haridwar was the Ganges source and that the river there issued from a rock “which the superstitious Gentiles [i.e., Hindus] image to be like a Cowes head [sic].”60 At the beginning of the 19th century, the Gaumukh was again a stone, according to a “native” surveyor in British employ. He stated that two miles from Gangotri;

is a large stone, situated in the middle of the bed; the water passes on each side, but a small piece of fragment is disclosed above the surface, to which fancy may attach the idea of the object [i.e., a cow’s mouth].61

The first British official to reach Gaumukh was J.A. Hodgson and his account from 1817 similarly problematises the identification. He travelled to the river source with a Brahman from Gangotri and further evidence of the contem- porary disjunction between local and Pan-Indic traditions concerning the site

56 On this and other water sources known as Gaumukh, see Eck (2012: 135–136). 57 Fraser (1820: 466–473, 478–480). He presumably refers to Sharaf ud-Din’s Persian text Zafarnāma, translated into English by J. Darby in 1733. 58 Hedin (1909–1913, vol. 1: 210). 59 Hedin (1909–1913, vol. 1: 282). 60 Hedin (1917–1922, vol. 1: 146). 61 Raper (1812: 506). the mountain at the source 215 comes from his report that; “[t]he Brahmins say, they have never heard of any rock or place called the cows-mouth … or anything like it either in sound or signification.”62 By 1832, however, the British traveller Thomas Skinner located the ‘cow’s mouth’ at Gangotri, noting that it had been located at Bhaironghati (the Jadh-Bhaigirathi confluence 6 miles below Gangotri).63 This suggests that the application of this myth to the Bhaigirathi source is no older than the second quarter of the 19th century. Also notable is that the earlier use of the myth was not in connection with Manasarovar or Badrinath and their claims to being the Ganges’ source, but with that river on the plains of India. We are dealing with different cycles of mythology whose applications may or may not overlap. That disparate bodies of uncollated knowledge of the region existed in pre- modern India can also be seen by the fact that the Persian historian Mhd. Karim Ferishta (c. 1560–c. 1620) recorded that the Jamuna and Ganges sources were within the lands of the raja of Kumaon (under which he includes Garwhal).64 But this was still unknown to the British when Major James Rennell published his famous map of 1782,65 and neither was it known to Purangir, a renunciate of the Giri order, who some time prior to 1773 travelled to Manasarovar believing that lake to be the Ganges source.66 Purangir was an extremely sophisticated and learned individual. His guru had visited Tibet before him and Purangir served as an intermediary between the East India Company and Tibet, accom- panied eic emissary George Bogle to Shigatse in the 1770s and also visited Peking in the entourage of the Panchen Lama.67 Yet despite having access to bodies of knowledge in the keeping of the British, Chinese, and Tibetans as well as his own order, he remained unaware of the actual geography of the Ganges’s source region.68

62 Hodgson (1822: 104, 117). 63 Skinner (1832: vol. 2, 42). 64 Rawat (2002: 36, citing Mhd. Karim Ferishta, History of the Rise of Mohomedan Power in India till the year 1612, [J. Briggs, trans.], vol. iv, 549–550). 65 Hedin (1909–1913: vol. 1: 214–215). 66 Hedin (1909–1913: vol. 1: 18–19). 67 On Purangir[i] see Lamb (2002: 383); Teltscher (2006); also see Petech (1988: 49–62). The use of renunciates as diplomatic agents was common throughout the region; see for example Aris (1988: 204) concerning the Jebtsundamba Khutuktu of Mongolia’s use of an Indian renunciate as a messenger in 1723. 68 The account recorded in Duncan (1801: 45) concerning the travels of the renunciate Praun Puri is similarly removed from scientific geography. He describes his journey to 216 chapter 9

It was only in the first decades of the 19th century, as part of British scientific and strategic initiatives, that the Ganges source(s) were finally established in their geographical reality after several decades of speculation and preliminary enquiry. The Survey of India began its investigation into the issue in 1807, when Lt. William Webb, with Captain Felix Raper and the Anglo-Indian, Captain Hyder Jung Hearsey, were allowed by the Gurkha Governor of Srinagar to proceed north from Haridwar. After a difficult journey up the Bhaigirathi they turned back before the Jadh confluence,69 but sent Jung Hearsey’s assistant up to report on Gangotri. Webb and his party then explored the Alakandanda route to Badrinath, establishing that it carried a greater volume of water than the Bhaigirathi and was thus from the scientific perspective the main source of the Ganges. Webb travelled on past the village of Mana (north of Badrinath), and then west to the traditional source of the Alakananda, the 550 feet high Vasudhara falls which derive from Satopanth Tal (Hindi: tāl = lake/pond) and ultimately from Satopanth bamak (glacier). These falls Webb accepted as the true Ganges source,70 although another branch of the Alakananda actually continues north to a source below the Mana-la (only a few miles—albeit separated by a mountain range—from the easternmost source of the Jadh Ganga: see maps 8, 9). That northern section of the Alakananda is locally known as the Sarasvati, the toponym of a major river in the classical sources that has now dried-up. In 1815, with the Gurkhas vanquished, a well-connected private traveller, James Fraser, became the first European to reach Gangotri and two years later the surveyors Captains John Hodgson and James Herbert reached Gaumukh.71 As Survey of India officers, Hodgson and Herbert naturally collected informa- tion of strategic and political importance and among the information they

Manasarovar, and states that the Muslim area of “Teree Ládac” is two days to the west of this site, and that seven days to the south is “Cailafa Cungri”. The source of the Alakananda lies four days from there (in an unspecified direction), and five or six days south of there are Kedaranth and Badrinath. This account can certainly be read as implying that “Cailafa Cungri” is Sri Kailas, but it may also imply an entirely distinct or simply vaguely recalled understanding of the landscape. 69 Colebrooke (1812). Webb’s map of the Ganges reproduced with his article shows nothing above Gangotri, and makes no mention of the Jadh Ganga. 70 For a critical popular account of Webb’s mission and subsequent controversies, see Allen (1982: 57–77). 71 Oriental and India Office Library, l/p&s/12/4172, “Documents provided by the Tehri state on 28 February 1935”. On routes through these regions, see Sanan and Swadi (1998: 271). the mountain at the source 217 obtained was that the village of Nilang on the Jadh Ganga was the regional Tibetan frontier town. Its actual status reflected the overlapping sovereignties of the frontier. Quoting the officiating Brahman at Gangotri, who had visited Nilang, Hodgson stated that;

The people considered they belonged to Tibet and paid tax to a collector from Chaprang [Tsaparang], but also gave the Raja of Bashir a blanket per man every three years and some raisins [!] to the Garwhal Raja.72

The status of Nilang later became contested, for the modern nation-state sys- tem of fixed borders did not allow for such local incongruities and in their search for a secure northern frontier for India, the British sought a border along the mountain watershed. At the Simla Convention in 1913–1914, when the Indo- Tibetan border was agreed between the two powers, not only Nilang, but all of the sources of the Jadh Ganga above Nilang were included within India. While signing the Convention, the Tibetans objected to their loss of territory and a long-running dispute developed as to whether Nilang was within Tehri- Garwhal or Tibet, which was further complicated when Bushahr claimed it as their territory! Rather embarrassingly for the British, the Tibetans were able to cite British maps in their favour, including one attached to a book written by Sir Charles Bell, British India’s foremost Tibetan expert who had played a major role in defining the border at the Simla Convention.73 The Tibetan claim did not stop at Nilang. In their view, the border was at Gum-gum nala, a ford below the Jadh Ganga/Bhaigirathi confluence just above the hamlet of Harsil.74 They claimed it extended eastwards from there to the Mana-la above Badrinath, thus embracing Gangotri and Gaumukh. (But they were less certain about these sites, and after consideration of their place as Hindu centres agreed they could remain within Tehri-Garwhal.) The Tibetan claims, at least to Nilang, were difficult to refute. Earlier travellers, including East India Company envoy William Moorcroft, had considered Nilang Tibetan and Bushahr, whose intervention was regarded by the British as pro-Tibetan,

72 Hodgson (1822: 91). 73 The history of the Nilang dispute is detailed in Oriental and India Office Collection, l/p&s/12/4172; (various documents), which is the source for what follows. A concise summary of the issues is provided by Lamb (1989: 365–377, also see map 5, 561). Also see Ottley (1940: 27–29). On Bell, see McKay (1997). 74 Skinner (1833: 79–80) records a ruined temple near Mukhwa that the Brahmans attributed to the “Chinese”—i.e., Tibetans. There is also reputed to be a Buddhist stupa in Bagori village near Harsil; Nand and Kumar (1989: 119). 218 chapter 9 produced a copper plate inscription dated to 1667, in which the raja of Garwhal agreed that his border was below Harsil. Gum-gum nala was visited by a repre- sentative of the Tsaparang Dzongpon (Governor) as late as 1919, apparently in pursuit of Tibet’s claim to sovereignty.75 Nilang and its nearby offshoot Jadhang were only inhabited during the summer months by a few dozen Jad families engaged in cross-border trade.76 Like the inhabitants of Kuti and other places outside of pre-colonial state structures, their identity as ‘Indian’ or ‘Tibetan’ was more a product of colonial classificatory systems and political necessities than any indigenous claim; their payment of taxes to three states evidences their frontier status. When the matter came to a head in the 1930s, British colonial officials were actually prepared to concede territory beyond Nilang as Tibetan, but the claim lapsed in the face of more significant issues in Lhasa. The colonial era transition from frontiers to boundaries produced many such anomalies in Asia and the results of demarcating the boundary line on the very practical basis of the Himalayan watershed could be subsequently exploited, as this example occasionally has been by Communist China. The process involved defining not only territory, but population groups (invariably un-empowered minorities outside of the dominant state culture), as belonging to one or other nation-state; an identity previously unknown to them. In this case the people in question were fortunate in being classified as Indian. But most important to our enquiry is the evidence that the Jadh Ganga and Bhaigi- rathi sources only came under Indic rule in the modern era. The question of identifying the Ganges source raises interesting issues. In the modern world such decisions are made within scientific geography on the basis of the volume of water in various feeder streams. In the pre-modern world, however, no such precision was possible and the decision must have been reached through a variety of experiences and emotions. A hypothetical pioneering traveller in for example, the first millennium, travelling north from Rishikesh (where the Ganges emerges onto the plains), in search of the river source would, if they had no prior knowledge of its course, face a number of significant confluence points (see map 8). At Devaprayag two rivers, the Bhagirathi and Alakandanda meet to form the Ganges. Both of these rivers are formed by feeder rivers, of which the most prominent in the case of the Alakananda are the Saraswati and the Pindar, Rishi, and Dauli Gangas, which

75 Oriental and India Office Collection, l/p&s/12/4172; (various documents). 76 For first-hand accounts of the Jadh Ganga course, see Bhattacharji (1984–1985), Kapadia (1989–1990), and Ottley (1940). the mountain at the source 219 are joined at Rudraprayag by the Mandakini. The Bhagirathi is joined by an outlet of the Bhal and Bhilangna rivers at Tehri, and below Gangotri is the meeting of its two main sources, the Bhagirathi and the Jadh Ganga. Thus the actual source of the Ganges is not obvious and designating the number of its tributaries is dependent on the importance attached to particular feeder rivers. One could speak of (the classical) four rivers, the Pindar, Alakananda, Mandakini and Bhagirathi, or five with the inclusion of the Bhilangna, and so on. But it is notable that Gangotri is almost entirely encircled by the rivers which finally combine at Devaprayag to form the Ganges. The entire region can thus be sacralised (and the claim of any particular place, river, or mountain to be ‘the most sacred’ can usually be traced to sectarian bias, with Badrinath broadly classified as a Vaisnavite centre, Kedarnath as Śaivite, and Gangotri, perhaps reflecting its modern prominence, as Pan-Hindu, although with Śaivite leanings). But given the fundamental importance of the Ganges to Indic civilisation over the last three millenniums, the sacrality of its source region is axiomatic and locating Gangotri at the centre of this sacred realm is actually, from the perspective of traditional Sanskritic cosmology, entirely reasonable. Not only is it surrounded by the various Ganges’ sources but within a radius of 50 miles (80ks) from Gangotri are Jumnotri, Kedarnath, Badrinath, and even Tsaparang and Tholing. Scientific mapping had removed the Ganges from the Meru-Kailas model, replacing it with the Karnali, which in turn potentially created new sacred geographies around the conceptually remodelled Ganges source. All that was needed to envisage this as the classical sacred landscape of Indic cosmology was a central mountain. Sri Kailas thus takes on a particular promi- nence as a mountain at the centre of this sacred circle of rivers. We shall return to this point in the following chapter.

The Indigenous World

While almost all of the literature on Gangotri concerns the Sanskritic world, an indigenous world may still be identified. Swami Tapovan for example, a renun- ciate travelling through the region in the 1920s, described the local people’s strong faith in territorial guardian deities, worshipped through annual com- munity rituals common to the wider region.77 Even today there are constant

77 Tapovanam (2003: 28–29). 220 chapter 9 reminders of that local world in ritual, culture, and understandings of land- scape. While the shrine at Bhaironghati, dedicated to Bhairon the tutelary deity of Gangotri,78 is now largely abandoned,79 worship of local deities continues at other shrines such as the Naga site at Mukhwa. (As noted, such shrines are usually situated outside the village, in contrast to Sanskritised temples within villages.) Annual festivals still take place at Naga lakes and while sacrifices are being eliminated, they still survive. Local people are often reported to object to Himalayan mountaineering,80 and whether expressed in terms of Sanskritic or local sacralities there is some apprehension concerning outsiders in this region. The fear is that their igno- rance of the correct behaviour required by the local deities will lead to actions that upset these powers, with the ensuing punishment likely to fall upon the local community. In 1928, for example, Swami Tapovan travelled with a small party of renunciates to a lake between the Jamuna and the Ganges, thirty miles north of Uttarakashi. He refers to this lake, the site of a bi- or triannual local festival of bathing the village deity in the lake, as Dhumdithal [Dodital], the source of the Asi Ganges tributary. At the local village, he found that;

some of the women and the old people objected to our attempt. They said that the lake was the abode of gods … and men could not reach it … If somebody arrived there in a spirit of adventure … the gods would send down heavy hailstorms, destroying all the crops in the villages.

But “a small band of men with progressive views”, including a guide of “the Brahmin caste among the mountain folk”, ruled that the renunciates’ worship at the lake would be beneficial, and took them there.81 Thus while in this case facilitated by local agents with a wider than normal world-view, grounds for local opposition to outsiders, even renunciates, are apparent. The account by James Fraser, whose visit to Gangotri came only a decade or so after the temple there was built, is worth quoting in full, for it suggests other issues limiting the pace of transformation from local manifestations of the Western Himalayan Cultural Complex to World religions.

78 Skinner (1832: vol. 2., 47) however, refers to “Bhairo, a saint after whom the ghat … is named.” 79 Sundaranand (2001: 54). 80 An Indian climbing party on Kinnaur Kailas reported local opinion that as Śiva’s residence it was unclimbable; but the (Hindu) mountaineers responded that “as climbers we had challenged Kinner Kailas and not Lord Shiva.” Kumar (1981: 86). 81 Tapovan (2003: 27–28). the mountain at the source 221

In former times no temple made with hands was provided for the worship of the deity, but …, the piety of Ummr Singh Tghappa [sic], the chief of the Ghoorkha conquerors appropriated a sum of money … for the erection of the small building which is now placed there; and it by no means clearly appears whether he has in truth done an act pleasing or disgusting to the goddess: on this subject the Pundits answers were by no means explicit. Jumna prefers simple worship at the foot of her own and natural shrine and has forbidden the erection of temples to her honour; and probably were it not for the comforts which have accrued to the holy Brahmins of Gungotree … their attachment to ancient customs would have induced them to declare that the original mode of worship was also here most acceptable. It is remarkable, if there be not some superstitious reason for the non erection of temples, that among the number of donations presented to this establishment by the many pious persons who have visited it either themselves or by proxy, in the lapse of so many centuries, there should never have been any thing of the sort, nor any provision for a building. Whereas at Kedernauth and Buddreenauth, places not in themselves as holy as this, there are places of worship. It is, therefore, fair to suppose that the want of them here is not from neglect, but in consequence of some superstitious prohibition.82

Fraser does indicate here a pre-modern Gangotri sacrality, but clearly that sacrality was local.83 We have seen that the texts of many Indic schools of thought, Epic, Tantric, and so on, contain explicit statements concerning access to the sacred heights of the Himalayas; it was the land of the gods and ordinary persons could not and should not, travel there. Fraser’s statement adds another dimension to that prohibition, that it reflected local understandings. The inner realm of the Himalayas was beyond the world of temples and tem- ple Brahmans. Renunciates did not need temples and their images, nor did followers of the local belief system have them. Naga stones and suchlike were placed in geomantically auspicious outdoor locations, shrines, not temples. When renunciates entered that local world and confronted and Sanskritised local deities, they did pave the way for the construction of temples, but these developed organically from the renouncers’ shelters and after gaining local acquiescence could attract elite (Sanskritic) patronage. The sudden erection

82 Fraser (1820: 467). 83 Channa (2013: 258–259, 271) notes that the Jad have drawn Gangotri into their sacralised territory only since their relocation from Nilang in the early 1960s, prior to which it “did not play any significant role in their lives”. 222 chapter 9 of a temple circumvented that organic process; it was an alien structure inap- propriate for an impure, untransformed local space. Here we might extrapolate a Himalayan process. Until recently, beyond the inhabited Sanskritic world and beyond the agricultural fringes of settlement with their focus on female fertility or ‘mother’ goddesses, was an inner ring of mountain territory whose inhabitants held to the local deity system of the Western Himalayan Cultural Complex. That territory was textually stated to be open only to renuniciates, who could transact with local populations and deities through personal charisma and siddhi power. If, as non-state agents, they were killed by a hostile local population acting to forestall their deity’s wrath—as certainly occurred on occasion84—it was inconsequential. But if they were able to establish themselves, they served as agents of gradual trans- formation to the understandings of a World-religion, ultimately paving the way for temple-based popular pilgrimage. For Śaivites, those mountain realms were Śiva bhumi,85 the widest impli- cations of which could only be understood by the enlightened vision of the renunciate. Only they could properly perceive the heavenly reality behind the prosaic—the land where Arjuna or Rama had walked, the yogini behind the village woman, the deity manifest in a rock. It was a dangerous place, a frontier which, remembering Heesterman, we might call the Zone of the Consecrated Warrior. But this zone has shrunk over time. Once explored and purified by the presence of the Consecrated Warriors (rsis, Pasupatas, Naths, Giris and their like), it was opened to the ordinary Sanskritic pilgrim. Temples, roads, and settlements followed. Temples marked the assimilation of territory into the Sanskritic world and as they were erected higher and higher in the mountains so the local world, the Zone of the Consecrated Warrior, shrunk. Over time Badrinath, Kedarnath, and finally Gangotri were transformed from heights where (ideally at least), only renouncers ventured, into centres of Sanskritic culture drawing on the mythology of Epic or Puranic literature. Today that Zone has shrunk dramatically. As the Himalayas are transformed into tourist playgrounds it exists only beyond the roads and new settlements, in isolated hamlets or unpopulated heights. That process is now almost complete. But the sudden imposition of a temple at Gangotri before the region had been prepared by the sustained presence of renunciates was a cultural leap that

84 See, for examples from the eastern Himalayas, Huber (2011: n. 14 272). 85 In 2003, at Manimahesh Kailas, I asked a Giri renunciate, Pasu Ram Giri (chela of Sri Sri 108 Giri Maharaj), who had then spent July–October there for the last five years, if the surrounds were Śiva bhumi? “Yes” he replied: “Everything for 2000 miles around is Śiva bhumi.” the mountain at the source 223 challenged renunciate agency, short-circuited organic progression, and thus upset local sensibilities.86 More than a century passed before Gangotri became properly Sanskritised territory, and that transformation was due to renunciate agency. Again, just a few individual renunciates were responsible for that transformation, which concerned not only Sri Kailas and the Ganges sources, but, in a linked devel- opment, the Kailas-Manasarovar complex. In the following chapter we will examine that development, and discuss the Kailas mountain at the centre of the Ganges’ sources.

86 The implications of a temple in terms of its representing the spiritual supremacy over the surrounding landscape of a particular belief system are clearly recognised; e.g.; the construction of a Hindu temple (‘Baba’ mandir) near the Nathu la in Buddhist Sikkim is not universally popular. chapter 10 Sri Kailas: The Epic Prototype?

Modernity Meets the Mountains

We have seen that prior to the 19th century neither Kailas-Manasarovar nor the source of the Ganges were part of a consistent Pan-Indian body of knowl- edge. Textual references to the upper Himalayas did contain some reflections of actual geography, but also imagined spaces and cosmologies and the vari- ous accounts were internally inconsistent and collectively contradictory. This is consistent with the existence of discrete bodies of knowledge which arose in different regions, gotras, sects, etc., and means we may identify Vedic, Epic, var- ious Puranic interpretations, Tantric, royal, Vaisnavite, Śaivite, Buddhist, local and many other understandings of the Ganges source and the sacred mountain and lake(s) above it. These traditions may already have been in communica- tion before the British period for the appearance of regional mahatmyas (the Manasa- and Kedara-khandas), appears to date to the immediate pre- or early British period and thus may reflect local processes. But the mahatmayas, if indeed they do predate the British, were the last note of traditional understand- ings before the scientific mapping revolution. The Survey of India provided an almost entirely distinct body of informa- tion about the Himalayas. This challenged the earlier forms of knowledge; what was once a mythological and sacred visionary reading of the landscape now became a modern map. But far from discrediting the earlier bodies of knowl- edge this added dimension only enhanced the appeal of the site. As with other forms of traditional knowledge (such as medicine), a successful reformulation occurred as traditional knowledge was re-interpreted in the light of scientific knowledge. Those who led this process actively embraced science as a means to bet- ter interpret their earliest textual sources and believed that the new body of knowledge actually explained and validated much of the traditional. They con- cluded that many aspects of the mythical land of Kailas could be matched to reality, that upper Himalayan geography fitted the descriptions of the land through which had travelled Arjuna and the Pandavas, Rama and the rsis. Where ancient Hindu cosmology had drawn on the model of a central moun- tain at the source of great rivers there was now such a place. By emphasis- ing correspondences rather than contradictions in the texts, and by assuming toponymical correspondences of the Gokarna=Gangotri kind, they constructed

© koninklijke brill nv, leiden, 2015 | doi: 10.1163/9789004306189_012 the epic prototype? 225 a new understanding of the sacred landscape that was informed by earlier bod- ies of knowledge. In what follows we identify the key agents in this process and discuss their reformulations of the ancient sources.

Renunciate Responses to a Modern World

There are close parallels between the modern Indian constructions of the sacred geography of Kailas-Manasarovar and of the Ganges headwaters. Nei- ther site can be reduced to a single historical narrative. Both are imagined heavenly realms transformed to earthly locations, and the history of both is a tangle of competing mythologies, shifting toponyms, and manifestations of dif- ferent cultural groups. What unites them however, is that both constructions involved a small and closely linked group of modernist Indian renunciates. We will see that Western writers have had a determining influence on the modern understanding of Kailas, but these renunciates not only assisted in that con- struction but also responded to it with their own constructions. The two sacred geographies—of Kailas-Manasarovar and of the Ganges’ source—were mutu- ally reinforcing, yet expressed in different cultural idioms. They represented not so much a dialogue, as a synergetic translation. As noted, from as early as the Vedic period onwards Indic renunciates trav- elled increasingly widely in the Himalayas, venturing beyond the shifting cen- tres of political power into local worlds where they acted as Sanskritising agents. In the British period both Kailas-Manasarovar and the Bhaigirathi and Jadh sources of the Ganges remained spaces outside of direct imperial control with colonial interests, such as they were, focussed on Badrinath and Kedar- nath which were under colonial government authority. This allowed renunci- ates in Tehri-Garwhal to occupy both traditional and modernising roles in the late 19th and early 20th centuries, which made this state a dynamic centre for the organic transformation of Hindu beliefs. Re-imagining, re-inventing, and re-creating earlier bodies of knowledge in line with scientific geography, the renunciates found new ways in which to be the primary interpreters, indeed authorities, of the sacred Himalayan spaces. We have seen that many Hindu authorities accepted the Western conclusion that the history of Indic religions was one of ritual and moral decline from pure spiritual origins, a doctrine that could be equated to the Hindu concept of the Kaliyuga. One response was that reformers sought to refocus Hindu beliefs on the earliest textual sources of their religion in order to dispense with elements they now understood as ‘tribal’,or ‘non-Aryan’ adoptions polluting the purity of Vedic lore. This tendency was not entirely new to the Indic world, but as was the 226 chapter 10 case with so many strands of belief, political circumstances favoured its wider emergence at a particular time. The doctrine of decline carried with it the implication of an earlier ‘Golden Age’. The interests of the Vedic-orientated reformers seeking to recreate that imagined Age coincided with the broader Indian nationalist movement in that colonial subjection could be seen as a product of the decline and a recre- ation of a Golden Age a means of restoring their nation.1 Hindu reformers, not least renunciates, were consequently able to build coalitions of interest with the newly emerging Indian political class. The waning power of their earlier patrons, the royal courts of medieval India with their demand for the Atharvavedic rituals of power and the orthodox Brahmanical rituals and sym- bols of legitimacy, saw significant elements of the renunciate movement seek patronage from that new elite. To do this they used a number of strategies that accorded with their new beliefs. They joined opposition to widow-burning, embraced scientific progress, and even de-emphasised caste—creating a mod- ern Hinduism for a modern India. This intellectual adaptability seems char- acteristic of the renunciates throughout Indic history. Just as we saw with the Magi, they remained at the forefront of socio-political developments by their adaptability, initiative, and malleable social authority. In the context of this study, there were two particularly influential figures behind the reform movement. One was the blind Punjabi Sanskrit grammarian, Swami Virjananda Sarasvati (1779–1868) who was initiated in the Himalayas by Swami Purnananda Sarasvati, but found his patrons in Rajasthani royalty. Swami Virjananda sought to counter the perceived decline in Hinduism by distinguishing the literature and tradition of the Vedic rsis from later writings. Those which proved deviant or later were to be rejected.2 The second figure was Swami Virjananda’s disciple, Dayananda Sarasvati (1825–1883). From a Gujerat (Punjab) Śaivite Brahman family, Swami Dayanan- da was also initiated by one Swami Purnananda Sarasvati (1824–1883; presum- ably not the same individual as Virjananda’s guru). He then became a pupil of Swami Virjananda after living an itinerant existence from 1846–1860. Swami Dayananda followed his teacher’s repudiation of scriptures deviating from Vedic precepts, in particular the Puranas. He thus rejected most popular Hinduism, including caste and the ban on widow remarriage and he preached,

1 There is a considerable literature—much of it highly theoretical—on these social develop- ments in colonial India; see for example, Jones (1989); for a convenient summary, see Robb (2002: 218–245). 2 Lajpat Rai (1967: 21–25); also see http://www.aryasamaj.com/enews/2009/dec/2.htm; http:// ravitiwari.in/rtpaper3.pdf; accessed 9 September 2015. the epic prototype? 227

“a rationalistic monotheism … supported by a new interpretation of the past.” His vision bought him into contact with Western Theosophists, although they later turned away from his dictates.3 His followers recall that;

His soul was full of purity and [the] greatness of India’s glorious past … He was taught to have implicit faith in the ancient books written by Rishis … to wage incessant war on the falsehood of Puranic Hinduism and restore the true teaching of the Vedas.4

In 1875, to further those goals, Swami Dayananda established the Arya Samaj, one of many 19th century Hindu reform movements (and one which, perhaps ironically, was opposed to pilgrimage). This influential group;

was a movement chiefly among urban commercial castes (…), often edu- cated and prosperous. They were creating an organised, regulated and proselytising Hindusim … a disciplined body of the faithful and the book.5

Following these principles (although increasingly being prepared to allow some validity to the Puranas), was the Kailash [sic] ashram in Rishikesh. This was founded in 1880 by Sri Swami Dhanraj Giri Maharaj (1811?-1901/11?) on land gifted by the raja of Tehri.6 Around 1922–1923, the head of this ashram was Swami Jnananandagiri (1896–1969) who is said to have visited Kailas- Manasarovar in 1924 wearing only a loincloth. But Swami Jnananandagiri was clearly part of a modern world. Not only was his hair cut short (as was that of Swami Dhanraj), but he also gained a science diploma in Prague and a doc- torate from Liverpool university. Later the author of a number of scientific papers in the field of radiology,7 he eventually abandoned the renunciate world for a distinguished career in scientific academia in America and India. Swami Jnananandagiri’s importance to us lies in the influence he had on two of his disciples, Swami Tapovan and Swami Pranavananda.

3 Johnson (1994: 107–115; quotation from Jones [1976: 33], cited on 111). 4 http://www.indiastudychannel.com/resources/40860-Maharshi-Swami-Dayananda -Saraswati.aspx. Accessed 4 July 2010. 5 Robb (2002: 235). 6 Giri (1976: 80–81): References to the Archarya (abbot) of the Kailash ashram vary somewhat, apparently reflecting internal issues. Aside from two brief periods under a Puri, the Archayas have all been Giris. 7 Pranavananda (1983: n. 2 25). 228 chapter 10

Swami Tapovan

Sri Swami Tapovan Maharaj (1889–1957), was a Kerala Brahman of the aris- tocratic Nair family. Previously noted here as the author of the Sri Gangotri Kshetra Mahatmyam, he first went to Gangotri in 1918 and in 1921 built his still- surviving hut there. Swami Tapovan is however, most associated with the plain of Tapovan above Gaumukh, where he carried out his religious practice and which, according to his disciple, is named after him rather than the reverse.8 It was Swami Tapovan who provided the names of many places including Sri Kailas and Meru to the scientific survey of the Gangotri region in 1936.9 The survey was led by Major Gordon Osmaston, who noted that;

New maps were badly needed because of the increasing number of trav- ellers coming to the Himalayas with a genuine interest in the high glaci- ated regions. These highest areas had not been visited by the surveyors when the existing maps were made sixty and seventy years ago, and the few details shown were unreliable.10

It seems entirely logical that Osmaston would take Swami Tapovan as his informant. The Swami was the most prominent religious figure associated with the area and the most authoritative presence in the region at that time who could have provided the required information. The Gangotri pandits are unlikely to have had a close knowledge of the terrain above Gaumukh,11 and would presumably have advised Osmaston to seek out the Swami for such information. Around the time of Osmaston’s visit a mountaineer, J.A.K. Martyn, desig- nated a “Chandarpur Kailas”12 on a map, some 12–15 kilometres southeast of Sri Kailas. This toponym, not otherwise known, identified a mountain known today as Chandra Parbat (6728 mtrs) and there may have been some confu- sion between the two. But the possible existence of an oral tradition of a Kailas mountain in this vicinity is indicated by two early 19th century travellers, both of whom were generally reliable observers. William Moorcroft recorded two

8 Pranavananda (1983: 66, 70); Interview with Swami Sundaranand, August 2008. 9 Interview with Swami Sundaranand, August 2008; also see Sundaranand (2001: 112, re “the plain of Sundarvan named after the author by his Guru.”) 10 Osmaston (1939: 126). 11 Both Tapovan and the areas above Nilang were used by Jad shepherds; Channa (2013: 126–127). 12 Martyn (1938: map, 82). There is no evidence for Chand influence in the Gangotri region. the epic prototype? 229 mountains known as “Cailas or Mahadeó Ka Ling”—Kailas Manasarovar in Tibet and another near Gangotri.13 The surveyor J.A. Hodgson reported that a Gangotri Brahman who had visited Nilang informed him that from there, “the Cylás peaks towards Gangotrí are seen to the right”,14 which is where Sri Kailas is located from that standpoint. But in both cases the identification may be qualified. Moorcroft’s terms sug- gest a possible confusion with the extraordinarily prominent Sivling (i.e.: Śiva’s lingam) mountain above Gaumukh, while Hodgson qualified his statement with a note stating that “Cylás is a general appellation for high ranges always covered with snow.”15 This was a problem his successors also found, but it did not preclude the understanding of a specific Kailas as we saw in regard to Kin- naur. Thus it is possible that the toponym ‘Sri Kailas’ was coined by Swami Tapovan, but it is also possible that it was within an existing body of (oral) knowledge. But we know that Swami Tapovan did name features of the landscape around Gangotri and to understand the process by which he identified them we must look more closely at his world. Educated in English and able to quote from philosophers such as Tolstoy and Schopenhauer,16 the Swami showed an early partiality to the spiritual world. Abandoning formal education he lived as an independent intellectual, becoming editor of a magazine inspired by influ- ential nationalist Hindu reformer and Congress party leader Gopal Krishna Gokhale (1866–1915). This brought him to the attention of the Sankaracharya of Dwaraka math, who invited him to Calcutta. There, and in subsequent travels, Swami Tapovan met many great religious figures of his time before taking san- nyasi in 1923 at Nasik, where he was initiated by Swami Jnananandagiri. After further periods of study with various teachers and journeys to sacred sites in India and Tibet, including two visits to Kailas-Manasarovar, the Swami centred his world around the Gangotri region. At Gangotri he fulfilled the expectations of an advanced renunciate, writing Vedanta commentaries on the Upanisads and other spiritual reflections including, as we have seen, the mahatmya of his adopted home.17 He was thus the key figure in opening Gangotri to popular pilgrimage and promulgating the understanding of it as a Pan-Indian sacred centre.

13 Moorcroft (1816: 415). 14 Hodgson (1822: 91). 15 Hodgson (1822: 92). 16 Tapovan (2001: 14, 34). 17 Tapovan (2003: v–xiii); Bharati (1963: 162) describes the Swami as, “one of the most cele- brated contemporary commentators on Vedānta philosophy.” 230 chapter 10

Swami Tapovan’s career cannot be seen in isolation. For example, although initiated into the Giri order he is not easily associated with any particular sect of renunciates. Though a Śaivite in personal ritual he is not easily distinguished as either Vaisnavite or Śaivite by his writings. While its introduction reveals that he observed a Śaivite ritual calendar,18 one of his best-known works is a Hymn to the Vaisnavite centre of Badrinath (whose Mahant sponsored his second Kailas-Manasarovar journey). Swami Tapovan was thus a Pan-Indian Vedantist Hindu; modernist, monotheist, non-sectarian, and Vedic orientated. His concern with tradition can be seen in his travelling to Kailas-Manasarovar via the Mana-la, “because [he believed] it was the only route used by the ancients on their journey to Kailas.”19 Yet he was also influenced by modernist tendencies, as can be seen in that, along with other reformers he embraced scientific enquiry, rejected Tantric practice and did not consume intoxicants.20 Swami Tapovan was literate in Malayalam, English, Sanskrit, and a number of regional languages. While resident in a Princely state he was linked to the nationalist struggle in the sense of seeking a better India. He was aware of the modern world and actively sought to integrate it into his understanding of the teachings of the Vedic rsis. He could not ignore the scientific world, the new Indian political class who came to patronise him included many who embraced the concept of scientific modernity.21 Swami Tapovan actively sought therefore an encounter with scientific knowledge, a tendency that followed from his initiating guru, Jnananandagiri. Swami Tapovan did not found an order, initiate any disciples,22 hold any official position, or establish an ashram. But he was an outstanding figure in a lineage which drew on the earlier reformulations of Swami Virjananda Saras- vati and Dayananda Sarasvati (apparently) passed via a Swami Purnananda (who is difficult to identify23), and transmitted to him by Swami Jnanananda-

18 Tapovan (1984: i–vii). 19 Tapovan (2003: 220). 20 Interview with Swami Sundaranand, August 2008. 21 His journey to Kailas-Manasarovar was arranged by the brother-in-law of the then Nepali Prime Minister, who was a devotee; Tapovan (2003: 172–173). 22 Tapovan (2003: xi). 23 Swami Jnananandagiri’s guru was Swami Purnananda (1834–1928); As noted, Swamis Virjananda Saraswati and Dayananad Saraswati were also initiated by (another) Swami Purnananda. It is difficult to draw up precise lineages of these figures, not least because of the replication of names, but also because their initiating guru is not necessarily the figure most influencing them. The idealised concept of unbroken lineages of instruction seems less formalised in Indic traditions in contrast to Tibetan. It does seem clear, however, that there were close links between the Sarasvati and Giri lineages. the epic prototype? 231 giri.24 It was from the knowledge held by or developed within this lineage that the modern sacred geography of the Gangotri-Sri Kailas landscape emerged (and which also had a significant influence on that of Kailas-Manasarovar). As the figure who synthesised the teachings of his lineage and promulgated a new vision, Swami Tapovan was the primary agent in that construction. Although he did not have a formal chela, Swami Tapovan did have several pupils of whom the most renowned is Swami Sundaranand. A keen photog- rapher (he is also known as ‘Photo-Swami’), and author of the outstanding photographic study of the region,25 Swami Sundaranand is a Brahman born in Andhra Pradesh in 1926. He first travelled to Gangotri in 1948 and at the time of writing was still resident there in Swami Tapovan’s old hermitage. The Swami is well-known throughout the Himalayas. He numbers among his acquaintances numerous Prime Ministers, politicians, mountaineers, and travellers, and he has acted as a guide for various civil and military Himalayan expeditions in addition to carrying out his own spiritual practices in the mountains. Like Swami Tapovan, Sundaranand has come to know this region extremely well. He has taken the exceedingly difficult—once thought legendary—route from Gaumukh to Badrinath (via the Kalindi pass), eleven times. Like his predeces- sor, he also has travelled to Sri Kailas (Swami Tapovan’s visit being in 1921 when he travelled on to Mount Kailas in Tibet).26 As a well-connected Brahman renouncer following the modernist path, Swami Tapovan appealed to modern Indian society, where he found his pa- trons. While accepting that patronage to support his travels and print his sacred writings, his authority was only enhanced by the fact that he actually seemed to have no desire for wealth, power or fame. He lived in humble surroundings near the Ganges source, undertook mountain pilgrimages, and accepted only a handful of students whom he did not initiate into a sect. He represented the ideal renunciate, the very image of the Himalayan rsis of popular imagination

24 The suffix ‘ananda’ is often used in the names of prominent renunciates rather than ‘giri’. This is difficult to account for. It may be that as titles such as ‘Maharaj’, ‘108’, ‘1008’, ‘Sri’, ‘Mahaguru’, and so on are endowed upon them, sectarian affiliation becomes less important and a Pan-Indian identity is emphasised. (This is suggested by Ghurye [1964: 83]). Also possible is that the term giri retains certain negative connotations in dealing with wider society and is thus quietly dispensed with as renunciates grow in status. However, Sarkar and Roy (1957: chapter iv) state that the ten names of the Dasnami divisions are divided among the four maths and that at Joshimath, “[n]ovices take the title of Ananda.” 25 Sundaranand (2001). On Sundaranand see Malhotra (1983: 141–154). 26 Sundaranand (2001: 129–130). 232 chapter 10 and textual account. For those embarrassed by what they perceived as the greed and excesses of many Hindu gurus, the ascetic Swami Tapovan provided an alternative as an archetypal renouncer who resisted the lure of worldly plea- sures. He is still revered in the region today, with his writings and photographs of him on sale in Gangotri bazaar, and we can assume that his ideas had a wider support system among his renunciate contemporaries. Certainly they did in the person of his devoted disciple Swami Sundaranand. Through writings and teachings of his own and of his followers, Swami Tapovan’s account of this region became the predominant model. Given his patent lack of interest in wealth, power and fame—all of which he could certainly have obtained had he wished—there is no reason to doubt Swami Tapovan’s motives in his construction. Following a lineage emphasising Vedic study, he allowed the authenticity of a rather wider range of texts than Swamis Virjananda and Dayananda, with his commentaries on the Upanisads and embrace of Epic and Puranic references. His criteria was also open to revelation by those of his own Brahmanical renunciate traditions.27 But he must have genuinely believed that his own vision of the landscape was that of the rsis, that like them he was the possessor of visions of divinely revealed knowledge. Conceptually, and in the eyes of his followers, the revelatory visions of the Swami were entirely valid in the traditions of the Vedas. Swami Tapovan’s construction was broad-based, drawing on and synthe- sising a range of sources including functionaries of the local religious world. Given his non-sectarian background (at one point he cites the Narada Bhakti sutra; “Prostrations to Shri Vishwanath, the Lord of Kailas”!28), the Swami felt free to draw on other religious traditions and bodies of knowledge including those revealed by British geographers and scholars of Hinduism. He must have known of the Manasa- and Kedarakhandas—and the doubts over their antiq- uity. He must also, as a South Indian like so many other renouncers (and ritual priests) in the Himalayas,29 have bought his own pre-conceived images of this region with him just as Westerners take such images to their encounter with new lands.

27 Swami Tapovan (2003: 157) considered images useful devices to bring awareness and his return to the ancient texts did not exclude the Puranas. A process by which the ‘fundamentalism’ of sectarian founders is lessened over generations seems characteristic. 28 Tapovan (2001: 1). The Narada Bhakti sutra is dated to around 1000ce. by R. Raj Singh (2006: 76). See however, Bangha (2011: 140) who implies a 16th century date as more likely. 29 i.e. the pujaris at Badrinath, Kedarnath and at Pasupatinath (Kathmandhu), are also South Indian. the epic prototype? 233

How Swami Tapovan, seeing himself as divinely inspired, read the land- scape of Gangotri as the sacred ancient site of the rsis was in academic terms, through a complex combination of factors including religious vision, intuition, and learned correspondence of toponyms. In his understanding certain places existed in the texts, therefore they must exist on the landscape. He even refers to aspects of this process in his writings. Of one site he wrote; “[t]here is now no means of ascertaining the Puranic name of this holy place but it is beyond doubt that the spot was once the favourite abode of rishis.”30 In other words, his reading of the landscape revealed that a particular site was sacred. Similarly he describes the (sadly no longer surviving) birch-forest (Bhoorjavanam) en route to Gaumukh as “a favourite resort of saints and ris- his”;31 the former probably true of his time, the latter assumed of the past. The point is that once this way of reading the landscape is entered into, with the appropriate learning and culturally assumed authority, any feature of it can be identified as having sacred associations according to the intuitive visions of the beholder. (Visions characteristic among cannabis users, not least mod- ern renunciates there, but Swami Tapovan rejected such vision-inducing devi- ces.32) Swami Tapovan, along with Swamis Jnananandagiri, Sundaranand and Pra- navananda (see below), are, according to Sundaranand, the only renunciates to have visited Sri Kailas in the last century.33 These men were well aware of Kailas-Manasarovar. Given their scientific tendencies, they understood the geographical layout of the entire region. They must, therefore, have been struck by certain paradoxes. Most notably Sri Kailas was situated at the centre of the Alakananda, Bhaigirathi, and Jadh sources of the Ganges. Whether they knew this mountain by that name or themselves named it, they recognised that its location implied its supreme sanctity. Yet knowing the Kailas in Tibet, they also accepted the sanctity of that site. While using the texts as a guide to reading the landscape, Swami Tapovan accepted that the Puranas were not literal truth and did not always “agree with experience and reason.” But he considered that they contained;

many valuable truths or at least portions of the truth … Those ancient rishis had recourse to fiction and allegory and exaggeration so that they

30 Tapovan (2003: 26); emphasis added. 31 Tapovan (2003: 43). 32 Interview with Swami Sundaranand, August 2008. 33 Interview ibid. 234 chapter 10

might bring home to the minds of people at all stages of mental culture, those eternal truths which they wanted to impress upon them.34

Where contradictions could not be resolved, he did not critically confront the issues. Ultimately he saw the gap between the academic and the spiritual as “unbridgeable” and one that as a devout Hindu he would not engage with;

I do not intend to tread upon those dangerous grounds. Indeed, criticism is like a boundless sea. Once you get into it, you can hardly get out of it … In all spiritual matters (…) it is faith that matters, not intelligence.35

Ultimately contradictions and paradoxes were more easily resolved by the religious understanding than by the academic.

It matters little [Swami Tapovan concluded] O Mother Ganga, whether you originate at Vishnu’s big toe or at Sri Gomukha … the fact remains Ganga is Ganga. O Mother, thou art the Supreme Mistress of all the world.36

We will see that Tibetans visiting the western Himalayas were similarly reading that landscape in the form conditioned by pre-conception, by cultural under- standing of religious Heroes of the past, and charismatic self-authority. The renunciates were operating in a location far from central institutions, they implicitly claimed the right to exercise their autonomy and the authority to interpret the divine world in which they existed. The location in which they operated was a liminal one, a transitional frontier zone that required different qualities and responses to those considered necessary at the centre. As mas- ters of a sacred landscape their autonomy was accepted. They had gone beyond worldly realms and so were free of its dictates. Once the construction of the Gangetic centre was formulated in broad outline, its details could be filled in and its claims supported by additional identifications and revelations from Swami Tapovan’s support networks and successors. Study of earlier texts could only add to possible identifications, with those texts examined for supporting evidence of the construction rather than critically examined from the standpoint that the texts emerged from an entirely

34 Tapovan (2003: 42). 35 Tapovan (2003: 35–36). 36 Tapovan (2003: 266). the epic prototype? 235 different context and even culture from that of the contemporary world. While the academic approach stresses the lack of continuities, the emic perspective understood the texts and the world in which they were used as fundamentally unchanging. We may link geo-political developments to this construction process. Brit- ish-Indian expansion brought about the need to define their northern bor- der. While Tehri-Garwhal state was not under direct British rule, any actions affecting India’s external border with Tibet were a British concern. Centring Gangotri in the Sanskritic world was not however, antithetical to colonial inter- ests, indeed the reverse was true and it can be seen as a political strategy to define that border and to link the population inside it more firmly to the centres of Indic religious, political, and economic worlds. The promotion of pilgrimage there certainly served the economic interests of the local popula- tions and raised their social status in the Hindu caste system. This is not to suggest that this was more than a fortunate conjunction of interests between religious and political expansion. But that conjunction did strengthen the new understandings of the Gangetic realm, with both British political and Indian religious authorities supporting the construction and as we see throughout this history, the desire of central authorities to strengthen their borders through the creation of universalising symbols has been of primary importance in the devel- opment of Kailas mountains. Yet it is notable that in contrast to the promotion of Kailas-Manasarovar in European languages, there seem to be no significant Western language works focussing on the Gangotri region. Indeed the definitive colonial text on Kailas37 (see Chapter 15) managed to entirely avoid mentioning Gangotri and sought to steer pilgrims through the British-controlled Badrinath route to Tibet rather than via Gangotri. The processes stimulating interest in Kailas-Manasarovar included a significant political dimension in the sense of strengthening ties between British India and the Tibetan government and the absence of such stimuli here in regard to Sri Kailas does support a primarily emic construct initially at least, with its intended appeal primarily Indic. There is one possible interpretation of Swami Tapovan’s vision that cannot be disproved. This is that it was not a modern imagined (re)construction, but the public revelation of an unbroken chain of restricted knowledge dating back to the Vedic rsis. Given the traditions of ‘secret knowledge’ orally transmitted from guru to disciple it is perhaps plausible, but the possibility need not detain us here for there is no evidence to support it. But it should be noted

37 Sherring (1974: [first published, 1906]). 236 chapter 10 as a powerful emic conceptual device used by Hindus (and their New Age associates) to further the authority of such visions.

Swami Pranavananda

Swami Pranavananda was, and indeed remains, the predominant Indian au- thority on the Kailas-Manasarovar region. He was born in Andhra Pradesh in 1896 and after graduating from college in Lahore, briefly worked for the Indian railways. He then joined the Congress party and was active in the ‘non- Cooperation’ movement before taking sannyasi around 1926.38 In his later writ- ings however, he still referred to Mahatma Gandhi as “the world’s greatest man.”39 Pranavananda’s guru was the scientist-sannyasi Swami Jnananandagiri and Pranavananda was thus guru-bhai (brothers by virtue of sharing the same guru) to Swami Tapovan and heir to teachings that sought to blend the Vedantic with the scientific. Inspired by his guru’s visit to Kailas-Manasarovar, Pranavananda made his first journey there from Ladakh in 1928 and his second via the Jadh Ganga route in 1935. He then became a regular visitor to Kailas, including a stay of 16 months there in 1943–1944. In all he completed 23 parikramas of Kailas and 25 of Manasarovar and while he does not seem to have known Tibetan and was largely unconcerned with the Tibetan understanding of the site, he acquired considerable empirical knowledge of the complex. As a modernist, his enquiries were rooted in science. He was a member of the Royal Geographical Society and collected fossils and rock samples that he sent to the Geological Survey of India. He even took soundings of Manasarovar from a boat he had carried up there, and he revisited the question of the river sources in the vicinity.40 In this and other scientific areas he was often critical of Sven Hedin’s find- ings,41 but his most famous work included a page long quotation from Hedin (on whom see Chapter 15) describing the physical beauty of the sacred site— something emphasised throughout his work. He also borrowed Hedin’s famous description of Manasarovar. In his words it became, “a pure turquoise set

38 Snelling (1990: 273–274). 39 Pranavananda (1983: 96). 40 Pranavananda (1946). The collections of material objects he gifted to Lucknow museum are unfortunately not now identifiable, or even known, there. 41 See for example, Pranavananda (1983: 28, 30–31). the epic prototype? 237 between the two mighty and equally majestic silvery mountains … Kailas … and … Gurla Mandhata.”42 The rationalist aspect of Pranavananada was prominent in his writings. Not- ing “sensational articles” in the press concerning Tibetan mystics, he observed that he “did not come across any great siddha or a yogi worth mentioning in the whole of Western Tibet.” But he did not deny that such individuals existed and proposed his own guru Jnananandagiri as an example of an “advanced soul.”43 While concluding that such souls were as rare in western Tibet as anywhere, he defended the sanctity of the site itself;

It is really regrettable to find some people fabricating curious and funny stories which are utterly false to trade upon the credulity of the innocent and religious-minded folk. There is no doubt, however, that the surround- ings of the Holy Kailas and Manasarovar are highly charged with spiritual vibrations of the highest order …44

Pranavananda wrote two overlapping books on Kailas, both of which were reprinted.45 While the first had a more scientific emphasis, the second was a pilgrimage guide in the modern form of a tourist guide-book. Along with notes on such matters as ethnography, flora, fauna, the administration of the region, and its history and culture, his Kailas-Manasarovar detailed the routes from India to the site, provided lists of provisions necessary for the journey and souvenirs to buy, gave the estimated expenditure necessary, and even provided the names of persons en route who could be relied upon to help the pilgrim. The pilgrimage, he assured his readers, would “engage the attention [of] young or old, man or woman”,46 although due to the risk of robbery he advised that “intending pilgrims and tourists should go in batches and take a firearm.”47 Pranavananda was, perhaps even more than Swami Tapovan, a scientific modernist, frequently citing the discoveries of British explorers and implic- itly accepting the conclusions of Charles Sherring and Sven Hedin as to the antiquity of the pilgrimage there. Yet like his guru-bhai, he seems to have had little contact with the British, finding his influential patrons among the newly

42 Pranavananda (1983: 200). 43 Pranavananda (1983: 25). 44 Pranavananda (1983: 26). 45 See Pranavananda (1949/1983) and (1950). Earlier versions of these works exist, in both English and Hindi. For details see Pranavananda (1983: 239); Snelling (1990: 268). 46 Pranavananda (1983: 22). 47 Pranavananda (1983: 190). 238 chapter 10 emerged nationalist-intellectual class. In addition to Swami Jnananandagiri he thanked a long list of high-ranking patrons in his books,48 and his personal status is indicated by Jawaharlal Nehru’s contributing a foreword to his Kailās- Mānasarōvar. Pranavananda did claim that the beauty of the site would appeal to all, but his intended audience was Indian. He promoted it as a sacred site to Hindus and his books were originally printed in small quantities in Calcutta for an educated local audience. Publishing in the last years of British rule, Prana- vananda looked to a post-colonial Indian nation and while adopting modernist views influenced by the West, his was a Hindu Kailas-Manasarovar. While Swami Tapovan became the leading authority on the Gangotri region and drew on both his traditional learning and visionary authority to iden- tify features of that landscape with those of ancient Indic texts, his guru-bhai Pranavananda was more of a scientific compiler of information about Kailas- Manasarovar. Both were devout Hindus and as a result neither man publicly explored the contradictions of their field of knowledge.49 Rather they sought to reconcile those contradictions where possible, but otherwise to leave such issues to others. Neither Swami seems to have ever articulated what they must have considered, that the location of Sri Kailas at the centre of the Ganges sources made it a strong candidate for identification with the Kailas mountain of the Epic and Puranic literature. How they reconciled their knowledge of the two Kailas mountains is nowhere recorded, but must be located both in their own faith and in the wider cultural characteristic layering of sacred knowledge in which overlapping sacralities need not be reconciled. Before we consider further the identification of Sri Kailas with the ‘original’ Indic Kailas, we shall turn to an anthropological analysis of the sources of authority among such individuals as Swami Pranavananda in the modern era, which may well reflect historical realities.

The Construction of Authority

The authority of Swamis Tapovan and Pranavananda to define sacred land- scape was accepted by the broader Hindu community. To better understand their authority and constructions we must focus on that community.50 Despite

48 His patrons included federal and state government ministers, industrialists, academics, and local aristocrats; see Pranavananda (1984: xx–xxiii, including photo section). 49 Swami Pranavananda stayed at Gangotri for two years with Swami Tapovan; interview with Swami Sundaranand, August 2008. 50 Much of this section draws on my own fieldwork in the western Himalayas and other sacred sites of India over the last four decades. the epic prototype? 239 the rhetoric of renunciation studies have shown that renunciates are not di- vorced from wider society and indeed that they form their own society.51 They live and act as part of a small though highly influential world of regional or Pan- Indic renouncers, a world in which the participants generally know each other personally or by reputation. For the ideal Śaivite renunciate, who has left behind the Devibhumi, the lower agricultural realms where the mother-goddesses rule, the landscape in the upper Himalayas is the divine vision and reality of Śivabhumi. He (or she) has entered the realm of their deity and as Śaivite renunciates mimicking that deity they are at home, and act and are expected to act within the parame- ters of behaviour attributed to a deity. As they partake of the Śaivite cosmos they do not require temples or more precisely—for temples are focal gather- ing points that offer warmth and shelter—they do not require the images of the deities which are the focus of common worship by householders visiting temples. The renunciates are trained to visualise the precise appearance and physical attributes of the deity (in this case Śiva in some form), and to read the landscape as permeated with his presence. Pilgrims acknowledge the renunci- ates as divine in the same manner that they do the images, bowing before them and making offerings to them. Therefore, in this ideal theoretical vision, the renunciate oscillates between being Śiva and being one of the fully-realised beings described in traditional texts as dwelling with the deity in Śivabhumi. Those texts describe certain fea- tures of the landscape and if, as we have seen, they do not collectively provide a consistent description, the particular texts that have informed the individ- ual renunciate’s vision will have prepared them to expect certain features in the landscape; the Kailas mountain on which dwells Śiva or the hermitage of Kubera near Badri, for example. The renunciate entering Śivabhumi becomes a part of the visionary landscape inaccessible to the less spiritually heroic figure. The modern renunciate movement has derived much of its inspiration from the upper Ganges region.52 A number of renunciates who practiced there have become prominent not only at a Pan-Indian level, but in the West, although others have remained obscure. While they may be discussed as a category, it is difficult to discuss them collectively in appearance, practice or belief. Some

51 See for example, the accounts of their society in Hausner (2007); Pinch (2006); van der Veer (1988). 52 One thinks of renouncers such as Sri 108 Swami Kailashanand-ji Maharaj who founded the Kailashanand Trust in 1954–1955, and in recent years the renowned Om Giri, now Om Giri Maharaj, long resident at Gaumukh and now Gangnani. 240 chapter 10 have cut their hair, others proudly retain it. Some are celibate, some are not, some have followed a single guru, others have sought teachings from many. Although non-sectarianism is an aspect of their modernist promotion, sec- tarian affiliation is undoubtedly a strong part of the individual renunciate’s identity. Yet in practice sectarian lineages both intersect and combine. Swami Jnananandagiri for example, whose name identifies him as a member of the Giri order of Dasnami sannyasis, drew on teachings deriving from Swamis of the Sarasvati order. This is not only an individual process. The Kailash math founded by Dhanaraj Giri and later headed by his disciple Govinandagiri,53 is today run by Swamis of the Sarasvati order.54 Similarly the Giris, who are par- ticularly prominent in the region today along with renouncers of the Naga sect (which is not directly connected to the local understanding of Naga deities), include a number of individuals such as Maharaj Om Naga Giri, who is a mem- ber of both sects. It becomes clear that the strict injunctions and complex divisions in Brah- manical texts are—and surely always were—idealised versions of reality. In practice authorities and spokesmen of sectarian groups do not speak for or gov- ern the lives of members who practice on the frontier of the sacred rather than at the central institutions of that order.55 In reality, a fluidity exists that is not immediately apparent to those whose research is based on texts and/or infor- mants at the central institutions of ascetic orders,56 but which is clear from prominent renouncer’s biographies and from fieldwork among them. Despite the dominant presentation of their traditional aspect the more prominent of those in this region at least, emerge from a modernist back- ground. The individuals we have highlighted were all born into the Brahman caste, were well-educated, not only in Indian traditions but to some extent at least in English and were employed in modernist professions such as journal- ism prior to taking sannyasi. What is particularly notable is that none of those whose backgrounds are easily discernible have risen through traditional San- skritic education and ashram or math-based learning from early childhood. Indeed, in each case known even their initiation did not follow the textual norms maintained by central institutions. There is, therefore, a paradoxical relationship between the renouncers in this location and the central institutions of Hinduism. Many of the institu-

53 Abhedananda (1987: 137). 54 Tapovan (2003: ix). 55 This seems characteristic of their frontier location, with the core tenants of the centre weakened at the periphery. 56 For an exception, see however the important work of van der Veer (1988: 174). the epic prototype? 241 tions are founded by renouncers (such as the Kailash ashram), yet the most charismatic and successful spiritual agents do not emerge from those insti- tutions (which produce monastic and temple administrators and functionar- ies). In any case the renunciate community actually maintains autonomy from such centres, acknowledging their validity but defending their own authority to act outside the institutional strictures in such matters as for example, accept- ing disciples or obtaining spiritual revelation outside of established channels. That freedom enables them to found their own orders; the system is thus self- replicating. Paradoxically, what seems to elevate the status of renouncers is not their having passed through the normal institutionalised Brahmanical learning pro- cess, but factors such as personal charisma, recognised achievements such as difficult journeys to famous tirthas,57 and political acumen—the ability to attract powerful patronage. Renouncers with a modern education are now bet- ter equipped to appeal to patronage from the educated classes than those from a traditional background, being better able to frame their teachings in an idiom of scientific modernity. (Being born a Brahman is probably not essential to that appeal today, but still carries an implication of traditional authority and dharma that reassures patrons of their genuine spiritual calling.) At a local level, acknowledgement of local deities, a facility for empathy with local people and skill in teaching in the widest sense, all enhance their reputation. But the actual demonstration of their spiritual strength through such endeavours as remaining at Gangotri throughout the winter, or traversing the regions above Gaumukh, play the major part in the development of their status. In fulfilling the ancient ideal of a Himalayan rsi, they increasingly gain recognition for spiritual achievement, a recognition that then outgrows the local world. Before attracting Pan-Indian patrons, the renouncer must gain followers at a local level in the form of non-elite householders who refer to him for practical and spiritual advice, and also those from a wider area—usually Brahmans— who are seeking a guru and are thus potential initiates into the renouncer’s spir- itual lineage. Yet these ties may not be formalised. Swami Tapovan for example,

57 It is interesting that extreme acts of self-mortification by renouncers, while attracting considerable attention in the West from the time of the classical Greek commentators onwards, does not seem to be an automatic path to Pan-Indian renown. While adding to status at a local level, it may be that this does not fit with the modern presentation of Hinduism as a rational and scientific faith, yet it does not seem to have been a path to higher recognition even in the pre-colonial period. The issue requires more analysis. 242 chapter 10 did not actually initiate anyone into a lineage or sect, but several prominent renouncers consider(ed) him their guru. The emphasis in modern renouncer’s spiritual biographies, both written and oral, is always on guru rather than sect.58 Yet hallowed as the guru-shishya (relationship) is and however idealised it may be, it is not necessarily a simple one. A renouncer initiated by one guru may subsequently fall out with him and follow another, while those seeking initiation may be refused, or after a period of discipleship, leave without initiation. Others are initiated by one guru but with his blessing study under another or with many others. This may lead to them seeing a later teacher as their pre-eminent guru, as in the case of Swami Sundaranand, initiated by Swami Dayananda Giri but primarily loyal to Swami Tapovan. Initiation may even be claimed directly from a deity,59 and renunciates may be accepted as such within the community without formal initiation, sectarian membership, or even a guru if they possess the necessary personal charisma and/or are manifestly genuine practitioners.60 Renunciate autonomy is also manifest in their practice, which allows a wide body of specialities to be developed. Yoga, bathing, fasting, a vow of service, silence or other penance, consumption of cannabis, the chanting of mantras, the making of mudras, and so on, may or may not be included in any particular renunciate’s practice. Initiation by a guru involves the bestowing of a spiritual name and a personal mantra, after that teaching is tailored to the individual within the broad understanding of the lineage or sect. The individual may also seek their own path and learn from fellow renouncers, in contrast to those who undertake a formal course of instruction at an ashram in places such as Haridwar or Rishikesh. An individual’s mastery of a particular practice is recognised by the wider renunciate community,61 and may lead to him being known by his practice—‘Mudra Baba’, for example. The question of whether they are Tantric is almost irrelevant in the field. Tantric practices are recognised

58 Here my findings differ from the received understanding of the importance of initiation into a particular spiritual order as dictated by textual and ashram authorities. 59 One such renunciate I encountered in Haridwar claimed to have been instructed to take sannyasi by his deity, who appeared to him in a dream. He gave the term “manasa- sannyasi” (‘mind sannyasi’) to describe himself and others of that ilk. Hausner (2007: n. 12 211) notes the existence of the term vaimukh for renunciates without a guru; Mallinson (1993: 53) records the use of the broad term ‘siddhas’ for renunciates who have had no formal initiation, but are accepted by others. 60 See the example in Tripathi (2007: 250–252). 61 My own pursuit of visiting each of the Kailas mountains for example, was regarded by the renunciate community as within (admirable) Giri practice. the epic prototype? 243 as one body of knowledge and practice that can be adopted but there are many such bodies and virtually no restrictions on practices that might be followed, briefly or for longer periods. The academic urge to classify particular practices as Tantric (or otherwise) is not matched by renunciates, at least in this region. Yet if sectarian affiliation, guru lineage, and ritual practice are more fluid than is generally assumed, broad divisions do shape the renouncer’s activities. Certain locations are almost entirely Śaivite, others almost entirely Vaisnavite, and the renunciate’s external markings—both traditional and modern—clear- ly indicate their spiritual persuasion. Both sect and guru are quickly revealed on meeting other renunciates and may be grounds to exclude them—aghoris, for example, renowned for their cemetery practices, may not be welcome every- where. But again this seems to be affected by personal charisma, a Sita-Ram follower, for example, may be welcomed and given a prominent place at the dhuni (fire) of a Śaivite site if his personality is appropriate. The renouncer’s embracing or otherwise of cannabis and other lifestyles may also be an impor- tant factor here, with its consumption crossing sectarian boundaries and cre- ating its own communities. The modern renunciates of this region, or at least the elite among them, are therefore, autonomous agents operating in a sacred landscape in which they partake of divinity through their identification with the deity, a Tantric mode that has permeated their practice. They are endowed with and claim the charismatic and traditional authority to define, to reveal, and to pronounce. In the tradition of the rsis their visions are passed on and may, as in the case of the Gangotri mahatmya, become recorded as texts. Their audience is an educated and often sophisticated section of the Indian world which includes both the elite and the intellectual classes, who sustain, promote and encourage the dominant knowledge of that culture. Thus a modern renunciate such as Swami Tapovan had the authority to provide both the toponyms of scientific geography and the modern reading of the landscape that sought to locate the Vedic world in the contemporary, as well as the authority to compose the mahatmya with its spiritual explanations of the landscape. Through his charismatic authority he gained the audience to receive his teachings and enshrine them in Indian and world beliefs.

On Sri Kailas as the ‘Original’ Indic Kailas

Descriptions of Kailas in Epic and other early literature are impossible to con- vincingly reconcile with geographical realities because the Kailas of those texts was a heavenly, not an actual, place. Despite indications that the descriptions 244 chapter 10 had roots in some knowledge of Himalayan (or Pamir) realities, it is highly unlikely that the various accounts all relate to the same roots. Yet the Epic accounts do imply that their Kailas mountain was close to the source of the Ganges, and a strong case can be made for that ‘original’ mountain, in as much as it can be said to exist, being Sri Kailas and for Swami Tapovan’s attribution of its name reflecting some version of historical realities. Certainly the extraor- dinarily distinct phallic appearance of nearby Sivling peak makes it another strong candidate for a Kailas model, while entirely different historical trajec- tories might be considered to suggest that Muzh Tagh Ata was the mountain referred to,62 but there is much in favour of Sri Kailas. If we depict it as the centre of a mandala (something apparently never specifically done in Indic tradition in the Kailas context), Sri Kailas is sur- rounded by the Ganges’ head-waters and sites like Gangotri, Badrinath, and Tholing. If the names Swami Tapovan gave are traditional (which is Swami Sun- daranand’s understanding), then the sacred geography of the Sri Kailas region is of great significance. From Gangotri you follow the Ganges southeast towards Gaumukh, or the ‘Cow’s Mouth’ glacier, the actual source of at least this branch of the great river. But while river sources may be determined by scientists, they are not always obvious to the viewer. In 2008, the retreat of the Gaumukh glacier meant that the main body of water was not emerging from that glacier,63 (see photos 17, 18). Instead there was another water source emerging from under the north side of the glacier and that flow could be traced back a hundred yards or so under the glacier to

62 Muzh Tagh Ata massif is situated in the far west of China near Tashkurgan and Tajikstan. Staal (2003) points out that Muzh Tagh Ata and the Oxus and Tarim/Yarkand rivers would have been known to the early Indo-Europeans who migrated down from the Oxus to the Indus. Certainly the persistence of the Tarim and Oxus in the classical 4-river schema of the Abdhidharma literature etc., suggests conceptual roots in that complex. Further evidence for this is that Muzh Tagh Ata was a source of soma (see n. 48 36, this work). We might also note the existence of a pass in the vicinity of Muzh Tagh Ata named Subash (Turkic: ‘Origin of the Waters’) pass, and that the Tiznap river that runs through Tashkurgan is a tributary of the Tarim, which might explain how that river is referenced in early texts as one of the four/five rivers of the cosmic mountain. The Bönpo’s understanding of their religion as deriving from Tazig and the Ölmolungring construction might also be considered in this regard. (We might also consider whether this schema was inherited from the Kushan or Zhang-Zhung polities, as their territorial horizons reached those rivers.) The point is that there is no single source for the Kailas concept. 63 Glacial movement may be rapid; the Gangotri glacier retreated 600mtrs (14.6 mtrs per annum) between 1935 and 1976 (figures from Geological Survey of India). This rate has now greatly increased, see Nand and Kumar (1989: 23). the epic prototype? 245 a stream coming from a glacier to the north. Standing at that point looking up the Gaumukh glacier, there is a Mount Meru—the other name so significant in classical Indic cosmology—above a glacier on your right marking the route to reach Tapovan, the beautiful glacial valley dominated by Shivling. But on the left, to the northeast, following that river flow from the glacier to the north, is the Raktavarna Bamak—the ‘Blood-red [coloured] glacier’,64 which leads up to the Sri Kailas glacier, which runs northwest. From there it is possible to cross onto another glacier that descends to the Jadh Ganga course (that river course followed, incidentally, by Heinrich ‘Seven Years in Tibet’ Harrer to cross to Tibet over the Thaga la65). The most southerly Jadh Ganga route can be used to meet the route from Badrinath to cross the Mana-la. Sri Kailas is thus located at a point between two main glacier-fed Ganges tributary sources, and is roughly equidistant from the third main tributary, the Sarasvati-Alakananda. More than any other mountain it can be seen as the peak that manifests the Ganges, the site where the heavenly river descends to earth. The movement of glaciers means we cannot be certain of the location in previous millenniums of Gaumukh glacier (which is currently retreating), but the 2008 hydrology is unlikely to have been unique. Early pilgrims, with no means of ascertaining which stream had the greater volume of water, could thus have envisaged Sri Kailas as the mountain source of the great river of north Indian civilisation. Given the conclusions of Gruenendahl and others already noted and consid- ering the importance of the Ganges river to the Indic world of the Epic period, allied to the very real possibility that anyone tracing the Ganges to its source might well have found themselves at the base of Sri Kailas as my own field- work suggests, Sri Kailas does seem the likeliest candidate for the Epic Kailas. The sacred geography of the region, with Mount Meru on the other side of the meeting place of the Raktavarna and Gaumukh glaciers nearby and Sri Kailas situated between rivers feeding the Ganges, is evidence for this. Even the northward turn of the Raktavarna glacial source may be considered spiritu- ally significant, as we have seen in the case of Ganges at Varanasi and the Indus near its source.

64 The Raktavana gets its name from the red soil in the waters. Rakta, meaning ‘red’, also car- ries the Tantric meaning of menstrual blood (of the Goddess); Gray (2007: 160), suggesting it as of alchemical interest. 65 For a map of his route see Brauen (2002: 24). I am advised by Isrun Engelhardt (University of Bonn), personal communication April 2009, that Harrer’s diary does not mention Sri Kailas. 246 chapter 10

If we identify the various Ganges sources as the rivers of the classic cos- mological model, we are left with the problem of the lake in that model, for the Kailas mythology emphasises the lake as much as the mountain. But that mythology speaks of one lake, never of two, which as noted is a major prob- lem with the reconciliation of Kailas-Manasarovar and the early Indic sources. There are no lakes of note at the Bhaigirathi or Jadh Ganga sources, but there are several in the vicinity of what seems to be the earliest known source of the Ganges, Badrinath. Two streams that feed into the Alakananda have lake sources. There is Deo Tal on the Mana la, which is the source of the Sarasvati, and there is the Arwa Tal at the Arwa river source. The former lies on the route to Tibet, the latter on the difficult traverse from Gangotri to Badrinath. But nei- ther lake or feeder stream seems of great sanctity. There is however, another lake, Satopanth Tal, at the traditional source of the Alakananda branch of the Ganges. A triangular-shaped lake, around 500 yards (450 metres) across and 1200 yards in circumference, Satopanth Tal is some 15 miles (22 kms) from Badrinath via Mana, though only around 8 miles due west as the crow flies. Situated at a height of 4,402metres, Satopanth (Skrt: lit; ‘the path to the truth’) Tal lies above the Vasudhara falls, the traditional source of the Alakananda. It is thus the lake at a traditional Ganges source and it is sacred to the people of Mana,66 who wash their dead there before cremation and later scatter the ashes of the deceased there. Satopanth Tal is often frozen over.67 But the Swiss geologist Augusto Gansser bathed there in 1936, and records that;

the sacred Lake of Satopanth … [lies] in front of a spur of rock … On the shore, about 20 feet below the flood-water mark, human bones and rags of clothing can be seen, relics of the dead who had been committed here to the sacred waters. On a big boulder at the acute end of the triangular lake are the remains of faded red-and-white prayer flags; and under the shelter of this rock the pilgrims have built a … night shelter … some of them spend

66 The people of Mana, like those of Kuti, claim to be Hindus who came down from Tibet. Discover India magazine (Issue 54: 1998) has some notes and illustrations of Mana but I have not located academic studies. See however, Jerarth (2001: 121), who records the name of the local deity of Mana as “Gantakaran”. See however, Sherring (1974: 358–360), who describes the wider appeal of this ‘Bell God’.Also see, for a recent account of a silent ascetic who lived at Satopanth in a small concrete hut, wearing only a dhoti; Chakrabarti (2003: 167–172). 67 See the photo in Sundaranand (2001: 219). the epic prototype? 247

weeks up here in summer drinking the water and filling bottles with it to take home … the water was swarming with tiny, fiery-red plankton crustaceans. Ducks also come hither on a visit.68

Written sources concerning Satopanth Tal are overwhelmingly Sanskritic and Vedic references to the mountains as the home of departed souls problematises classifying the funereal rituals held there as entirely a local tradition. It is unclear if the local population identify it as a Naga lake. Yet as a site understood to have been passed by the Pandavas en route to heaven and as a lake at the source of the historically most prominent branch of the Ganges, it is a candidate for ancient Sanskritic sacrality and for having been a model for at least some aspects of the sacred lake of the Epics, Puranas and Pali texts. In support of this thesis is the fact that the (Sanskrit) lake toponym Manasarovar might be rendered as ‘Lake of Mana’ and thus applied to this lake, enabling the concept of the Ganges taking its source in Manasarovar to be recognised here. Ultimately, the early Indic Kailas was an imagined place, a heaven, a poetic vision, and later a derivative cosmological centre. But if the early textual ac- counts of this site reflected any actual geography, the earthly mountain to which they referred was most probably Sri Kailas and the earthly lake Sato- panth Tal.

68 Heim & Gansser (2004: 196, also see photo 212 between 210–211). Illustrations

figure 1 Tibet’s Tise—Mount Kailas south face courtesy of toni huber

© koninklijke brill nv, leiden, 2015 | doi: 10.1163/9789004306189_013 illustrations 249

figure 2 Kaplas Kailas: a glimpse of the peak above the insurgent’s zone of contestation

figure 3 Manimahesh Kailas, the view from the lake 250 illustrations

figure 4 The Manimahesh massif, north face

figure 5 Kinnaur (Kinner) Kailas range courtesy of rafal beszterda illustrations 251

figure 6 Kinnaur peak (centre); the small peak immediately beside it is known as “Śiva’s lingam” courtesy of rafal beszterda

figure 7 Adhi (or ‘Baba’ / ‘Chhota’) Kailas; viewed from Joylingkong 252 illustrations

figure 8 Sri Kailas and surrounds googlearth. landsat. digital globe. 31° 01’20.32” n, 79° 10’35.95 e. july 11.2014. accessed january 16, 2015. illustrations 253

figure 9 Sri Kailas. From a photograph by Swami Sundaranand courtesy of swami sundaranand 254 illustrations

figure 10 Muz Tagh Ata (near Tashkurgan in Sinjiang, prc) is another possible Kailas prototype illustrations 255

figure 11 The survival of local traditions: pole shrine (darcho) outside Kuti, near the Adhi Kailas 256 illustrations

figure 12 Naga deity Vasukinag and his faithful companion Jitmanji: Bhadrawah Mandir (Doha dist: Kashmir) illustrations 257

figure 13 Sanskritisation: modern Hindu symbols overlaid onto the local protective deity of Kuti (Kumaon district) 258 illustrations

figure 14 “Śiva & Parvati at Kailas”: early 20th century. The absence of any link to the Ganges and the depiction of two mountains implies a local Kailas (Manimahesh?) is intended. courtesy of chamba state museum, hp illustrations 259

figure 15 Depiction of the Tibetan Buddhist Tise pilgrimage in traditional form at Tragyam Gompa, Dolpo credit: edward h. worcester 260 illustrations

figure 16 Jigten Gönpo, The Drigung Kagyu hierarch who initiated that sect’s movement into western Tibet (see Heller 2005) private collection, zürich illustrations 261

figure 17 The Gangotri Glacier at Gaumukh, source of the Bhagirathi branch of the Ganges. Note the major stream on the left of the picture, which emerges from the Raktavana (bamak) Glacier. 262 illustrations

figure 18 The Raktavana (bamak) glacier face and stream which joins the Bhagirathi Ganga at Gaumukh (see Photo 17) illustrations 263

figure 19 Satopanth Tal, the lake at the traditional source of the Alakananda branch of the Ganges. The Manasarovar prototype? 264 illustrations

figure 20 Charles Sherring’s Western Tibet and the British Borderlands, which constructed the modern understanding of Kailas credit: university of california digital library illustrations 265

figure 21 Swami Jnanananda, renunciate, scientist, and guru of Swami Pranavananda and Swami Tapovan 266 illustrations

figure 22 Swami Pranavananda, renunciate, scientist, and author of the classic guidebook to Kailas-Manasarovar illustrations 267

figure 23 Swami Tapovan, renunciate, Kailas pilgrim, and Gangotri savant. Painting from a photograph taken by Swami Sundaranand the day before the samadhi of Swami Tapovan. courtesy of swami sundaranand 268 illustrations

figure 24 Swami Sundaranand at his Gangotri hut, originally established by his guru Swami Tapovan, whose portrait is visible to the left of Swami Sunderanand illustrations 269

figure 25 Lama Anagorika Govinda, German Buddhist and New Age apostle of a Universal Kailas

section 3 Tibetan Histories

chapter 11 Tibet’s Tise

Introduction

The modern understanding of Kailas-Manasarovar as a ‘World mountain’ is largely shaped by Indic perspectives and owes little to Tibetan understandings of the site. While Tibetan ritual practices enhance the modern experience of its sanctity the perspectives these performances represent are still poorly reflected in the modern construction. In this and the following two chapters we situate the site in Tibetan history, considering the disparate perspectives presented by the Buddhist and Bön faiths which both lay claim to particular representations of the mountain-lake complex. We are well-informed as to the nature of Tibetan pilgrimage, particularly to sacred mountains. The great Italian Tibetological pioneer, Giuseppi Tucci, advanced the earliest model in which a tribe dwelling around a mountain came to identify it as their protector-deity.1 The mountain provided the necessities of life for the tribe and its herds and in return was worshipped through annual or biannual rituals carried out on the mountain-side. The mountain and/or mountain deity (for the two identities may subtly merge), came to be under- stood as being in a relationship with the tribal ancestors and leaders (as we saw was also the case in Korea). Thus the identity of tribe, rulers, and moun- tain were linked and the mountain served as a tribal identity marker vis-à-vis neighbouring tribes, each of which was in a similar sacred association with a particular mountain. Although Tucci’s ideas survive more recent studies have refined this model and brought out the extensive variations in the world of Tibetan mountain deities, which exist in diverse contexts and ideologies.2 There is neither a single

1 On whose scholastic legacy,see de Castro & Templeman(2015);for Tucci on sacred mountains, see in particular Tucci (1980: esp., 193, 202, 219). 2 There is now an impressive body of literature concerning pilgrimage and sacred space in Tibet, to which this present study is greatly indebted. See the bibliography for details. To briefly note only major edited collections and monographs, see Blondeau (1998); Blondeau & Steinkellner (1996); Buffetrille (2000); Ekvall and Downs (1987); Gutschow, et al., (2003); Huber (1999); (1999b); and (2008); Karmay (1998); Kind (2012); Macdonald (1997); McKay (1998); Ramble (2008); Tautscher (2007). For wider theoretical and contextual considerations, see in particular Huber (1999a) and a convenient summary of her thesis by Buffetrille (1998).

© koninklijke brill nv, leiden, 2015 | doi: 10.1163/9789004306189_014 274 chapter 11 model of such deities nor even a specific Tibetan term for ‘mountain god’. The most common term—yullha—actually indicates a ‘village’ or ‘territorial’ deity, but in central Tibetan usage does frequently refer to a mountain deity (and the term is used here in that sense). The yullha is usually depicted as a fearsome warrior, to whom prayers might be offered on the following lines;

In going to war, be our general. In trading, be our greatest trader. Whilst robbing, be our head of robbers. Lead the army to the land of the enemies, destroy them and achieve victory.3

A distinction can be drawn between sacred mountains presided over by such yullha (usually understood to have been converted to Buddhism), and ‘gnas ri’ mountains, where the presiding deity is from the Buddhist (or Bön4) pantheon. This distinction is in some senses one of stages in a process. As Toni Huber states;

One can only consider all Tibet’s cult mountains in terms of a gradient along which they vary, from the “model” néri [gnas ri] at one extreme to the “typical” local cult mountain at the other, with most falling some- where in between.5

Tise (as the Tibetans call the mountain now identified in India as Kailas), is at the “extreme” end of this gradient in that its presiding deity is entirely Buddhist. No certain trace remains of those elements that Samten Karmay and Katia Buffetrille isolated as defining a yullha mountain, notably a close relationship with local tribal leadership. What is prominent at Tise are the elements char- acteristic of a Buddhist gnas ri mountain; ritual circumambulation, a historical ‘opener’ of the pilgrimage, and the “ritual appropriation of space, in which writ- ten sources serve an important function.”6 This indicates that either Tise has been so thoroughly overlaid by Buddhist concepts that it has been completely

3 Invocation to the mountain deity Gangmar (sgangdmar), cited in Diemberger (1998: 46). Very similar sentiments or formulas are cited by Charles Ramble (1998: 128) from a recitation by the lha bon (priests) of Khyenga village; “[i]f we go trading, make us the foremost traders, If we go to war, make us the generals …” 4 Huber (1999: 23) identifies “the cult of both Buddhist and Bönpo néri [gnas ri] as conforming to exactly the same cultural pattern.” 5 For a discussion on yullha and néri mountains and the complex range of forms these sites have taken in space and time, see Huber (1999: 21–35). 6 Karmay (1994: 115); Buffetrille (1998: 20–25; quotation from 21). tibet’s tise 275 detached from a local context and all trace of its earlier sacrality eliminated, or more probably that the mountain only became sacralised by Buddhism. As can be seen in Chapter 14, there are a large number of deities which have been identified in disparate Indic and Tibetan texts in association with Tise. But we cannot determine a single ‘classic’yullha with a fixed identity, even if we surmise that ‘Tise’ is the name of that deity.7 The earliest references to regional sacrality concern not the mountain but the lake, Mapham (Manasarovar). While those accounts are heavily mythologised, the identity of the presiding deity of the lake seems largely consistent in texts dating back to the middle of the first millennium ce., if not earlier. The lake is said to be ruled over by the serpent deity Madros, a Klu (Skrt; Naga), whose most common name in Pali, Sanskrit or Tibetan equates to ‘Not Warm’. (As the residence of a serpent deity, Mapham (mtsho: Tib; ‘lake’) may be typologically associated with Western Himalayan Cultural Complex Naga lakes.8) But there is no such consistent tradition concerning the mountain and the number of deities associated with it seems to indicate a more complex history. The want of any early mention of a Tise ‘mountain deity’ in Tibetan literature suggests that the spiritual significance of the mountain was not of the antiquity or the stature of the lake. The multiplicity of gods with which Tise was associated in literature from the second millennium ce onwards also suggests that the identification of its presiding deity was a contested process, not an organic development. The Buddhist take-over of Tibet was a process that operated on numerous levels, with a fundamental aspect being the subjugation (’dul ba) of the land- scape and its autochthonous deities.9 It was this ritual and visionary process that ultimately transformed Tibet into a Buddhist realm, with the understand- ing that the deities themselves and the landscape in which they manifested had become subject to Buddhism. Yet despite tales of Madros embracing the teachings of World-religions, the predominant Tibetan Buddhist mythology of Tise was, as will be seen, a magical triumph ending in the expulsion of the Bön

7 A point noted by Buffetrille (1998: 27). 8 Further indication that the Tise/Mapham region may have been associated with the Western Himalayan Cultural Complex is the fact that Ladakh, to the west also has no sacred moun- tains, or pilgrimages to them; although mountains are home to renunicates. Yullha there seem more on the local territorial deity model of the Western Himalayas. They are, for example, gen- erally seen as coming from elsewhere rather than being linked to the foundation of territory or clan. (This may be a significant division). See Dollfus (1996). With exceptions, the same is true of Zanskar, see Riaboff (1996). 9 On which see Ramble (1999: 15); also see Huber (1997: 246). 276 chapter 11 practitioners there, along with a stock myth involving the subjugation not of a local deity, but a Brahmanical god. This again suggests the mountain became sacred in a period of sectarian competition. That competition involved definitions of territory that suggest a cultural understanding of particular mountains as protectors and markers of political and cultural frontiers. It is notable that the principal gnas ri mountains of the Tibetan plateau are situated on the external borders of the Yarlung Tibetan state and the Kailas mountains of India might also be located in that context. As Huber has noted, “the definition of larger scale (i.e., nonlocal) forms of ‘Tibetan’ identity was certainly an active feature of ritual life at borderland néri sites up until 1959”,10 while there are also Chinese and Turko-Mongol parallels to such understandings. We shall however, reserve that discussion for the following chapter.

The Early Period

Apart from inscriptions,11 most of the earliest Tibetan sources consist of mate- rial found in the caves at the central Asian Buddhist centre of Dunhuang. Most of that material, however, dates to shortly before the caves were sealed around the end of the first millennium, only to be revealed again in the European colo- nial period. A study of the Dunhuang Tantric material in the Stein collection for example, “did not find any manuscript that could be firmly dated to before the mid-tenth century.”12 We have in any case no proper sources for central Tibetan history before the early 7th century ce., and the early history of the Ngari (mNga’ ris) region within which Tise and Mapham are located is almost entirely obscure.13 Ngari (which very broadly equates to western Tibet), only

10 Huber (1999: 24). 11 On which see Richardson (1985). 12 Dalton and van Schaik (2006: xxi; for a succinct account of the background of this material, see xi–xv). 13 There is an intriguing but inexplicable reference in Emile Schlagentweit’s Buddhism in Tibet (1863: 63), stating that a ruined Buddhist monastery dating to around 137bce existed “on the slopes of the Kailas range.” Schlagentweit’s source for this is given as another pioneering work, Christian Lassen’s four volume Indische Alterthumskunde (1849: vol. ii, 1072). The citation is however, incorrect, and I am unable to locate any mention of the monastery in Lassen’s work. (My thanks are due to Prof. Werner Menski (soas) for assistance with the German language text.) Schlagentweit’s claim was repeated in two other prominent works which did not give the source; Blavatsky (1888: vol. 1; 35.) and Waddell (1895: 19). If the ruin existed, we must tibet’s tise 277 emerges into recorded history when its independent existence as a principality called Zhang-zhung came to an end in the mid-7th century. The historically shadowy Zhang-zhung became, as we shall see, something of a blank canvas for future imaginings of history and identity. Most notably, in the context of the emergence of a systemised Bön religion in the 10th century ce., it came to be remembered as the ‘heartland’ of Bön, with Tise remembered as what Tucci called the “soul mountain” (lha ri)14 of that realm. But literature advancing that understanding only developed many centuries later. Tibet had emerged as a Central Asian power from origins in a dynasty rul- ing the Yarlung valley and that expanding principality conquered Zhang-zhung around 644, killing the Zhang-zhung king and annexing the region.15 That a distinct local identity remained is indicated by a failed rebellion in 677,16 and by the promptness with which the region re-emerged as a centre following the collapse of the Yarlung dynasty after the assassination of its last king in 842. But Ngari’s 10th–11th century re-emergence in the form of several histor- ically interwoven principalities, notably Guge-Purang (Purang being the dis- trict encompassing Tise/Mapham), was under a Buddhist dynasty claiming Yarlung royal antecedents; the old Zhang-zhung ruling aristocracy had been vanquished. There is no evidence that Buddhism existed in any formal institutional sense in Zhang-zhung.17 But the Yarlung take-over meant that the ‘First Dissemina-

rewrite much of the early history of the region, presumably linking it closely to the Kushan empire, but I found no mention of any such ruin in the writings of any of those who have travelled widely in the region. Emile Schlagentweit relied heavily on information gathered by his brothers, who had travelled in Western Tibet in 1855, and their papers in the Royal Geographical Society might throw light on the reference. 14 Tucci (1980: 219). See however, Walter (2009: 142–153) who states that bla, which Tucci translates as “soul”, did not have this meaning in the earliest texts, but is there a political term meaning ‘government’ or ‘governing power’. (But the application of the term bla ri to Tise may be late; i.e.: post-phyidar). For a nuanced discussion of the term in this context, see Davidson (2005: 220–223). 15 Vitali (1996: n. 319 221). Any scholar working in this area owes an enormous debt to Vitali for this extraordinary work. Kingdoms …, may conveniently be read in conjunction with Vitali (2003: 53–89). Also see, Richardson (1988: 30). 16 On which see Tucci (1956: 105–106). 17 Vitali (1996: 276) notes the earliest Buddhist structure in the region, “[t]he ancient temple of dPal.rgyas … located in Zhang.zhung khri.sde. Its foundation dates back to the time of Srong.btsan sgam.po, when it was built to prevent mtsho Ma.pham, from overflowing (and flooding the great sKyin.Thang river in Tibet).” No other information about this structure, or its precise location, is known; email communication, Roberto Vitali, July 2007. 278 chapter 11 tion of Buddhism’ (snga dar) in Tibet was at least proclaimed in the newly conquered region in the late 8th century.18 Any immediate impact was lim- ited, but there is (much later) evidence for Buddhist activities in Ngari by the mid-9th century,19 and Guge-Purang’s new rulers were famously instrumental in initiating the phyidar (phyi dar), or ‘Later Dissemination of Buddhism’, on the Tibetan plateau in the late 10th and early 11th century. It was the phyidar that fully established the plateau as a Buddhist realm. We have seen that within canonical Indic Buddhism there were bodies of knowledge that included a heavenly mountain named Kelasa, and a heavenly lake, or lakes, that went under various names. But the late Mahayana-Tantric forms of Buddhism that took root in Tibet and became fully established in the phyidar produced very different understandings of the regional sacred geogra- phy. The early Tibetan sources from Dunhuang confirm that the Indic cosmo- logical understanding of a central World-mountain had passed to Tibet dur- ing the first dissemination of Buddhism,20 if not earlier. That World-mountain was not however, in any way related to or identified with Tise or any other earthly mountain. Indeed, in the early material there are scarcely any refer- ences to individuals or processes at Tise or Mapham, or associations of the region with any particular sacrality. In those sources Ngari is only a periph- eral region referred to largely in the context of Yarlung expansion and had the complex been of great regional significance (as the ‘soul mountain’ of the conquered state, for example), we might expect some mention of it in that con- text. One explanation for its obscurity is suggested by Michael Walter. While mountains are referred to in metaphors, Walter observes that, “we have not a single text relating to a ‘mountain cult’ or its rituals … in Old Tibetan [i.e.; pre-phyidar] literature.”21 Walter thus problematises established assumptions concerning the antiquity of Tibetan mountain cults, suggesting that these ideas may only have emerged into Tibetan culture in the late Yarlung and early phyidar periods.22

18 Vitali (1996: 165); and concerning the edict, see Richardson (1980: 62–73). 19 Vitali (1996: 164–169). 20 See for example, Dalton & van Schaik (2006: 124); also see Dietz (1988: 113–115); Walter (2009: 290). 21 Walter (2009: 230–231). We note that the periodisation of Kailas-Manasarovar given by Frederick Thomas does not contain historical elements of Tibetan belief prior to the 10th century; see Takeuchi, Quessel & Nagano (2011: 51). 22 Walter notes that despite the prominent mythology of the descent of the legendary first Yarlung king onto a mountain, that site did not become the centre of any state cult; Walter tibet’s tise 279

Certainly if we rely on Tibetan textual sources, the ‘ancient status’ of Tise takes on a mirage-like quality. In much later sources layers of beliefs are pro- jected back onto that ‘ancient’ Tibet, and academic enquiries assume, a priori, that the site was a sacred centre in that ancient world. But the Tibetan sources, like the Indic, fail to confirm any exalted status for Tise before the 2nd millen- nium ce. Yet so wide-spread among early Iranian, Turkic, Indic, Chinese and south- east Asian cultures were beliefs in mountains as embracing some sacral char- acter—be it funereal or heavenly—that the absence of such a belief in ancient Tibetan culture would, in the regional context, seem a major anomaly. Absence from the very limited sources that we have could simply imply that the sacrality of mountains was too obvious to need constant reiteration, and certainly the use of mountains as metaphors in the early sources suggests they had some culturally emotional resonance. As the early Tibetan sources represent the interests and perspectives of the Buddhist elites the absence of references to sacred mountains in those sources may only indicate the absence of a moun- tain central to their interests or identified with the emerging state. The plateau was inhabited by a variety of distinct ethno-liguistic groups at that time, and the Buddhist elites may, for a number of possible reasons—religious, political, and social—have separated themselves from the strata that held the moun- tains sacred.23 The handful of relevant sources do, however, reflect a culture in which the inhabitant’s relationship with the environment was mediated not through

(2009: 230–240, n. 41 277), on which also see, Kirkland (1982: 257–271). Kirkland notes the early introduction of the tradition of the cosmic mountain Ri-lab lhun-po (in thd, p. 81/85f.). There are numerous references to mountains in association with Tibetan royalty; see for example, McKeown/Stein (2010: 28, 73, 75–78, 143–148, 155, 157, 160–161, and especially 333). But many of Stein’s references are from Chinese sources or suggest a wider notion of territorial deities rather than specifically mountain deities, while none of the Yarlung dynasty inscriptions translated by Richardson refer to sacred mountains; see Richardson (1985). Further investigation is needed, and see the archaeological input of Aldenderfer (2011: esp., 31–32) who links domestic rdo ring to a wider mountain cult (but not to Bön). I am indebted to Toni Huber for the point that by that time the Chinese already had very strong and sophisticated Buddhist and Taoist mountain cults which were a focus of imperial power, and that the Yarlung dynasty Tibetans are known to have sought information concerning Mount Wu-tai shan (on which see Huber 2008: 46–47). Chinese input is thus worthy of further examination. 23 We might consider that the use of mountain metaphors for kingship in the early literature represents an implicit appeal to a section of the population holding to mountain sacrality. 280 chapter 11 mountains, but through lakes (such as Mapham), inhabited by Klu or other spirits. That suggests that the pre-Buddhist culture of the peoples of Ngari, or a significant element of them, involved concepts shared with the West- ern Himalayan Cultural Complex. That does not preclude the existence of other sacred elements in that realm, however, and the survival today of moun- tain worship traditions in those areas of Tibet least influenced by Buddhist authority does suggest their indigenous (or regional) origins, while their state- less character also lends credence to their being of considerable—even pre- Buddhist—antiquity. If such phenomena were the preserve of oral traditions then understandings may have been lost or be at most dimly reflected in (later) written accounts. Thus the absence of references to Tise in any systemised religious context in the early sources is not definitive evidence of an absence of belief in its sacrality. It may point to a wholly local or a factional understanding of its sanctity in that period.24 But that Tise occurs only as a metaphor or as a geographical locator is consistent with the conclusion of it being at the very most of local significance until it was ‘Buddhacised’25 during a period of assimilation and subjugation of local forces.26

24 Evidence of its local status is seen in Vitali (1996: n. 516 327–328) which gives the example of Tise and Mapham being used as metaphors for eternity in a treaty around the mid-11th century. The use of mountain or other deities of place in oaths is however, known else- where in the Himalayas (e.g.; Sikkim’s Khanchendzonga), and may be based on their use as metaphors for eternity rather than religious status. 25 On ‘Buddhacisation’, see Buffetrille (1998). 26 As, for example, in the reference to them in pt 1287/Ms.250 Pelliot (The Old Tibetan Chronicles), cited in Haarh (1969: 403). In this foundational account of the battle of the early King Gri gum his/a demon’s[?] body is cast into the Tise glacier—which I read as a metaphor for the ‘far horizons’. See the discussion and alternative translation of this (which does not, however, alter the point that this is a metaphor) in Cutler (1996: 79–82). As he and others have noted however, the tale derives from an existing myth. Also see Uray (2003: 256). In another version (cited in Haarh 1969: 343) the king’s body is thrown into a river, while the final redaction of the otc is dated by Stein to post-842 (i.e.: long after the Yarlung take-over of Zhang-zhung), McKeown/Stein (2010: 118). Richardson further problematises the legend by noting later traditions locating the event near Gyantse and 19th century Indian pandit Nain Singh’s finding of a similar story associated with Lake Dangra (Danr ra G.yu mtsho). Richardson concludes that a location in Kong-po is most likely, but that the myth is an “amalgam of hazy memories”; Richardson (2003: vol. i; 156–157). For recent translations and comment see, Hill (2006:92) and Zeisler (2011: 126–131). tibet’s tise 281

Early Renunciates in the Region

In later sources the Indian Tantric master Padmasambhava is credited with a primary role in establishing Buddhism in Tibet. There are tales of his visit- ing Tise and a reference from the immediate pre-phyidar period to the region attracting wandering religious practitioners. A Dunhuang text dated by Mathew Kapstein to probably 965 or 977, mentions an Indian Buddhist prac- titioner who travelled to Ngari around that time. There he, “practiced the pro- found rites of service and attainment at Mt Tise, and performed ablutions in Lake [Mapham]”, before travelling on to Samye, Gansu and China.27 Yet wandering practitioners commonly acknowledged local sacralities, and while numerous Dunhuang texts relate to Tantric rituals28 there is no evidence that ritual worship of mountain or other territorial deities was then a particular concern of the esoteric specialists. We know that in the last centuries of that millennium communities formed around charismatic practitioners of esoteric rituals of the Tantric type and that these groups sought out particular locations of which the Tise/Mapham region seems archetypal. The Bhairavakula Tantra, for example, like the Cakrasamvara Tantra discussed below, recommends the practice of its rites in sites of which this region seems an exemplary example: “On the banks of a river, on the most excellent of mountains, in a meritorious pilgrimage spot, in an abode of Śiva, in a Śakti pilgrimage spot, in an excellent seat (of a deity), or in a cremation ground …”29 Much later sources reinforce the claim that the Tise/Mapham region was one such site attracting renunciates of various persuasions. A late 15th century chronicle, in contexts that seem devoid of mythological elements, mentions a practitioner reputed to have spent a year in a meditation cave beside Lake Mapham during the late 8th century.30 It also states that in the early 10th cen- tury, although “Bön” practitioners had apparently been expelled by the new Buddhist rulers, the kingdom still had practitioners of the Dzogchen (rDzog chen) teachings from the Zhang zhung snyan rgyud lineage31 (established in the 8th century), who were in retreats throughout the region and whose prac- tices overlapped the two faiths. One is stated to have meditated at Tise,32 yet

27 Kapstein (2006: 10–12). By this period, however, Buddhism (and thus patronage) existed in Ngari at least at court level. 28 On which see Dalton & van Schaik (2006). 29 Cited in Hartzell (1997: 869). 30 Vitali (1996: n. 223 166). 31 On which see Blezer (2010). 32 See Vitali (1996: n. 323 223–226) re “Ma-hor stag-gZig”, whose name suggests a western or central Asian origin. 282 chapter 11 he is just one of a number of practitioners associated with different locations in the region. There seems no particular orientation towards Tise or Mapham, rather the general area is indicated as suitable for religious practice. Similarly the reference to a gathering of “cotton-clad practitioners” (i.e.: non-, or post- monastics), at the 1057 royal enthronement33 indicates their presence in Ngari at that time, but not specifically at Tise/Mapham. The pre-phyidar literature does not associate the site with any particular Buddhist deity and any such association before at least the early 13th century is legendary rather than historical. Any earlier Buddhist presence therefore, was presumably because it offered an appropriate location suitable for meditation practices within the wider context of Tantric ritual practice intended to foster identification with the practitioner’s tutelary deity (see below, for example), as well as offering some source of patronage. If there was a sacred focus it was on the lake rather than the mountain, but neither site possessed any trans-regional sacred identity in themselves and there is no evidence that resident practition- ers were specifically associated with that particular landscape. Of course, as discussed in regard to the sacrality or otherwise of mountains in the early literature, the lack of evidence for renunciates at Tise could be due to the very commonality of their regional presence, obviating overt need to refer to them. But any such presumption must answer the question of patronage, which is in any case difficult to account for, not least in the case of Buddhist renunciates prior to king Yeshe Ö’s ascension. We are left with the possibility of esoteric practitioners of the type Yeshe Ö (959–1036) critiqued (see below), having established local followings there or obtaining some other (trading?) income. Tibetan sources do contain one well-known account of a late 8th century Indian Buddhist master’s presence at Tise/Kailas. Buddhaguhya (or Buddh- agupta in some accounts),34 is regarded as an early propagator of Dzogchen and the author of several commentaries included in the canon of the Nyingma sect of Tibetan Buddhism. While on pilgrimage to Mount Potala (in south India), Buddhaguhya was told by an old woman (a common literary trope), to go to Kailas,35 where he would gain magical (siddhi) powers. While meditating there he was invited to Lhasa to propagate Buddhism. The invitation was in the form of a letter from the Tibetan king, whose invitation was accompanied by a large quantity of gold-dust. Buddhaguhya replied that his personal deity, Manjuśri,

33 Vitali (1996: 363). 34 Roerich & Choepal (1996: 170, 191); van Schaik (2004: 186–187); Karmay (2007: 61–63). 35 As we saw in the last chapter, we cannot be certain that the Kailas referred to was Tise. tibet’s tise 283 advised him against making the journey. He did, however, recommend partic- ular texts to the King, including Mahayoga Tantras and the Dharmamaṇḍala sūtra (discussed below).36 Although a number of scholars have been inclined to accept this account as genuine,37 and the letter is included in the canonical Tangyur (Bstan-’gyur),38 there are several reasons to doubt its veracity. The principal objection is that the earliest source for this story is in the Bazhed (sBa bzhed), an account of the establishment of Buddhism under Yarlung king Trisong Detsen (reigned c. 754–797). The dating of this text is problematic for it includes earlier material, but it is unlikely to significantly predate the 12th century,39 and (probably) the earliest version of the text (the dBa’ bzhed)40 does not mention the story. Furthermore, the Bazhed account is by Nyingma apologists whose hagiography of Buddhaguya is intended to legitimate one of their esoteric lineages. By projecting back his presence at Kailas and recommendation of particular texts, these become legitimated in the context of the debate over the validity of early Nyingma literature.41 Thus it is probable that the account of Buddhaguya is a later interpolation for a sectarian strategic purpose, including perhaps, a claim to historical association with a sacred site then in the process of transformation and contestation. We have therefore, only a single reference from the last decades of the first millennium ce., (that mentioned by Kapstein), which indicates that Tise had sacral associations, and that reference is within the period in which Buddhism was becoming prominent in Ngari. The only earlier references to the complex (as will be seen), are metaphorical and devoid of sacral association, and that the plethora of later references post-date the phyidar and the systemisation

36 Davidson (2002: 153–159) includes translations of the letters; also see Dargyay (1994: 102–103); Germano (2002: 229–232). 37 See, for an interesting discussion of the issues and the findings of other scholars, Cutler (1996: 54–74). Cutler accepts the validity of the account. Also see, for a more detailed examination of the sources, Lo Bue (1988). Karmay (1998: 25) doubts the antiquity of Buddhaguhya’s letter on the basis of prophecies it contains. 38 Lo Bue (1988: 790–791) cites this as the sNar thang edition, vol. 94, ff. 387r–391v: Lo Bue’s sources do not lead him to state more than that, “it is conceivable that he may have travelled as far as western Tibet.” 39 On this text and the issues surrounding it, see Kapstein (2000: 26, 62–63 and esp., n. 11 212–214). 40 Wangdu and Diemberger (2000: n. 55 34); also see Huber (2001). 41 Davidson (2002: 154–159, esp., 158). Also see Dargyay (1994: 99–114) whose hypothesis that Buddhaguya was probably the first to put into writing the idea of Songsten Gampo as a Boddhisattva might also explain his prominence in the sources. 284 chapter 11 of Bön further indicates they are a product of later processes. So too does the fact that there is no evidence for a pre-Buddhist Tise yullha. If we may allow the possibility that this was acknowledged only in local oral traditions and maintained in the form of art, dance, or song,42 the absence of a consistent identity for this deity even in later texts problematises any claim to systemised or lasting belief. That Mapham was locally sacred is, however, suggested by wider cultural understandings in the region and the early Indic mythology including both cosmology and Naga-conversion narratives that came to be applied to the lake, as well as later references which seem devoid of mythical context. That would be consistent with the region becoming one where renunciates confronted or co-opted local deities, and in the phyidar it was just such local spirits that the Buddhist project sought to overcome through their ritual transformation into Buddhist protective deities. But as will be seen it was only after the phyidar that the mountain became more prominent and was identified with a specific deity: Demchok Heruka (Skrt: Cakrasaṃvara).

The Phyidar and the Unreformed Renunciates

After the establishment of a new kingdom ruling over Guge-Purang in the early 900s,43 the region was transformed by the ascension there of King Yeshe Ö in the third quarter of the 10th century. In that era the Tise/Mapham sites did briefly pass to the control of another member of its royal family, the (Buddhist) king of Kunu (a region including parts of Bushahr). The Kunu king circumam- bulated Mapham, indicating, if an account from 500 years later is to be accepted as historical, that the lake was of some religio-political status.44 The ensuing history of the Ti-se/Mapham region in what seems a particu- larly significant period for our enquiries is preserved in the “Royal Geneologies of Ngari” (mNga’ ris rgyal rabs). Dated to 1497 and attributed to the abbot of Tholing monastery,45 this work is our main source for the history of the Guge- Purang kingdoms and enables us to contextualise religious events during this period within wider political developments. As with other Tibetan historical accounts, however, it is not a history in the Western academic sense.

42 On songs, see Buffetrille (2000: 100–105); Hein (2007); Bellezza (2005); and Gamble (2010). 43 Vitali (2003: 54). 44 Vitali (1996: n. 321 223–224). 45 Vitali (1996: 93–95). tibet’s tise 285

The Geneologies describe how Yeshe Ö gained control over local goldfields that were to provide the economic basis for the kingdom’s development and how he reorganised, systemised, and centralised control over law, land-hol- dings, taxation, and defense of the state. In the religious sphere, self-conscious- ly identifying with the heritage of the Yarlung dynasty, Yeshe Ö is credited with initiating the phyidar, the process which fully indigenised Buddhism in the Tibetan world.46 The phyidar involved several interlinking processes. Indic Buddhist texts and teachings were collected and translated into Tibetan to form the canon- ical foundations of Tibetan Buddhism. This was accompanied by an ambitious project of temple-building sponsored by local rulers, with these centres cre- ating institutional settings from which to spread the canonical teachings. The most significant aspect in our context, however, was the imaginal transforma- tion of the Tibetan landscape and its resident deities into manifestations of a Buddhist world. The phyidar was an extraordinarily dynamic enterprise. It saw considerable (and probably unprecedented) movement across the Himalayas by artisans and craftsmen, translators and religious figures, as well as by those employed in the service of these groups; porters, pony-men, and so on. It must have resulted in the spread of a greatly enhanced knowledge of the routes through the Himalayas and it attracted numerous religious practitioners from the Indian sub-continent. An outstanding figure in the phyidar and the key agent in the regional temple-building process was Rinchen Zangpo (958–1055),47 remembered as the ‘Great Translator’. He was among a group of youths sent to Kashmir by Yeshe Ö to study Buddhism and he returned in 987, around a year after a royal edict issued by Yeshe Ö decreed the adoption of Buddhism within his realm, which act may be considered to have initiated the phyidar. Rinchen Zangpo was subsequently credited with establishing temples along much of the Sutlej route to northwest India,48 and along with a series of monasteries built in this period (most notably the establishment of Tholing in 996), these structures laid the basis for the regional development of Buddhism.

46 Vitali (1996: 249–250, also see 324–325). While there were earlier initiatives by Yeshe Ö’s father aimed towards developing Buddhist structures in Ngari, it was only under the patronage of Yeshe Ö and his successors that Buddhism was actually institutionalised there. 47 On whom see, Tucci (1988). A modern biography taking into account sources found since Tucci’s pioneering work is a lacuna. 48 Vitali (2003: 56). 286 chapter 11

Many of the new temples were constructed on sites already sacred in indige- nous understanding, thus overlaying Buddhist structures onto local sacred spaces. In addition, the territorial deities of the region were re-conceptualised, ritually converted to Buddhism and assigned a role as protective deities within the Buddhist pantheon, a conversion process in which Rinchen Zangpo was apparently particularly skilled. It is clear however, that other religious beliefs and practices—generally subsumed in the Buddhist sources under the rubric of ‘Bön’—continued to be wide-spread and that those elements challenged the Buddhist project. YesheÖ is recorded as initiating the suppression of Bön, yet the situation was more complex than a simple binary opposition. There is for example, evidence that key figures such as Rinchen Zangpo studied Bön texts,49 and in the earlier literature it was not Bön that was depicted as the main opposition to Yeshe Ö’s Buddhacisation project but rather the practitioners of the esoteric Tantras. This tends to confirm the absence of any authoritative Bön religious institutional structures actually existing to oppose Buddhism at the turn of the millennium (an issue discussed in chapter 13). It also indicates the tensions between extant forms of Buddhism within the process of its transmission from India. While Tantric practitioners such as Padmasambhava are renowned for bringing Buddhism to Tibet under the Yarlung dynasty, the post-dynastic mi- lieu was very different. In the absence of any dominant religious authority or structures, the institutional and social collapse of the Yarlung Buddhist realm created a religious vacuum. This was filled by charismatic individuals; local shamans, former attendants of the now-ruined temples, and the already power- ful cemetery-dwelling renunciates with their deity-centered rituals that floated free of categories of ‘Hindu’ or ‘Buddhist’. Such individuals established con- siderable followings (principally but not entirely among the laity50), for their diverse practices. These famously involved animal sacrifices, sexualised rituals (sByor ba), and ritual ‘liberation’ (sGrol ba: i.e. killing). The result was that as Davidson put it, “every sort of drooling schizophrenic could claim to be a fully enlightened siddha as he cut the head of yet another hapless beast in the name of liberation.”51

49 Vitali (2003: 228–229). 50 For an example of these practices and problems arising from them among the regional elites, see White (1996: 308–309). White cites the (Kashmiri) Rajatarangini concerning the 936ce downfall of King Cakravarman after his adoption of Tantric rituals, which led to the Tantrics taking over his court. 51 Davidson (2004: 79); also see, for an example of this tendency involving such esoterica as cemetery practice, corpses at the centre of the mandala, the selling of human flesh, yaksas as sexual partners, etc.; Davidson (2004: 204–205). tibet’s tise 287

Following the re-establishment of central political authority in Ngari,52 Yes- he Ö famously issued an ‘ordinance’ (bka’ shog) around 985, directed against;

You tantrists, who live in villages … Indulging in the ten evil ones and taking on the mode of life of dogs and pigs … You say “we are Buddhists”, but your practices, Show less compassion than an ogre … By offering faeces and urine, Semen and menses to pure divinities Alas! You will be reborn in a mire of rotting corpses … By way of retribution for killing animals through your ‘rite of deliver- ance’, Alas! You will be reborn as an ogre. By way of retribution for indulging your lust in your ‘sexual rite’, Alas! You will be reborn as a uterine worm … If these practices, like yours, bring about Buddhahood, Then hunters, fishermen, butchers and prostitutes, Would all surely have attained Enlightenment by now.53

After this edict Tantric practices involving sexualised rituals did continue, but they were considered the preserve of advanced practitioners who had left the monastic setting and taken to the ‘wandering’ phase of practice (caryā).

52 Re “heretical” lineages recorded in this region, see Vitali (1996: n. 315 214, 217–218). 53 For the full epistle, with Tibetan text and discussion, see Karmay (2003); also see Seyfort Ruegg (1984); Cantwell (1983: 107–118). Yeshe Ö’s ordinance was part of a wider regional response to the more extreme aspects of Tantric practice, demonstrating how far these teachings had spread. A western Indian text of the 10/11th century similarly asks; “if liberation is to be identified with the excitation of the female genitalia, would not even donkeys be liberated? Indeed why are rams and bulls not liberated?” White (1996: 172), citing the Rasarnava [‘The Flood of Mercury’]). In Kashmir around this time the philosopher Abhinavagupta reinterpreted Tantric scriptures to remove the extreme elements; White (1996: 7). That became an increasingly popular strategy. In Burma the esoteric Aris sect were suppressed; Seyfort Ruegg (1984: n. 25 133); also see, particularly for sources on the Aris, Bhattacharya (1996–1997: 52–67). A somewhat later response came in Japan in the mid-13th century, where “an orthodox Shingon backlash” paved the way for the suppression of such practitioners in the 15th century; White (2001: 20); also see Faure (2001: 544). While the responses may not have been directly connected, their initiators must have had similar interests, which seem to centre on the consolidation of religious and political power at royal courts. 288 chapter 11

Alternatively they became secret practices carried out with some degree of discretion,54 or practices domesticated within the Nyingma sect, whose monks were permitted to marry. The centralisation of royal power required bringing religious institutions under royal command and the suppression of alternate or independent sources of religious authority. In Ngari, where the assassination of the last Yarlung king by a Tantric practitioner was part of royalties’ family memory,55 those who claimed the authority to ‘liberate’ anyone they disagreed with were an obvious target. Given that Yeshe Ö’s ordinance also complained that “as the ‘sexual rite’ has become popular the different classes of people are mixed”,that suppression was also designed to reinforce social hierarchies in the new kingdom on the earlier model.56 There was also an economic factor. By centralising religious power in state structures, the new kingdoms could channel religious funding by the laity into state institutions patronised by, and consequently supporting, the royal court.57 Money and goods that had previously gone to individual teachers could now be diverted to state-approved institutions, thus removing the economic foundations of practitioners outside of their authority. In the wider context these changes suggest that fundamental issues of au- thority and of social structure were involved. An individual renouncer, living outside the social boundaries and surviving on roots and berries or occasional charity posed no threat to the established order. Such behaviour may have had a political dimension in that a renunciate’s sojourn in the wilderness brought that land within the ‘known’, and tempered its impurity by spiritual practices there, yet in that sense the renunciate was an agent of the culture and author- ity of tribe or state, not a threat to its power. But when renouncers grouped together or formed fixed communities outside of central control they became a potential threat to state control of ideology and society. Thus these alternative

54 See for a modern example, Campbell (1996: 2) who states that “[H]e requested that I become his sexual consort … despite the fact that to outsiders he was a very high-ranking yogi-lama of the Kargyu lineage, who as abbot of his own monastery, had taken vows of celibacy.” 55 The assassination of King Langdharma by Lalung Palgyi Dorje; the latter has however, also been identified as the 9th abbot of Samye monastery; see Karmay (1988: 9). 56 The possibility that this was a public health issue might also be considered: on sexually transmitted diseases in Tibetan society in the modern period, see McKay (2010: 41–60). 57 The religious communities that formed around charismatic renouncers might hold great wealth; van Scheik (2010: 75) cites vows (from Oriental and India Office Library manu- script, Tib j348/2) that include “not coveting the wealth and jewels of vajra siblings.” tibet’s tise 289 power structures had to be assimilated or suppressed as part of the centralisa- tion process. The established elite response to potentially divisive religious communities was to control them and limit their activities to specific, generally marginal spheres of authority in order to ensure that they served the wider aims and interests of the authorities. Patronage was the principal means by which this could be achieved. By offering reliable financial support to religious groups, and cutting off their alternative funding sources, they could be brought under control. This process was a conscious strategy, specifically recommended to kings in eminently realpolitik texts of statecraft since that most Machiavellian of political guides, the Arthasastra.58 As noted with Tantric alchemical economies however, some renunciates did maintain funding sources not easily controlled by central authorities. Many were obviously flexible enough to exploit a variety of financial opportunities, including patronage and trading networks as well as popular support, in order to maintain their independence. As noted, this flexibility has been character- istic of Himalayan renunciates and an important aspect of their continuous viability, enabling an organic continuity through on-going reformulation of identity, role and practice. The phyidar did succeed in absorbing much of the anti-structure impetus and creating firm foundations for Buddhist expansion in Tibet under state con- trol. But the esoteric avenues of spiritual expression could not be entirely closed off because according to the criteria adopted for establishing authenticity many such practices proved to be canonical.59 Confronted by a wide variety of tra- ditions claiming to be Buddhist, the Tibetans had turned to the Holy Land of India, the source of Buddhism, to ascertain their validity and concluded that

58 Attributed to Kautilya, Brahman adviser to Chandragupta Maurya, much of the Arthasas- tra is almost certainly later. It includes strategies for gaining, preserving and increasing royal power in what is conceptualised as an amoral world where the strong vanquish the weak. Thus the concept of divine kingship is presented as a strategy for fooling the ignorant, while agent provocateurs are recommended to test the loyalty of officials and subjects. Spies were to disguise themselves as renunciates and renunciates themselves made ideal spies, able to travel freely and mix in all social circles; Arthasatra 1.11.7.1–18 and 21–23; 13.2.172.1–18: (Rangarajan 1992: 477–478). Esoteric and occult skills were also to be utilised; Arthasastra 14.3.88, (Rangarajan 1992: 509.); on which see Olivelle (1987). 59 The issue of textual authenticity had already arisen in central Tibet. As early as c. 814 the authorities of Samye monastery had ruled that as Tantric practice had become “perverse”, the Tantras could only be translated and circulated with royal approval; Snellgrove (1987: 443) citing the sGra sbyor bam po gynis pa. But see Snellgrove (1987: 456–457) regarding the limited success of the edict; also see Karmay (2008: 5). 290 chapter 11 texts for which a Sanskrit original existed were authentic.60 Indic authorities such as Atiśa (982–1054), the abbot of the great Vikramaśila monastery, were then invited to Tibet by the Guge king to advise on the ideal utilisation of these practices. Atisa apparently recommended that the esoteric practices be forbid- den to those bound by the monastic code (vinaya), but he is unlikely to have wanted them suppressed; before becoming a monk Atisa himself had stud- ied esoteric Tantras while a disciple of Mahasiddhas such as Naropa (c. 956– 1040).61 A modified form of renunciates thus emerged, who were to be central to the Buddhist encounter with the Tise/Mapham sacred complex.

The Return of the Reformed Renunciates

The gathering of Buddhist masters (choskhor) which began at Tholing in 1076,62 brought together a large number of local and foreign Buddhists in what can be seen as the culmination of the Guge-Purang royal monasticisation project.63 But while Tholing remained the monastic centre of Ngari, new power struc- tures were developing in central Tibet and increasingly the subsequent focus of religious development moved to Ü-Tsang (central Tibet). Famine, rebellions by Dardic subjects, and invasions by gar log tribes from the northwest weakened Guge-Purang, which split into separate kingdoms in the late 11th century.64 The power of the various Ngari kingdoms subsequently varied on the ‘galactic polity’ model, and shifting sovereignties and weak central government allowed the rapid re-emergence of alternative and non-monastic religious tendencies as historical agents. The collapse of the unified polity in Ngari initiated a new phase of religious construction and a new emphasis on the conversion of landscape and local deities. While state-sponsored monastic Buddhism had been firmly established

60 The Nyingma developed their own canon, on which see Eimer & Germano (2002); also see Davidson (2004: 104–120). 61 Snellgrove (1987: 147–149, 493). 62 On which see Shastri (1997). 63 One interesting guest was the ritualist of whom it was said that the, “many great Lamas and officials (bla dpon) … had no choice but to bow to him. It was known that those who wouldn’t would be ‘liberated’ by the coercive spells of Yamāntaka.” Martin (1982: n. 39 68–69). This indicates that some of the esoteric ritualists critiqued in Yeshe Ö’s ordinance were sufficiently powerful or had made sufficient concessions to enter into the new power structures. 64 Vitali (1996: 324–352). tibet’s tise 291 in the region, the principal agents of change in the ensuing period were again renunciates. But they were now operating under a new paradigm, organised into sects and articulated in terms of a new ideal, at least overtly purified of its more extreme tendencies and brought more closely into association with the new monastic structures. These Buddhist renunciates were at least nominally monks and essentially they formed a loose alliance of interests with monastic Buddhism. Institutional acknowledgement of this was the posthumous eleva- tion of certain renunciates to an idealised status that remembered them as enlightened masters. In addition, at least at the elite level Tibetan Buddhist authorities in this period were attempting to define more sharply the doctrinal identity of both Buddhism and within that framework, sectarian orders. The early concern with the challenge from esoteric practitioners was replaced by a more formal artic- ulation of opposition to a new ‘Other’ in the form of Bön, whose claims to mas- tery of indigenous traditions contested Buddhist control over local territorial deities and their realms.65 The two belief systems used their own formulations of indigenous categories of deities and ways of seeing the landscape as part of that contestation.66 Thus Rinchen Zangpo is said to have, “selected only those places for establishing [temples] … which were either the centres of Bon-po faith or the local gods.”67 If Buddhism was not to become absorbed into the general religious world, as to some extent happened in India, it required a strong, continuing articulation and manifestation of its distinct character. While we are poorly (if increasingly

65 Clearly there are many elements of Tibetan religion that do not fit easily into categories of Buddhism or Bön, while both systems have been used to identify competing elements of political interests. But while many commentators would emphasise the similarities and co-existence of the two systems, competition is both specifically articulated in accounts of the period and an important factor in the history of Tise/Kailas. 66 Here we must differentiate between those elements that exist within blurred boundaries, that is, local practices and processes as well as non-elite understandings and worship, vis- à-vis the central agencies of the disparate faiths. While the introduction and development of Buddhism into any Asian society has always been marked by relationships with estab- lished local deities and beliefs, such compromises must also be understood as a means to an end. The central institutions of Buddhism may take on indigenous characteristics, but their relationship to local sacralities is always shaped by the understanding of their subjugation by Buddhism; there is no question of equal status. (I am indebted to Wayne Shipp for pointing out that the introduction of Buddhism into the usa has not involved an encounter with local deities, although it has entered such an encounter with local faiths, notably the New Age, but also indigenous [‘Native American’].) 67 Thakur (2001: 35); also see Thakur (2011: 209–218). 292 chapter 11 well) informed as to the circumstances in which Bön acquired sufficient dis- tinct identity to be seen as constituting a threat to Buddhism (and clearly this is tied to the interests of particular clans and established aristocracies), it is dif- ficult to locate that process earlier than the period in which Tibetan Buddhism itself took on a sectarian character. Thus, in sum, while we may legitimately explore and even emphasise overlapping understandings in relevant contexts, competition and contestation are the most prominent aspects of the historical processes centred on the Tise/Kailas region in this period.

Milarepa versus Bön?

In Tibetan understanding, Tise became established as a Buddhist mountain as the result of a contest between the magical powers of two renunciates, one Buddhist, one Bön. The tale is reminiscent of that ancient form of ritual warfare in which two sides each put forth their greatest warrior, with the result of their combat accepted as deciding the entire conflict, and as implying divine favour to one or other side. The Buddhist Hero of this myth was the archetype of the reformed model of the renunciate, Milarepa (1040–1123?), who is recorded as visiting Tise in 1093.68 There are various accounts of his motives for the journey. According to the prosaic account in the Blue Annals (1476), he travelled there “in order to labour for the benefit of the crowds of nomads.”69 His (1488) biography attributes the journey to his guru Marpa recommending the site to him as his karma was ripe to meditate in such places.70 An 1896 guide-book to Tise however, states that he went in fulfilment of a prophecy by the Mahasiddha Naropa, the principal guru of Milarepa’s guru, Marpa. According to this prophecy, Milarepa’s visit there would result in Tise and Mapham becoming the meditation places of his followers.71 Not only was this the case, but Milarepa’s tutelary deity was to become the Buddhist deity most closely associated with Tise. Whatever may have sent him there, Milarepa’s visit is framed in terms of a Buddhist-Bön competition for the mountain. In the most prominent of a num-

68 Vitali (1996: 323). Milarepa was from a clan with close links to Bön and aspects of his biography indicate his knowledge of that tendency, although of course that does not rule out that he opposed it. 69 Roerich and Choepel (1996: 433). 70 Evans-Wentz (1951: 163–165), the translation is by the Sikkimese Kazi Dawa Samdup, although the terminology is heavily Christian-influenced. 71 Fillibeck (1988: 73). tibet’s tise 293 ber of accounts of his marking his presence on the landscape, he is described as emerging victorious from his contest with the Bönpo renunciate, Naro Bön- chung. Milarepa, and by implication Buddhism, had demonstrated their supe- rior powers and were henceforth supreme over Tise with the Bön sportingly allowed their own mountain nearby, that now known as Bönri. The land- scape was thus understood to be transformed; what had formerly been the “divine countenance” of the Bön deity Gekhod was now the “divine palace of Cakrasamvara.” But this central mythology is problematic. Naro Bönchung does not appear in early Bön sources and the earliest version of this story, by the 3rd Karmapa, dating to a little over a century after Milarepa visited Tise, was simpler and less polemical. It stated only that Naro Bönchung challenged Milarepa’s claim to the mountain, but that while the Bönpo “had obtained some miraculous powers he was unable to win.”72 A more critical account of the events occurs in a mid-19th century Bön guide-book to Tise which ‘disowns’ Naro Bönchung. It states that;

Once (Milarepa) went on pilgrimage to Mount Tise with some disciples, and fully cognisant that since the beginning of time it had been the “mountain of the la’” [i.e., bla ri ‘soul mountain’] of everlasting Bön and a retreat centre for its followers, he was overcome by feelings of hatred and attachment and thought he would ruin the reputation (of Bön). He changed the name of his disciple Rechungpa73 to ‘Naro Bönchung’ and together they discredited Tise, and without any respect for the truth they filled the land with lies … The episode when Naro Bönchung fell from his drum while flying merely shows that his disciple Rechungpa had not obtained the signs of realisation.74

How then might we read these accounts of Milarepa at Tise as history in the Western understanding? On one level they obviously represent contestation over control of landscape and the ultimate victory of Buddhism (while allowing a continuing Bön presence), with all the implications that that contestation

72 Martin (2001: 117–119). 73 Rechungpa (Ras chung rDo rje grags pa: 1083–1161) was a pupil of Milarepa. His literary function in this account is primarily as a vehicle for Milarepa to explain himself. On this figure, see Roberts (2007: esp., 78, 92.) 74 Norbu (1995: 187 quoting the Tise dkar chag by Karu grubten Tenzin [dkar ru grub chen bsTan ’dzin], published and partly translated by Namkhai Norbu and Raman Prats; Rome, 1989). 294 chapter 11 and result has. Yet Milarepa is not considered to have ‘opened’ the site to pilgrimage,75 and another century was to pass before those who inherited the traditions of Milarepa actually moved into the area in any numbers, something that occurred in very different historical circumstances. Furthermore, Milarepa does not actually seem to have stayed long at Tise. In contrast, his biography states that he stayed for 12 years at Labchi (la phyi), another famous mountain pilgrimage site in southern Tibet. Most probable is that conditions at Tise were actually unsuitable for him to practice there, and that the later accounts of his presence are a construction of history in which visionary revelation serves as validation; something acceptable to the Tibetan world-view if not to Western historical analysis.76 Milarepa’s later prominence in Tibetan history is at least partly due to his popular appeal to the non-elite classes, for whom his life—its difficult youth, his use of ‘black magic’ for revenge against those who had wronged his fam- ily, and then his trials under Marpa and his ultimate freedom and enlighten- ment—provide an entertaining morality play that implicitly critiques the weal- thy monastics. (As such he serves as a ‘working class hero’ who was subse- quently eulogised by the very power structures he critiqued, a phenomena with numerous historical parallels.) But such accounts require agencies of

75 On the importance and visionary requirements of the ‘opening’ of a pilgrimage in the Tibetan tradition, see Karmay (1996: 9–10). 76 Whether Milarepa actually did visit Kailas, or whether he has been transformed by literary strategies into a transnational missionary on the model of Padmasambhava, the Pandavas, and other such figures, is an interesting question. Certainly accounts of Milarepa’s activi- ties at Kailas are difficult to reconcile with his biographies. They indicate he undertook a nine year retreat until around 1093, and met his disciple Rechungpa in, according to the Blue Annals, early 1094 to early 1095, and thus could not have been at Kailas with him in 1093. Milarepa’s visit, if such it was, must have been brief. It is notable that while he occupies a central role in the guidebook to Labchi written by the 34th Drigung heirarch, he is mentioned in the Tise guide book by the same author only briefly in regard to the competition with Naro Bonchung. We know that Milarepa’s life-story is the product of a complex vector of narratives and political strategies that developed over the three and a half centuries before Tsangnyön Heruka (1452–1507) compiled The Life of Milarepa. See, for example, Goss (1993); also see Quintman (2006). Ultimately, for the believer, Milarepa’s importance lies in symbolising the validity of the renunciate path and how karmic impediments can be overcome in a single lifetime. Davidson (2005: 255; also see on problems with Milarepa’s biography, 142–143), high- lights the importance of Milarepa’s songs in reaching the broader populace; Ruth Gamble highlights the importance of humour in these songs and narratives, see Gamble (2010). tibet’s tise 295 promulgation to spread beyond local fame and this portrayal owes much to the desire of later Kargyu chroniclers to firmly anchor their lineage and its validity at this location, as well as to the continuing acknowledgement of the validity of the non-monastic path to Liberation by Tibetan Buddhism in gen- eral.77 The predictions attributed to Marpa or to Naropa concerning Milarepa’s achievements were thus retrospective validation rather than historical expla- nation, for the truths they contain exist in the world of myth. Noteworthy is that Marpa is said to have stated that, “the Tisé peak (Mount Kailāsa) hath been mentioned by the Lord Buddha Himself as the Great Mountain, the abode of Dēmchog …, and a fit place for meditation.”78 That claim does not appear to have otherwise been made until it was advanced by Jigten Gönpo (see below) around a century after Milarepa’s visit to Tise.

Sectarian Developments

In the wider context Buddhist competition with Bön represented just one aspect of a period of struggle within the Tibetan religious world over authen- ticity and authority, a period that culminated in distinct identity formations. Such claims to a distinct identity (and to a global validity), were a departure from the indigenous world-view in which religious activities were essentially pragmatic, rather than concerned with ultimate transcendence.79 The emergence of Bön as a systemised belief system advancing a transcen- dent path appears to date only to the period immediately prior to, or around which the institutionalisation of Buddhism in Tibet manifested in the develop- ment of sects attributed to Tibetan founders. That occurred after the mid-11th century. This implies that the two tendencies were in a dynamic relationship of action and response and shared the experience of various historical processes, notably systemisation.

77 As Davidson observes, we have no writings by Milarepa himself and “his literary persona became something of a vehicle through which the various authors could express their own feelings and intuitions, ones they may not have wanted under their own names.”; Davidson (2005: 256). 78 Evans-Wentz (1951: 164). In the 1896 guide-book it is Gandhamadana (here separate from Gurla Mandata) that is prophesied by the Buddha; see Huber & Rigzin (1999: 128). 79 Huber (1999: 23) notes that in general terms the worship of the yullha addresses issues of immediate concern and that of the gnas ri mountain issues of worldly transience, but an overlap is observable. 296 chapter 11

No distinct religious order or institutionalised division had arisen in Tibet during the first dissemination of Buddhism, but the 1057 founding of Ret- ing (Rva sgreng) monastery by Atisa’s chief disciple was the precursor for the emergence of Tibetan religious orders with defined identities. Followers of this lineage characterised by emphasis on monastic discipline first developed a distinct sectarian identity under the name Kadampa (bKa’ gdams pa: lit: [those] ‘bound by command’). This distinguished them from followers of estab- lished traditions, notably those associated with Padmasambhava, who became known as the Nyingma (lit: ‘Old Ones’). The founding of Sakya monastery in 1071 marked the formal beginning of the sect of that name while the Kargyu (bKa’rgyud: ‘Oral Transmission’80), the sect most relevant to this work, emerged institutionally in the early part of the following century. Sects came to be defined primarily by their teaching and textual lineages, and in each case looked back to such lineages as deriving from within Indian Buddhist tradition; indeed that became fundamental to their claim to doctrinal validity. But the founders of the sects were—Atisa apart—those Tibetans who had bought the teachings in India and brought them back to Tibet. They were understood to be authorised holders of the Indian lineages.81 In the case of the Kargyu, that holder was a layman, Marpa (1012–1097?),82 who had made three journeys to India to obtain Buddhist teachings. Most significantly in this con- text he returned with authority to Tantric lineages believed to derive from the Tantric Buddha Vajradhara and received in visions by the Mahasiddha Tilopa. Tilopa had transmitted these to his disciple, the one-time abbot of the great Indian Buddhist monastery of Nalanda, the Mahasiddha Naropa.83 While the centrepiece of these teachings were those known as the ‘Six Yogas of Naropa’,in the context of Mount Kailas the most significant was the Cakrasamvara cycle. Having returned to Tibet, Marpa transmitted the lineages to his disciple, Milarepa, and while this archetypal hermit came to represent the antithesis of the monastic path, his disciple Gampopa (1079–1153) founded the Kargyu monastery of Dvagspo in 1121. Gampopa’s own disciples went on to found six sub-sects of the Kargyu, of which, as will be seen, four were to be prominent in the Tise region. These four were:

80 More correctly, “a line of transmission of an esoteric teaching from teacher to disciple.”; see Smith (2001: 40). 81 In obtaining initiation into particular traditions in India the Tibetan lineage holders were thus authorised to teach and commentate on that tradition. 82 Concerning his dates, see however Davidson (2005: 142–143). 83 On whom see Davidson (2005: 44–49, 142–147). tibet’s tise 297

(1) The Tshalpa (mTshal pa) Kargyu. Their Gungthang monastery in Tshal, near Lhasa, was founded in 1175 by Lama Zhang (Zhang Yu druk pa: 1123–1193).84 (2) The Pagmo Drupa (Phag mo grub pa) Kargyu. Founded by Gampopa’s disciple Pagmo Drupa (Phag mo grub Rdo dje rgyal po: 1110–1170).

Two of Pagmo Drupa’s disciples founded the following sub-sects of his tradi- tion:

(3) The Drigungpa (’Bri gung pa) Kargyu. Their chief monastery of that name northeast of Lhasa was founded in 1179/80 by Jigten Gönpo (’Jig rten mgon po: 1143–1217). (4) The Drukpa (’Brug pa) Kargyu, whose principal monastery of Ralung (Rwa lung) was founded in 1180 by Tsangpa Gyare Yeshe Dorje (Gtsang pa Rgyas ras Ye shes rdo rje: 1161–1211).85

When institutionalised into monasteries, these traditions generated consider- able wealth. In accord with Indic traditions, the texts and teachings obtained by Tibetans from India were commercial commodities. In what was understood religiously as an offering, Tibetans generally paid for them in gold,86 and on their return regained their investment through offerings from followers they attracted with their new teachings. Monasteries also attracted patronage by royalty (‘the state’), by the local aristocracy that was usually integrated into the establishment through incarnation lineages, abbotships, etc., and also by the laity. In addition, monasteries generated revenue through endowed land-

84 On whom see Yamamoto (2009). Yamamoto demonstrates the extent to which such figures were heir to numerous lineages, with the historical predominance of any one of these due to a complex variety of circumstances. Snellgrove (1987: 488), attributes the founding of this monastery to Gampopa (whose death he dates to 1169), but the differences need not detain us here. Also see Davidson (2005: 328–330). 85 Richardson and Snellgrove (1968: 137), and Samuel (1993: 479), state the monastery was founded by Lingrepa Pema Dorje; Gene Smith however, attributes it to Tsangpa Gyare Yeshe Dorje (Gtsang-pa Rgya-ras Ye-shes-rdo-rje; 1161–1211), the founder of the sect; see Smith (2001: 43–44) and also see Chandra (1968: 2). In some cases it seems the inspiration came from the founder, the actual achievement from one of his disciples. See Davidson (2005: 333) for a summary of Pagmo Drupa’s career. 86 On the implications of this in regard to the possibility that demand produced supply— that some Indians, “arrived … with armloads of esoteric works whose ink was barely dry”, see Davidson (2005: 232–234). 298 chapter 11 holdings, trading, and financial investments, and maintained forces sufficient for internal discipline and external defence.87 The largest of them were virtu- ally religious city-states. While sectarian identities took time to become fixed, and at first “there was no obvious territorial rivalry between the newly-formed sects”,88 this situation soon changed. Tibet in the 12th century was characterised by crime, disorder and factional fighting. The sources speak of famine and a breakdown of social order, with even pilgrimage and trade becoming unsafe. This impacted on resources and increased competition between centres, which must have con- tributed to identity formation. Monasteries in Lhasa and Samye were burned down in the fighting. Lama Zhang, who developed his own militia forces and territorial power, specifically built his monastery to protect visiting pilgrims yet it too was burned.89 This was a period when even renunciates were advised to arm themselves with weapons,90 and Drakpa Gyaltsen (1148–1216) the Sakya commentator (and uncle of the Sakya Pandita), warned that at places such as Tise, “there are many nomads … [who] … do all kinds of bad stuff … you’ll repent of having come, dead by a knife.”91 Reconstructing the events of this troubled period is difficult, but in the absence of any central political authority on the Tibetan plateau it appears that competition between the old aristocracies manifested in competition between the monastic establishments they patronised, with the Sakyapas emerging as the Kargyupas’ main opponents. Sakya produced a number of esteemed scholars and an outstanding leader, Kunga Gyeltsen (1182–1251), commonly known as the Sakya Pandita. Even before the Pandita’s rise to prominence however, Kargyu pioneer Pag- mo Drupa stressed the importance of his followers practicing at the sites— Labchi, Tsari, and Tise—that he associated with the tutelary deity of the tra- dition; Demchok Heruka (ie: Cakramsamvara).92 Yet we must suspect that the Kargyu’s subsequent orientation to the southern and in this case western peripheries of the plateau owed something to sectarian competition as well as to the disturbed conditions in central Tibet and the availability in Purang of royal patronage from a new king, Nam Lugyel (klu rgyal).93 As Vitali states,

87 For a discussion of this in the context of China, see Gernet (1995: 94–152). 88 Richardson (1976: 18–23). 89 Yamamoto (2009: 34–35, 174–204); Martin (1996: 44); Davidson (2005: 328–330). 90 Kapstein (2003: vol. ii, n. 38 537, citing the Maṇi bKa’ ’bum i.525). 91 Davidson (2005: 323, quoting Drakpa Gyeltsen’s Song of Realization, in Praise of this Place). 92 Petech (1978: 315). 93 Petech (1978: 316–317); also see Vitali (1996: n. 594 372–373). tibet’s tise 299 these practitioners, “were not sent to meditate at the holy mountain solely for spiritual reasons. Ensuring the survival of members of the [Drigungpa] com- munity necessitated the stable occupation of meditation places …”94 While individual Kargyu practitioners may have travelled to Tise somewhat earlier and reported favourably on prospects there, Jigten Gönpo, the Drigung hierarch, organised the first group of his sect’s meditators to travel there in 1191 (in addition to groups sent to Tsari and Labchi). They intended, according to later sources, to establish hermitages for future groups. But with the fragmen- tation of the Guge-Purang kingdom the weakened regions were increasingly subject to raids from external forces, notably “devastating” attacks from the south or southwest in 1193–1194. In these circumstances the first Drigung expe- dition had little success,95 and a group of Tshal Kargyu consequently delayed their planned journey to the region until 1195.96 From this latter group it was Sangye Tshalpa, a disciple of one of Lama Zhang’s followers, who was the first recorded member of his tradition to visit Tise. Sometime between 1195 and 1200 he established the Tshalpa regional cen- tre at Taga (rTa sga: to the north of Nubris). There he served briefly as abbot before being succeeded by a new arrival, the Tokden (rTogs ‘dan97) Dharma Sonam. But in 1200 Sangye Tshalpa was joined by an old companion, Marlungpa Wangchuk Sengye (1153–1241), a Tshalpa from a Bön lineage family with early associations with the Tise region. Marlungpa had studied under Lama Zhang, whose biographer he was, and from whom he had received Naropa’s Oral Transmission lineage.98 Marlungpa gained the Purang King as a follower, while Dharma Sonam established a formal yon mchod (‘patron-priest’) relationship with the king after curing him of a serious illness while at Tise/Mapham. Dharma Sonam is described as, “a great yogin … [who was] … greatly enterpris- ing in practical matters, rather more so than” Sangye Tshalpa,99 and the king granted him control over several water mills and monastic centres, underpin- ning the Tshalpa local economy.

94 Vitali (1996: 376). 95 The 1895 guide-book discussed by Huber & Rigzin (1999: 144) states however, that thanks to the siddhi power granted to Jigten Gönpo by the Naga king Madros, a bountiful harvest followed his arrival in Ngari, such that the Drigung Kargyu could even supply grain to Lhasa. Given the context of the famine conditions behind their westward movement, a good harvest would have had a major impact on their reception in the region. 96 Vitali (2003: 68–70); Vitali (1996: 372–373). 97 A proper study of the Tokdens (Kargyu renunciates) is a lacuna; see however, Russell (1996). 98 Vitali (1996: n. 320 221–223, 393–395). Yamamoto (2009: 48, 99) identifies him as Mar sgom or Mar pa lha dkar. 99 Vitali (1996: n. 644 396. citing Chos legs rnam thar, f. 10a, lines 4–5). 300 chapter 11

There were other Tshalpa meditators under another of Lama Zhang’s pro- tégés, but despite their having also received royal patronage their centres were placed under Dharma Sonam’s authority.100 His abilities gave his line of Tshal- pas a firm foundation, one that enabled them to maintain their position despite subsequent political changes. This culminated in his successor being given rights over Tholing monastery by the Guge king Gragspa (1230–1277), who had gained control of Purang and the passes to the south in the interim. Tshalpa ascendency was to be brief however, for the region passed to Sakya control after 1277.101 In the meantime, famine in central Tibet inspired two waves of Drigung Kargyu to follow up their 1191 initiative. A group of around 300 reached Tise in 1208 and a larger second group arrived in 1215 after famine had driven them from Labchi. These Drigung found support from the Purang king, whose realm had entered a period of prosperity and religious resurgence that saw both the Tshal and Drigung Kargyu given generous patronage.102 In these favourable circumstance, Tise/Mapham attracted a number of other meditators. These included members of another Kargyu sub-sect, the Karma— later to emerge as the most prominent of the Kargyu groups in Tibet,103 and most notably the Drukpa Kargyu. In particular we are concerned with Göt- sangpa Gonpo Dorje (rGod tshang pa mGon po rde rje: 1189–1258), one of the most influential renunciates to visit the region. He arrived there in 1214, by which time the Drukpa sect had already established three local centres. Although these were given over to the Drigung the following year, Götsangpa remained there for at least two years, and circumambulated Tise with a com- panion, one Dampa Tsang. It was Götsangpa who played the crucial role in the understanding that the mandala of his sect’s tutelary deity, Demchok, was inscribed on the landscape of Tise/Mapham. Yet the favourable conditions may have attracted too many renunciates, for the rights to practice at the site were now becoming contested by the various Kargyu followers, suggesting patronage and essential resources were not unlim- ited. Götsangpa’s biography tells of his being told by a Drigung renunciate, “This is our cave and you are not allowed to spend the winter here.” Displays of magi- cal power were a means by which competition between the renunciates of the different sects were resolved. One, for example, gained royal patronage after reportedly flying to the centre of Mapham while in meditation posture, and

100 Vitali (1996: 397–399). 101 Vitali (1996: 401, 442–445). 102 Vitali (1996: 372–377); Vitali (2003: 72–73). 103 Vitali (1996: 423). tibet’s tise 301 there giving teachings to Madros. Götsangpa too engaged in these displays out- shining, according to his biography, those of the Drigung by, “manifesting the thirteen gods of the [Demchok] cycle on the thirteen discs [of the mandala].”104 But it was the Drigung who became the dominant sect in the region from around 1215, at which date Vitali concludes that “the [Tise] pilgrimage was instituted in a stable and organised way”.105 They established close ties with local royalty that survived court intrigues and saw Drigung Kargyu serving as royal priests and abbots of the major regional monastery of Tholing and other religious centres. While other Kargyu sects were also patronised, the Drigung were paramount.106 Not the least of the Drigung’s diplomatic achievements was to find patrons among the rapidly expanding power in the north, the Mongols, who were increasingly influential in Tibetan affairs.107 The Drigung sent an emissary to the Mongols in 1219 and their patronage was formalised in 1229 when Ögedai Khan recognised the Drigung—and the Phagmo grupa—as the religious au- thorities in Western Tibet.108 The Drigung maintained their supremacy there until the temporary eclipse of their fortunes following the Sakya takeover of Guge in the late 1270s.109

Opening the Pilgrimage

Despite the Drigung’s dominance in the region, the highly significant ritual act of ‘opening’ the pilgrimage circumambulation route around Tise was attributed to the Drukpa yogin Götsangpa.110 This was a role he also filled at Tsari, one of the other great sites of Cakrasamvara.111 But he is particularly celebrated as the ‘opener’ of Tise,112 and also for his visionary identification of various western Himalayan sites as manifestations of the pithas and other territorial divisions of

104 Vitali (1996: 403–414). Further miracles are described in the Tise Lo rygus (1895), see Huber & Rigzin (1999: 145). 105 Vitali (1996: 407). 106 Vitali (1996: 400, 407, 410–411, n. 706 422–423); also see Sperling (1987: 33–53). 107 On Mongol-Kargyu ties, see Sperling (1990). 108 Vitali (1996: 403, 415, 418). 109 Vitali (1996: 411–413, 422–423). 110 Götsangpa is also remembered as founding the Stodpa Kargyu, a Drukpa sub-sect which gave rise to a number of other sub-sects; see Smith (2001: 45). 111 Huber (1999: 69). 112 Tibetan sacred sites may have more than one opener; see Buffetrille (2014). 302 chapter 11 the Cakrasamvara mandala. Among these were Tīrthapuri as well as Jalandhara and its Jvalamukhi shrine.113 As we have seen, the route he travelled was a specifically Tantric path, one including significant sources of alchemical agents that were of importance to Tantric, rather than specifically Hindu or Buddhist practice. According to the Tise dkar chag, Götsangpa’s discovery of the Tise circum- ambulation route occurred in a series of acts in which he was guided by deities taking form as animals or birds. These merged with rocks when their role was over. The entrance doors to the site were revealed by a dakini in the form of a wild yak, another significant site was revealed by (the deity) Mahakala in the form of a raven, and finally the Drolma la (sGrolmala), the pass over the highest point on the route, was revealed to him by 21 blue wolves, considered emana- tions of (the goddess) Tara. Yet the same text records that over a century earlier Milarepa and his Bön opponent had travelled around the mountain by that route.114 It appears that while the ‘opening’ of a pilgrimage has numerous levels of emic meaning in terms of the individual’s relationship with the sacred site and its manifest deities, in etic terms it represents the process of allowing access to the site for ordinary—i.e., non-Tantric—pilgrims. It thus marks the transformation, or ‘taming’ of the place from one too powerful for non-Tantrics to encounter, into one that enables the householder-pilgrim to partake of the site’s sacred power. Why this role is attributed to the Drukpa Götsangpa and not to a practitioner from the dominant Kargyu sect there, the Drigung, is unclear. Could it represent Götsangpa’s efforts to obtain popular support from the laity in the face of the Drigung’s royal patronage? We have noted accounts of earlier circumambulation of Lake Mapham, from this time on however, the central ritual observance of Buddhist pilgrimage to the site involved the circumambulation of Tise. Thus Götsangpa’s opening of the route signals a transference of focus from lake to mountain, marking the establishment of new understandings of mountains over Western Himalayan traditions of sacred lakes.115 In the wider perspective, the establishment of a profusion of what are now considered sects and sub-sects of the Kargyu tradition was a transitory or even concessional stage in the transformation from

113 While he recognised these sites as containing manifestations of the Śaivite deity Maheś- vara, he understood that deity as having been subjugated to the Buddhist deity. On this context of Götsangpa’s travels, see Huber (2008: 101–109). 114 Huber and Rigzin (1999: 134–136). 115 A gender aspect may also be read, with the focus shifting from the (female) lake to the (male) mountain. tibet’s tise 303 un-systemised communities of followers around a charismatic renouncer to monastic centres under local aristocratic and royal or ‘state’ patronage. In the post-Yeshe Ö reform environment, those communities that took on the new orthodox requirements and established monastic centres were best positioned to gain that patronage. But in making that transition, the Kagyu were able to retain the core elements of earlier identities involving esoteric or sexualised rituals and traditions of itinerancy and congregation in caves, mountains, and other remote and geomantically powerful places. Having established central monasteries on the approved model, those of their followers who were drawn to the renunciate traditions were, after adopting what must have often been a very thin veneer of monkhood, able to continue such practices within a now authorised framework. Again we see the ability of renunciation as a category to survive and prosper through a continual process of reinvention and response to changing circumstances. We also see that the category of renunciate in this period, as throughout, sanctioned violence. ‘Liberation’, or ritual killing, was one aspect that was prominent in the sources, but we may also identify sectarian leaders such as Lama Zhang and Ripa Nagpo116 who led military forces, or at least armed bands during periods of civil disturbance and political vacuum. We will also see that royal patronage for particular religious factions could draw local rulers into involvement in sectarian conflicts. The transformation of the sacred geography of the Tise/Mapham region from a local focus on the lake to a regional focus on the mountain must be located in the wider context of these social and political processes.

116 On whom see Vitali (1996: n. 705 422); Vitali refers to Ripa Nagpo as “the epitome of the ‘bKa.brgyud.pa of the thirteenth century. He lived the life of a hermit, developed mystical powers, and concerned himself with politics.” When the troubles erupted in central Tibet in 1290, he returned from Tise to lead a rebel army. chapter 12 Buddhacising the Mountain

Introduction

We saw that the earliest Buddhist presence in the Tise/Mapham region was probably generated by the tendency for esoteric practitioners to seek out re- mote and auspicious locations for their practice rather than any specific identi- fication of the region with the lake visited by the Buddha or any other textually- based understanding of it as within Buddhist sacred geography. None-the-less, those renunciates required patronage and the process of transforming a remote location appropriate to esoteric practice due to its sanctity in local culture into a textually-sanctioned Buddhist sacred landscape required a specific agency. In this case the agents were the local rulers, who offered support for Buddhist monastic and renunciate activities. Transformation also required explanatory conceptual devices. While the conversion of Ngari to a Buddhacised realm was a lengthy and multi-layered process involving the ritual conversion of numerous local people, sites, and deities, the historically dominant imaginary construction involved the instal- lation of the mandala of the Tantric deity Demchok onto the landscape. The myth of Milarepa winning Tise from the Bönpo may be better known, but his activities were primarily significant in the context of revealing the mountain as the abode of Demchok, his tutelary deity. His victory thus restored the cosmic order. The process of (re)installing Demchok ultimately relied upon the vision- ary authority of certain renunciates of the Kargyu order and the attribution to them of the power to read the landscape with the understanding of the fully- enlightened being. Toni Huber has pointed out that;

all traditional Tibetan [literary] genres depend upon and invoke two principal sources of authority to identify sacred sites. The main type is visionary authority, based upon the “pure visions” (dag-snang) of partic- ular landscapes experienced by Tantric lamas in altered states of con- sciousness, during meditation, dreams, and so forth … The second type is prophetic authority, in the form of citation and analysis of “prophe- cies” (lung-bstan) reputedly made by Buddhas, deities, recognised saints or highly realised lamas about the identity of particular places.1

1 Huber (1997: 306).

© koninklijke brill nv, leiden, 2015 | doi: 10.1163/9789004306189_015 buddhacising the mountain 305

Both such authorities were manifest in the sacralisation of Tise; the prophe- tic authority of Naropa and Marpa validating Milarepa’s role at Tise and the visionary authority of later Kargyupas that validated the understanding of Tise as the abode of the tutelary deity of their sect through the citation of prophecies said to have been made by the Buddha. Having discussed the religio-political aspects of the phyidar, the mythology surrounding Milarepa, and the move- ment of renunciate groups into Ngari in the previous chapter, we now look at the contested process by which the mountain was constructed as a Buddhist one.

The Cakrasamvara Tantra

Śambara/Saṃvara is frequently mentioned in the Rgveda, where he appears as a mountain-dwelling chief (or deity) whose forts are destroyed by Indra.2 But Cakrasamvara only emerges as a Buddhist deity in late Indic Tantric formula- tions as expressed in the Cakrasamvara Tantra. This is, according to a study by David Gray;

a Buddhist ritual text composed in India during the eighth century by an unknown author or group of authors …, developed in a non-monastic setting, and composed via the active appropriation of elements of [land- scape and3] both text and practice belonging to non-Buddhist groups, most notably the Kāpālikas.4

2 On Śambara/Saṃvara in the Vedic literature, see Parpola (2002: 261–262, 272–279); also see Parpola (1988: 210–211, 215–217, 261–263). In these and other works Parpola proposes links to the later Buddhist deity Cakrasamvara. Davidson (2002: 214), while also noting that a deity of that name is mentioned in the Arthasastra, sees no evidence for continuity. Such recurrence of names is common but the processes by which they were recycled is unclear. Did they still retain some cultural currency that gave meaning to their revival? Were they, like toponyms, part of a stock of “free-floating” spiritual indicators and symbols available to whoever needed them? 3 Gray (2006: 306–307). 4 Gray (2007: xv). Where not otherwise indicated, I have relied on this outstanding study. On this text, also see Gray (2001); Loseries (2011). Kazi Dawa-Samdup’s Shrīchakrasambhara Tantra: A Buddhist Tantra (1918. Edited by Ar- thur Avalon: Calcutta/London: Thaker, Spinks, & co), is not the actual Tantra of this name but a series of associated ritual texts. 306 chapter 12

As is the case with other works from the Mahāyoga (“Highest Yoga”) Tantra genre, considerable intertextuality with Śaivite Tantras may be demonstrated.5 The text actually centres on the worship of Cakrasamvara as Demchok Heruka, a fearsome deity whose form is stated in the text to be that of the Hindu Bhairava (a wrathful emanation/form of Śiva).6 The text clearly reflects the eso- teric practices of the cemetery dwelling renunciates of the period whose rituals were applicable to deities across boundaries of religious systems. As with other texts of this genre, the Cakrasamvara Tantra contains considerable material concerning the gaining of magical powers through transgressive rituals, con- sort practice involving the cultivation and sacramental consumption of sexual fluids, and the acquisition of magical powers of the type that, as we have seen, are described in a series of texts dating back to the Atharvaveda. The social setting of the Cakrasamvara Tantra is that of the itinerant renun- ciates gathering in isolated auspicious places and forming communities around a charismatic guru. The text informs the practitioner how to mentally generate or to recognise the dakinis and yoginis. Finding and subjugating these female spirits “who produce all powers” enabled the practitioner to gain magical power (siddhi) and as the text indicated those beings could be found at “the seats of yoga”, or in places where “the heroes who uphold great yoga dwell, even if it is in a land of caṇḍālas [outcastes] and barbarians.” While these female spirits could be found in every land, they were said to be abundant at each of the 24 sacred sites of Cakrasamvara, which were listed in the text.7 These sacred sites include four (Pulliramalaya, Jalandhara, Oddiyana, and Abuda) classified as pithas on the common model discussed in Chapter Four, along with 21 other sites classified under nine categories of “subsidiary seat”, “meeting place”, “charnel ground”, and so on.8 Among them were Kamarupa (that site still active in West Bengal?), Kuluta (Kulu region?), and Himalaya. While the latter is obviously the mountain range, it is again unclear whether a specific site in that region or the entire range is meant. Gray notes that the list ends with “and so forth”, which he points out, “… is significant, for it made possible the identification of alternate pilgrimage places as the tradition was transmitted to Nepal and the Himalayan regions, Tibet and Mongolia.”9

5 The extent to which Buddhist Tantras rely on material taken from Hindu texts (and vice versa) is highlighted in Sanderson (1994) and Sanderson (2001: n. 50 38 [re material from the Śaiva Vidyāpīṭha into the Buddhist Śaṃvara cycle], 41–47); but see Davidson (2002: n. 105 386). 6 Davidson (2002: 214) traces Heruka to a local Assamese “ghost or cemetery deity.” 7 Gray (2007: 266, 291, 329–337). 8 Gray (2007: 67). 9 Gray (2007: 329–333). Also see Gray (2006: 308–309), which notes that the text implicitly acknowledged that many of the actual sites were beyond Buddhist control. buddhacising the mountain 307

This was the case with its transmission to Tibet. As there was a Sanskrit origi- nal of the Cakrasamvara Tantra it was accepted as an authentic Buddhist work, included in the canonical Kangyur, and utilised by all of the later traditions. First translated in the late 10th century by Rinchen Zangpo (and an Indian pan- dit), and later taught by Atisa, the text exists in a number of versions along with many commentaries, which thus sanctified substantial variations in its inter- pretation.10 Of considerable significance is that the Cakrasamvara Tantra references the cosmogonic myth of the subjugation of Maheśvara (a form of Śiva), by Dem- chok, who then took possession of (the central World-mountain) Sumeru from the Śaivite deity. This well known myth is a version of a stock Indic literary device,11 which while functioning on multiple levels, serves most obviously to claim the supremacy of or to indicate the actual take-over of, the deity of one sect by another, and by implication the triumph of that belief system. Traceable to the influential early 8th century Buddhist Tantra Sarvatathāgata- Tattvasaṃgraha,12 the Mahesvara subjugation myth was greatly expanded in the commentarial literature deriving from the Cakrasamvara Tantra and its associated cycles.13 It is notable that the understanding validating this ver- sion was fully developed in Tibetan only in the late 12th or early 13th cen- tury work of Drakpa Gyaltsen. One of the leading Sakya assimilators of the Tantras, his primary intention was the defence of the validity of Sakya litera- ture.14 In the hands of the Kargyu, however, the Cakrasamvara Tantra was deployed as a device through which to identify cosmological understandings of the central World-mountain and its surrounds with an actual earthly site, or sites. Significantly the claim was actually implicit, in that the identifications were with “the mountain prophesised by the Buddha” rather than specifically to Meru/Sumeru.15 Different cycles of myth were thus not fully assimilated or

10 On the versions and related literature see Gray (2007: 15–20); also see Hartzell (1997: 31–32). 11 On which see Davidson (1991: 197–233); also see Davidson (1995: 547–555); Snellgrove (1987: 134–141); Buffetrille (1994); Buffetrille (1997; esp., 89). The accounts of Durga’s triumph over the buffalo-demon, and the Buddha’s triumph over Mara are other prominent examples of this device. A number of other examples of Buddhist assimilation of Hindu and local deity formulations are discussed in Gray (2005: 45–69). 12 Davidson (1991: 198–202). 13 See Gray (2007: 40–54). 14 Davidson (1991: 204). 15 An explicit Meru=Kailas equation is difficult to locate in pre-modern sources, but the available range of toponymical synonyms allowed this to be implied. See, for an early 308 chapter 12 reconciled, something characteristic of later Tibetan understandings of sacred sites where multiple sacralities co-exist. The visionary installation of the mandala of Cakrasamvara/Demchok onto the landscape of Tise, understood as the mountain said to have been prophe- sised by the Buddha in eclectic texts such as the Avatamsakasutra, effectively identified the protective deity of the Kargyu as the deity resident at the cen- tre of the world. It was thus, implicitly, the supreme divine principle. While the Kargyu had been driven by both famine and inclination to the periphery of the Tibetan cultural world, this transfer of cosmic to actual geography repre- sented a claim—an extraordinarily ambitious claim—to the centrality of that periphery. The Kargyu, in effect, claimed that the territory in which they were religiously paramount was nothing less than the centre of the world, the earthly location of the world of the gods. They now lived, it seemed, within the man- dala of their protective deity. Yet while ambitious, the Kargyu’s claim did not represent a radical departure from existing cultural understandings of a people, a territory, and a deity identi- fied with a local territorial feature. In addition, there was at least some textual precedent for the claim. In the late 11th or early 12th century the Kālachakra Tantra had situated a cosmological centre outside of India, albeit that centre was located in the mythological land of Shambala.16 Thus the Kargyu devel- oped rather than initiated a new line of understanding in the application of sacred geography. Crucial to the validation of the Kargyu claim was the con- cept of the mandala, and before discussing the basis of their claim, we turn to the relevance of that concept.

Mandalas: The Development of an Organising Principle17

The term mandala first appears in the Rgveda meaning ‘book’ or ‘chapter’, and it is thus an organising principle separating one section of the text from another. By the time of the Arthasastra (c. 150ce18), the term was used to indi-

example of this, van der Kuijp (2004: 328), who cites the Kargyupa, Darma Gyaltsen (1227–1305) as identifying Tise as the highest mountain of Jambudvipa and the mountain from which four rivers descend; Meru is apparently not mentioned. 16 Davidson (2005: 281–282). 17 My understanding of the mandala benefitted from discussion with the late Oliver Winrow, a soas doctoral student investigating precedents of the Cakrasamvara mandala in the “explicit logic” of the Rajmandala of the Arthasastra. 18 Trautman (1971: 184 and passim); also see Gray (2007: n. 120 39). buddhacising the mountain 309 cate a political or administrative unit,19 and later mandalic formulations retain that inherent structural character. In that they illustrate an understanding of sanctified space under the power of a single divine, or divinely ordained, prin- ciple, the mandala has conceptual roots in the Vedic sacrifices,20 performed on ground ritually demarcated to exclude the profane and the impure from the sacrificial site. The Vedic deities, generally in order of hierarchy or perhaps more precisely precedent, were invited onto that sanctified ritual space to par- take of the sacrifice. This assembly of deities could be expressed artistically, conceptually systemised as hierarchically situated around a central figure. In various forms that underlying principle of a formation of (classically four or eight) lesser entities around the central presiding figure of power came to be termed a “mandala.”21 As indicated by the Kuru claims regarding Kuruksetra, the concept of demar- cating a (moveable) power centre was long extant in Indic culture. It mani- fested in various spheres, in the rite of circumambulation (Tib: chos skor), for example, and in early theories of kingship and the state. Thus the Arthasas- tra situates the ideal royal palace at the centre of the kingdom. That model of the royal palace corresponded on the earthly level to the concept of the heavenly palace of a deity. Both king and deity were envisaged as central fig- ures whose power radiated outwards from the centre towards the surrounding impure and/or hostile realms. There was also a vertical linkage, for one aspect of what was taking shape here was the articulation of the king as a divinely sanc- tioned element, an intermediary between the earthly and the higher worlds whose rule replicated the heavenly remit of the kingdom’s presiding deity. The mandala was thus implicitly multi-dimensional, conceptually allowing vari- ants such as the later understanding that the individual body could be the site of a mandala. Mandala formations of power were, like the Chinese model of a supreme central mountain surrounded by the four marchmounts of each cardinal direc- tion, infinitely replicable. Each element, for example each peripheral deity in the mandala of a greater deity, could be located at the centre of its own man- dala in an overlapping and interconnected schema. The concept proved widely applicable, power centres such as temples for example, could be depicted as the centre of their mandala of spiritual and social power and influence or as

19 White (2003, n. 3 124, citing Gupta and Gombrich: 1986, n. 17 130) referring to the title of the 6th book of the Arthasastra. 20 “Like the Vedic sacrificial alter of which it is a streamlined form.”; White (2001: 9). On the origin of the mandala also see Luzanits (2008); Samuel (2008: 224–227). 21 Kramrish (1946). 310 chapter 12 peripheral in the mandala of a temple of greater regional or structural power. Given the ideal model provided, replication of the higher models and forms by lesser powers was presumed (a model that emerged in widely differing forms in sub-continental cultures22). The conceptual and imaginal malleability of the mandala model enabled it to become a (indeed, perhaps the) fundamental cultural referent through- out the Indic and subsequently Buddhist world, and an understanding of its subtleties and organic nature is crucial to the history of the Kailas pilgrimage. Discussing the breadth of its application, Huber observes that:

[i]n various forms and applications throughout Asia, maṇḍala have ser- ved as archetypes of the ideal city, models of the cosmos, blueprints for centers of royal power, templates for the operation of polities, networks for the distribution of resources, plans for sacred architecture, represen- tations of the divine mansion or palace and schemes for the distribution of vital energies within the human body.23

While that latter manifestation has been the focus of particular interest in the West, our concern here is with the mandala as a schema for territorial expan- sion by various spiritual power structures; a “conceptual tool for transforming the cosmos.”24 Ultimately it was the understanding that the mandala of a par- ticular deity could be inscribed onto—or in the emic perspective revealed as existing on—a landscape, that was at the heart of the wider sacralisation of Tise/Kailas within the Buddhist world. The disparate elements of the fully-developed mandala concept(s) did not emerge in a synthesised form until at least the latter half of the first millennium ce., with Lo Bue identifying the “earliest known text describing the conven- tionalised maṇḍala as we know it today [as] the Dharma maṇḍala-sūtra by Buddhaguhya (740–802).”25 Ritual application of that formation to the clas- sic Atharvavedic rites of pacification, subjugation, and so on, is indicated in

22 It is found, for example, in the manner in which the British Trade Agencies in Tibet (1904–1947) replicated the institutional model of the Residency in Sikkim, which in turn replicated the structures surrounding the Viceroy; more obviously one thinks of the Sanskritisation model. 23 Huber (1999: 26). On the Buddhist mandala, see Brauen (1997). 24 The phrase is Oliver Winrow’s. 25 Lo Bue (1988: 177). But see Luczanits (2008: 117) who points out that a circular formulation is not explicit in that text, and that later depictions of mandalas preserved in the caves of Dunhuang still held to the earlier square formation. buddhacising the mountain 311 that text,26 suggesting its usage in the Tantric circles that kept such rites. That understanding is confirmed by the predominance of representations of Tantric deities in the earliest mandala paintings (Tib: thangkas), and it is notable that most of these early depictions are of Cakrasamvara.27 Studies of Tantric origins are complicated by the fact that while many if not most of its constituent elements were manifest in earlier periods, the available evidence indicates that synthesisation of those “proto-Tantric” elements into a comprehensive system was a comparatively late and apparently rapid devel- opment. Davidson has convincingly located the Tantric Buddhist synthesis in a socio-political context, concluding that it occurred in the seventh century ce., “in the special circumstances of the rise of sāmanta feudalism.”28 Reflecting the structural hierarchies of feudal society,the “defining metaphor for esoteric Bud- dhism [became] that of the monk or practitioner becoming the … Universal Ruler”; “the overlord of a mandala of vassals.”29 Deity, king and Tantric practi- tioner all partook of or merged in the power centre of the mandala, while their vassals—lesser deities, princes or chelas—surrounded them and mediated the central power on its outward radiation. The king’s power over a specific territory was understood to equate to the power there of the deity with whom he was identified. Thus, when the king extended his power over new lands through conquest, so too did the divinity, the heavenly manifestation of the king, incorporate that new territory into its domain. Just as Indic kings commonly allowed lesser rulers to continue to control certain aspects of their kingdom as long as they acknowledged his overlordship, so too was it understood that the greater deity bound local deities to his will in return for which the local deities retained a certain degree of local autonomy. At the practical level of political power and at the level of belief in particular deities, the acceptance of local centres of power within a cultural realm, each replicating the essential structure and form of a central power, is characteristic of the south/southeast Asian “Galactic polity” model identified by Stanley Tambiah.30 This allows that centres of power wax and wane over time in a dynamic and flexible process and as David Gordon White observes, the “royal ideology of ‘galactic polity’ or the ‘exemplary center’ comprising the king, his

26 Lo Bue (1988: 221–224). 27 Luczanits (2008: 112); also see Luczanits (2009). 28 Davidson (2002: 166). 29 Davidson (2002: 2–4). 30 Tambiah (1976: 102–131). 312 chapter 12 deity, and the capital city, has been mediated by the mandala in nearly every premodern Asian political system.”31

Situating the Tise Mandala

The mandala was a device amenable to shifting cultural and historical con- texts and in the early part of the second millennium ce., the Kargyu applied it to the landscape of Tise in a visionary process on the model described in the Cakrasamvara Tantra. As noted, according to this text Demchok, having defeated Mahesvara, took his place on top of Sumeru in union with his consort, Vajravarahi. Surrounding them in the mandala formation were four yoginis and then three concentric rings of eight Bhairava deities, each of these deities and their consorts being associated with one of the 24 sacred places of the tradition (each of which places were thus ultimately sites of Cakrasamvara/Demchok). The periphery of the mandala was protected by eight wrathful goddesses, with the “mandala palace” then surrounded by the eight great charnel grounds— testament to the text’s origins in the cemetery cults of esoteric ritualists.32 In total, therefore, these divine beings made up the 62 deities of the commonest form of the Cakrasamvara mandala.33 Gray describes how, in a manner similar to visualisation practices outlined in other Tantric texts;

The most basic form of meditation practise for this tradition is the cre- ation stage (utpattikrama) practice in which one visualizes oneself as the deity Heruka, surrounded by his retinue in the maṇḍala palace atop the pinnacle of Mount Sumeru.34

31 White (2003: 25). 32 Gray (2007). On the cremation grounds at the periphery of the mandala, see Davidson (2002: 303). 33 For details of the names, positions, sites etc., of the Cakrasamvara mandala, see Tsuda (1978: 218–226); also see Gray (2007). 34 Gray (2007: 55). The Tantra actually refers to “the divine mountain” (i.e.; chapter 48; Gray: 2007: 364), rather than specifically to Sumeru/Meru; a synonym that permits multiple interpretations. Thus the later Kargyu identification refers to Tise/Kailas as “Himavat” or the “divine mountain prophesised by the Buddha”, rather than to Sumeru. It is the commentarial literature, from within a specific tradition of interpretation, that “fixes” such identifications; multiple interpretations are thus possible. buddhacising the mountain 313

More advanced practices such as the higher consecration rituals involve the practitioner taking on the external appearance of Demchok Heruka (ash- covered and wearing bone and skull ornaments), mentally becoming Demchok and taking the centre of the mandala, drawn at the chosen site according to the ordained ritual. The site should be, according to the Tantra, “on mountains, in caves, in groves, on river-banks, or in the primordially established charnel ground.”35 Yet this ritual practice did not in itself establish Tise as the accepted resi- dence of Demchok. Nor is there a single ritual event in which this occurred, or a single individual—a Milarepa or a Götsangpa—to whom the transformation was attributed. The understanding was that the mountain had been taken by Demchok from Mahesvara in cosmic rather than historical time. Thus the cru- cial element in the transformation was the myth of Mahesvara’s subjugation, which legitimated the site as the centre of the mandala,36 not the ritual, which was concerned with the individual seeking the power of the deity. Such rituals were presumably common, yet did not always result in that territory becom- ing accepted as the mandala of whatever divinity was the focus of the ritual. Indeed the rituals were supposed to be carried out in secret (though we might imagine that they did not always pass off unnoticed by residents or travellers). But for their formation to become accepted, the concept needed to be promul- gated and gain popular approval. The myth was the means by which this was articulated, not the ritual. In the emic perspective, the truth was revealed, not established by human action. We need to remember that the same process also occurred around the same time at various other places,37 including two other Kargyu centres with which Tise was associated, Labchi and Tsari. These three mountain sites were understood as representing respectively the Body, Speech and Mind of the deity.38 This was a time of great competition between sects, yet although there were objections to the identification of Tibetan sites with the sacred sites of Indic Buddhism—as will be seen—there was no Buddhist opposition to

35 Gray (2007: 162, and see 165–383 for a translation of the ritual). 36 cf. Ehrhard (1997: 336); “[p]erhaps it is the text that, in the Tibetan cultural sphere, brings forth the maṇḍala and sacred landscape.” 37 Other sites where the Cakrasamvara mandala was installed onto the landscape include Muktinath, on which see Buffetrille (1994 and 2012), and Kathmandu, on which see Decleer (2000: 48). But as noted, the addition of pitha places, charnel grounds and so on in the mandala means that conceptually there are a large number of sites at which Cakrasamvara is considered the ultimate principle. 38 Ricard (1994: 5). 314 chapter 12 the actual claim that these particular places were sites of the Buddhist deity Demchok Heruka. It was the means not the concept that was disputed. This suggests that the transformation of these sites on the periphery of the central Tibetan polity into Buddhist sacred territory was in the interests of Tibetan Buddhism generally, and the emerging Tibetan state more specifically. The support offered to the various Kargyu sects by the rulers of western Tibetan kingdoms certainly indicates that they were favourable to this process. It appears that the acceptance of this understanding was due to the contin- uing weight of sacred geography claims made by a succession of charismatic Kargyupas whose identifications relied upon their visionary authority, but that the support of the local kings and royal institutions cemented this understand- ing in the Tibetan world. The myth of Mahesvara’s subjugation and the myth of Milarepa’s victory over his Bön opponent were thus successful literary strate- gies that could also be taught to the majority of the population through sto- ries and songs, art, dance and other aspects of a systemised cultural produc- tion. This conclusion would support Huber’s thesis of socio-cultural factors as paramount in the transference of sacred geography from India to Tibet.39 An alternative explanation—which seems to lack a grounding in historical evidence but is theoretically possible—is that the ritual practitioner would interiorise the sacred geography outlined in the text. The relevant mandala could then be visualised as mapped onto the practitioner’s body, with cer- tain sites corresponding to particular parts of the body and projected onto a landscape.40 Such interiorisation may form a part of the practitioner’s visual- isation of landscape and identification of related features in a new landscape, but is unlikely to survive over time without a wider support system of pub- licity. David Gray notes that there are no references to such internalisation in the Cakrasamvara Tantra, but also that “the commentators had at their dis- posal a long tradition of sophisticated interpretative strategies that permitted the creative discovery or uncovering of the hidden import of the text.”41 A late 11th—early 12th century version of the text for example, specifically mentions Tibet and China and “legitimates the re-mapping of the sacred sites of the tra- dition to the landscape of Tibet and China.”42

39 Huber (2008: 108–109). 40 Huber (2008: 109), notes that “external reprojection of pīṭha is itself extremely rare in Tibetan discussions of Tantra.” 41 Gray (2007: 71); also see Gray (2007: 697); this was however, a well-known identification, see for an example in the context of Nath-Buddhist interaction, Schaeffer (2002: 521). 42 Gray (2007: 70). buddhacising the mountain 315

When we have accounts of how sites were recognised these always seem to involve not a ritual but visionary revelations or the intuitive understandings of enlightened masters of the tradition. The site is revealed or “remembered”, rather than being created through ritual projections (although the technologies involved in the process by which sacred territory is “found” are not all manifest in texts). But given the paramount significance of patronage and promulgation of such revelation (or projection) in embedding it into a culture in lasting form, this is normally a process rather than an event. Sacred sites (particularly those not connected to the historical founder of a religion), usually become sanctified by the accumulation of sacred associations over time. As A.W. Macdonald pointed out, “[t]he installation of a maṇḍala … takes time. A maṇḍala is not projected in its entirely and all of a sudden into a new site.”43 Nor, as will be seen in the next section, are all aspects of the mandala installation necessarily uncontested.

The (Re)Creation of a Sacred Buddhist Site

We saw that there is no evidence Tise was a sacred Buddhist site prior to the phyidar, or that it was subject to a process of Buddhacisation in which the yullha of the mountain was conceptually converted into a Buddhist protective deity. We do have later references to Mapham being circumambulated in the late 10th and early 11th century by Buddhist rulers from Kunu and the Malla dynasty of western Nepal. Those royal journeys can be contextualised as within the understanding of kings as possessors of those magical powers necessary to venture to sites otherwise open only to renunciates. They were also acknowl- edgements of local sacralities and claims to supremacy over them, and may indicate that the Klu deity Madros’s “conversion” to Buddhism was understood to have already taken place. But the first specific Buddhist association with Tise is Milarepa’s legendary late 11th century triumph over the resident Bön magician. Another century passed before the full-scale influx of Kargyu renunciates into the area began and as noted, Milarepa’s precise role and achievements at Tise are problematic. The legend serves as the Buddhist claim to the mountain as a ‘field’ of Buddhist activity and first appears in a very basic form a century or more after the event. That intervening century was one in which sectarian formation and contestation over the identification and sacralisation of the Tise region were interwoven processes, of which we know little.

43 Macdonald (1997: viii). 316 chapter 12

The identification of Tise as the abode of a Buddhist deity seems only to date to the influx of Kargyu practitioners into the region in the 12th–13th century and to be framed in the context of Kargyu efforts to root their traditions and author- ity there. As noted, they also grounded their tradition in two other mountain sites, Labchi and Tsari, and as will be seen the greater prominence of Tise over those sites today is the result of Indic historical developments. Tise was prob- ably a blank canvas prior to that time (unlike the lake), and not specifically a sacred space in any body of knowledge. Its distinct appearance is an argument against that, and it is possible that the Yarlung take-over of the region saw the expulsion/migration of a population group—or its elites—that did hold the mountain sacred. As we will see in the next chapter that understanding is part of Bönpo beliefs, but their evidence is very much later. Strictly speaking, on the basis of reliable sources we can only date Tise as a sacred Tibetan moun- tain from around the beginning of the 13th century (which is consistent with Vitali’s dating of around 1215 as the beginning of the pilgrimage). The legend of Milarepa served to authorise the Buddhist project and to root it and indigenise it in a “historical” figure who was already fundamental to Kargyu origin narratives and identity. The legend of Demchok Heruka, his tri- umph over the Śaivite deity Mahesvara and his right of abode on Tise then advanced that project. It implied divine purpose behind Milarepa’s activities and thus his enlightened awareness of the supreme sacrality of the site, while at the same time situating the tutelary deity of the Kargyu sect in cosmic— and thus eternal—time. The unfolding historical construction was at once religious, in subjugating the Bön claim to the site, religio-culturalist,44 in sub- jugating an Indic Brahmanical presence at the site, sectarian, in subjugating it to the Kargyu divinity, and (what we might call) modernist, in subjugating the Klu lake divinity within the mandala of the Tise-dwelling Demchok. Authority was rooted in divine revelation and charismatic authority and enhanced by cul- turalist and modernist sentiments as well as the dramatic appeal of a colourful and culturally appropriate narrative. That authority was also enhanced by royal patronage which enabled the promulgation of the new construction through monastic and popular performance, text, and ritual. It is Götsangpa’s opening of the circumambulation route around the year 1215 that is the most significant single act in this narrative process, in that it sym- bolises the transfer of primary regional sacral focus from Madros dwelling in Lake Mapham to a Buddhist deity resident in/on/above, the mountain, which

44 The term is coined to indicate what would be in a later period “religio-nationalist”; i.e., it refers to sentiments of exclusion towards “Others”. buddhacising the mountain 317 was then manifest as the “divine embodiment.”45 In other words, Götsangpa’s action marks the final subjugation of local territorial forces by those of Bud- dhism. Whether this also represents the triumph of a gender model, an indige- nous Tibetan model of mountain deities, or an Indic sacred model is unclear. Lake Mapham is considered female, but its presiding deity the Klu Madros is apparently male so the gender aspect is blurred. The Indic model of a central World-mountain is introduced in a layered form enabling multiple interpreta- tions but whether this represents an innovation or an adaptation of indigenous understandings of mountains remains unclear. It is safest to conclude, if only on the basis of its success, that the construction satisfied a significant range of socio-religious perspectives among the diverse population groups in the region, probably including those from Indic, central Tibetan, Central Asian, and even Western Himalayan Cultural Complex population groups. The Kargyu’s projection of the mandala of Demchok onto the Tise-centred landscape has certain implications in regard to historical distinctions between Tibetan and Indic cultures. In the sense that it represented a statement of indigenous identity it may well have suited the interests of the Ngari ruling classes, who were threatened by forces from the south and west. The primary rite of Indic religious practitioners was the ritual bath, a sacred site without a water source was almost unimaginable in an Indic context. Thus Naga/Klu lakes were prominent in Western Himalayan sacrality and Mapham was and at least ritually still is the primary focus for Indic visitors to the region. But the Tibetan relationship with mountains did not require a water source, while Bud- dhist literature contains numerous accounts of the Buddha and his followers overcoming Klu. Thus in prioritising the mountain as a sacred site the Tibetan Buddhists relegated both Hindu and local traditions to a focus on sites that were given lesser status, laying the groundwork for a greater civilisational divide after the closeness of the phyidar.46 The Kargyu projection of the mandala of their tutelary deity onto Tise was not an uncontested process. It was accompanied by certain integral claims that soon became the subject of criticism, most notably by the Sakya Pandita whose critique has attracted some academic discussion.47 In particular the Kargyu claimed that Tise was not just the site of Demchok’s residence but was the site of the Tantric pitha “Himalaya” and even the “snow mountain”

45 Huber (1999: 22). 46 Whether this deliberate cultural, geographic, and linguistic divide implies proto-national- ism might be considered. 47 See Broido (1987); Jackson (1994); and Mayer (1997: 79–105). For analysis and discussion of the sacred geography aspect of the Sakya Pandita’s critique, see Huber (2003: 392–424). 318 chapter 12 said to be referred to in the Abdhidharmakosa, with Lake Mapham equating to the Anavatapta lake of that text.48 The articulation of this equation was possible through revelation and interpretation within the vague and imprecise definitions of toponyms such as “Himalaya” as well as textual variations and additional and ongoing commentaries on them.49 The original source of these claims is unclear, for they were attributed to Marpa and Naropa in retrospective accounts of Milarepa’s sojourn at Tise.50 But that they attracted the Sakya Pandita’s criticisms around 1232 suggests they emerged from the circle of meditators sent out by Jigten Gönpo in the last decade of the 12th century, and if there was a single agent behind these claims he seems the likeliest candidate. Interestingly, Jigten Gönpo referred to the Buddha as travelling “to the banks of the Anavatapta and Manasarovar lakes”,51 thus equating Rakas Tal with Anavatapta (a point on which there may have been no consensus). The shifting identities of toponyms were apparently as fluid then as now, and certainly enabled considerable flexibility in such identifications. The Sakya Pandita’s criticisms appear in his “Discrimination of the Three Vows” (sDom-gsum rab-dbye), a work in which the modern reader can recog- nise a reasoned, logical critique of legalistic precision. The Sakya hierarch acknowledged the validity of the Tantric path of caryā—the “crazy wisdom” phase of those practitioners who wandered among cemeteries and pitha sites.52

48 For details, see Huber (2003: 392–424); also see Huber (1999: esp., 41–77). Further claims were also made as to the identification of Labchi and Tsari with sacred sites in India, which will not be considered here. 49 Tibetans also had access to Indic literature including the Epics and the Meghaduta, on the latter, see Smith (2001: n. 663 319). 50 Huber (2003: 392) cites a statement attributed to Milarepa; “This snow mountain [Ti-se] and the holy lakes, all three [sic! See n. 66, below], are a meditation place which was prophesied by the Buddha. And, if one does not respond to those who criticise them, then not only will these slanderers heap up sins but the greatness of this meditation place will also be violated!” The source is the 15th century biography, or hagiography, by Gtsang smyon Heruka, who considered himself an incarnation of Milarepa and who initiated a genre of Kargyu biographies; (Smith 2001: 61). That Milarepa should counsel the need to defend the identification against critics is surely a later invention, for the Sakya Pandita’s criticisms were made long after Milarepa’s passing. 51 “The Buddha … came to the great glacier mountain Ti-se (Ti-se Ri-Bo Gangs-chen) to the banks of the Anavatapta and Manasarovar lakes, and preached many sūtra texts …”; Martin (2001: 37), citing The Collected Writings (Gsung-’bum) of ’Bri-gung Cho-rje ’Jig-rten-mgon-po Rin-chen-dpal [Jigten Gönpo], (5.vols), Khangsar Tulku, New Delhi, 1969. 52 On the Kargyu as a sect “in which practical accomplishment in religion is given priority over interpretation and study”, see Martin (1996: n. 4 25, 33). buddhacising the mountain 319

But he emphasised that this phase could be properly undertaken only after monastic training, without which such practices were “useless”; an under- standing that was becoming if it was not already, fundamental to the Tibetan Buddhist clerical authorities. The Sakya scholar then carefully differentiated (indeed, much more carefully than many modern writers), the views of poets and the views of scholars. Poets he observed, might use “for vastness the meta- phor of the sky”, but “scholars could hardly be satisfied with descriptions that are not in accord with reality.”53 He also reiterated that it was forbidden to ven- ture to divine lands without Tantric initiation, that understanding fundamental to Indic conceptions of the sacred Himalayan realms. The Pandita then focussed on specific aspects of the identification of Tise with the sacred sites of Indian Buddhism for which the Kargyu claimed tex- tual sanction. He pointed out that the cosmological descriptions given in the texts did not match the appearance of Tise. The texts described elephants there, a fabulous Jambu tree, and jewelled lake-beds, while in reality none of these existed at Tise. Similarly he noted that the Kargyu description of the four rivers (the Ganges, Sita, Sindhu, and Paksu), and their circulation of the Lake Anavat- apta was far removed from the actual geography of Mapham/Manasarovar and its surrounds. Set within a wider critique of Kargyu interpretations of Buddhism, the rig- orous realism of the Sakya Pandita’s analysis of the claims for Tise stands out dramatically from any other account of the region. While otherwise swathed in visions, myths, and poetics by Asian and Western commentators, the sacred site was skilfully deconstructed by the Sakyapa hierarch, whose analytical approach pointed to a road rarely taken in later Tibetan historiography, but which indi- cates that such critical approaches are not the exclusive preserve of Western scholars. It is difficult to separate his critique from the framework of the contemporary struggle between the Sakya and Kargyu sects (which may also have had a per- sonal aspect54), and in the wider context from the beginning of the encounter with the looming power of the Mongols. But Huber locates the question of authenticity at the heart of the dispute, and certainly the empiricism of the critique is consistent with that conclusion.55

53 Here I rely on the translation by Huber (2003: 396–398), which is the source for all quotations from the “Discrimination of the Three Vows”. 54 Mayer (1997: n. 4 97–98) draws attention to this, on which see Shakabpa (2010: 232–233). 55 However ill-informed, as Robert Mayer has pointed out, the Pandita may actually have been as to the nature of “authentic” and “inauthentic” Indian Buddhist texts, see Mayer (1997: 91–96). 320 chapter 12

In seeking to defend their identifications, Kargyu apologists faced the prob- lem that they generally relied on Indic texts that did not employ a consis- tent schema in sacred geography nor maintain a single vision of the heavenly realms. In effect, they inherited not a single tradition of Indic beliefs, but a wide selection of different Indic bodies of knowledge concerning the site. This was further complicated by the existence of different versions of the texts, by the translation process, and by the plethora of commentarial literature, which resulted in a wide and often incompatible range of possible sources and read- ings of them, which the Tibetans rarely referenced precisely in such debates. Given that most Indic references to Kailas were in the Hindu Puranas, and that there was a paucity of references to Tise/Kailas in early Buddhist texts, the Kargyu defenders tended to primarily cite the Abdhidharmakosa, with all its complicating factors. Versions of this text did locate in the north of Jambudvipa (continent) an unnamed “snow mountain beyond the nine black mountains”, but the latter set were another shifting concept, and no consistent geography of what was intended to be a mythical landscape existed from which to draw firm conclusions. The visionary authority of Marpa, Milarepa, and other luminaries of their lin- eage was invoked in support of the Kargyu formulation. Indeed that remained their strongest validation for the application of Buddhist logic to mythologi- cal constructions of sacred geography, even that of a single text, was inher- ently problematic for their claim. The four rivers of the Abdhidharma lit- erature were particularly difficult to locate in reality. Of the Ganges it was claimed;

[T]he eastward flowing Gangā originally was a spring named mThong- ba Ran-grol which flowed out of a valley that lay in an easterly direc- tion from the lake itself. And later, it flowed westward from a mountain called sDul-chu, which resembles the mouth of an elephant, in the upper Gu-ge region in the west, after cutting through the middle of the lake itself.56

The Sindhu, associated with a “peacock-mouth” mountain, was that which became the Indus, after flowing from the north of the lake to the south, while the Paksu flowed to the east from the west of the lake, was associated with a mountain resembling a horse and became the Tsangpo-Brahmaputra. The Sita was said to flow north from the south of the lake from a mountain resembling

56 Huber (2003: 404). buddhacising the mountain 321 a lion and to flow into the northern ocean after crossing Ladakh, Baltistan, and Turkestan.57 Coming from a sect whose members were widely travelled in the region, this hydrological sacred geography represented an extremely strained effort to reconcile Indic text with Tibetan reality. The cosmological model required that if Mapham was Anavatapta (contra Jigten Gönpo!), it had to be the source of the Ganges and thus the “Gangā” was here identified with the Sutlej while the Sita imitated the reality of the Indus before its journey to the mythical northern sea. Attempts to transplant a particular body of Indic knowledge to Tibet also faced the problem that Indic texts identified only one lake in the sacred realm. The inhabitants of the Tise region were obviously well aware of there being two lakes on the plain to the south of the mountain, and on occasion they recognised three.58 As we have seen, Jigten Gönpo did refer to the two lakes as Anavatapta and Manasarovar on one occasion, but this differentiation was not predominant. It may be that the otherwise inexplicable inauspiciousness of Rakas Tal is a result of this problem, with Mapham/Manasarovar (in a not necessarily uncontested process that leaves occasional indications to the contrary), identified with Anavatapta, thus leaving Rakas Tal as something of an embarrassment to the claimed identity of Tise with the heavenly realm. There were other fractures visible too, though it seems that no-one chose to enquire too deeply into such issues. Was Tise Himavat (which translated as “snow mountains”), or Kailas (colloquially used, as noted, to mean “snow mountains”), or was it, or Meru, a “snow mountain” or some combination of these? Resolution was impossible. Aside from deliberately encrypted knowl- edge, different social groups could maintain different bodies of knowledge, with traders, for example, surely aware of the source of the Ganges and the routes of the Sutlej and the Indus. But while they allowed different percep-

57 Huber (2003: 404). 58 Vitali (1996: 154) describes the “Zhang.zhung-pa appraisal” of there being three lakes at the foot of Tise; “two Mu lakes … Mu.le.khyud mtsho … and Dug.mtsho Mu.le’i’, the former being Mapham and the latter Rakas Tal, with the third being the “rGyud lake in Gung.rgyud dngul.mtsho in Hor.ba”; Vitali (n. 205 154) also refers to four lakes in the Bön tradition: “Ma.phangmo.bya g.yu.mtsho, La.ngag gser.mtsho, Gung.chung dngul.mtsho, [and] Zom.shang lcags.mtsho.” Interestingly, in the context of the later inauspiciousness of Rakas Tal, Vitali (n. 208 155) also cites the Tise lo rgyus; “Formerly, the mighty dpal Ye.shes mGon.po.beng and [his] consort inhabited the palace of the black poisonous lake in the country of the srin.po-s, sited in the land of Lang.ka Pu.rang of srin.po Lang.ka mGrin.bcu (Dasagriva).” 322 chapter 12 tions of the site by those of different levels of enlightenment, the understanding that the heavenly realms of Anavatapta could only be reached by Hero-kings or great renunciates and Tantrics with magic powers was an understanding the Kargyu could not support. We must suspect, with Huber, that the economic benefits to the Kargyu from opening up the pilgrimage to all-comers were con- siderable, both in terms of offerings to renunciates, and the local economy generally.59 What is clear is that while the Kargyu always remained sensitive to the criti- cism by one of the outstanding minds of Tibetan history, their understandings became hegemonic in Tibetan society. The obvious contradictions were not sufficiently powerful in themselves to overcome the weight of visionary author- ity, elite patronage, and a cultural background in which numerous layers of belief existed. The financial benefits of pilgrimage must have been obvious from the first, and there were to be less tangible but powerful benefits to the later Tibetan state after the Tise pilgrimage was recognised at the highest level in 1616, when it was undertaken by the Panchen Lama. Before discussing Tise as a state-approved ritual and the continuing sacred geography exploration by the Kargyu to the west of the Tise-Mapham region however, we shall briefly summarise the later period.

From Sect to State

Although the various Kargyu sub-sects were firmly established in Ngari dur- ing the first decades of the 1200s, religious supremacy there was to be con- tested over the ensuing centuries. While control over trade routes was always a significant factor,60 sects patronised by and allied with the rulers of the vari-

59 Huber (2003: 406) observes that this also “reflects in part a collapsing together of the categories of Tantric “wandering” and non-Tantric circumambulation in the Tibetan pīṭha tradition.” 60 While trade was certainly of considerable local importance, it is difficult to reconstruct local economic history. Various regional trade routes intersected in western Tibet, but trade was essentially in private hands, smuggling was rife and even in the British period statistics were unreliable. Historically, desertification in the mid-second millennium al- tered economic patterns, presumably stimulating the growth of trade and pilgrimage as alternative sources of income. But the location of population centres along the Sutlej (a trade route when iced-up in winter, when agricultural labour is freed for trading expe- ditions), suggests that these sites were influenced by efforts to control trade. See for the economic background, van Spengen (2000); also see Mittal (1985). buddhacising the mountain 323 ous principalities were an integral part of wider political conflicts, which also involved Mongol leaders allying with various factions. Mongol power installed the nephew of the Sakya Pandita as the ruler of central Tibet in the mid-13th century, and the Sakya position in Ngari was thus bolstered. The Guge-Purang Genealogies are silent concerning the 1277–1372 period, which Vitali concludes saw the Sakyapas take the region (under Mongol suzerainty), from the previ- ously dominant Kargyu.61 None-the-less, a further wave of Drigung Kargyu meditators were sent to Tise, Labchi and Tsari in 1305,62 testifying to the continuing vitality of their tradition there. This was the period in which Purang-Yartse extended its rule across the Himalayas at least as far as Gangotri,63 and that communications with India remained open is indicated by the fact that between 1336 and 1354 pandits from eastern and western India were invited to the Yartse principality, though not, apparently to Purang.64 By around this time, however, the Mongol Yuan dynasty was collapsing in China and in central Tibet Sakya’s control ended in 1354 with power taken by the aristocratic lay monk, Changchub Gyaltsen,65 a patron of the Kargyu. The Guge-Purang region was then able to enjoy something of a resurgence of its separate identity, free from central Tibetan dominance.66 To the east however, a new sect, the Geluk, arose from a basis in the Kadam order. Its initiator was Tsong Khapa (1357–1419), an austere scholar, and the 1409–1419 founding of Ganden, Drepung and Sera monasteries, along with their success- ful implementation of the incarnation system to establish a lineage that was to become the Dalai Lamas, laid the foundations for their future hegemony. The Gelukpas soon established an influential presence in Ngari.67 Political power in Tibet remained contested but from 1435 until the 17th century the Kargyu sects were favoured by the most powerful forces, notably a series of aristocrats from Tsang. The Kargyu were able to retain their institutional power and enjoy the paramount patronage of these regimes, with the Geluk sect slowly emerg- ing as their main competition after the Sakyapas’ eclipse. But in the absence

61 Vitali (1996: 556–557, and for detail on this period, 414–520, etc); also see Petech (1980: 85–111). 62 Vitali (1996: 411–413). 63 Vitali (1996: 466–467). 64 Vitali (1996: 458–459). 65 On whom see van der Kuijp (1981: 277–327). 66 Vitali (1996: 476–480, 501). 67 Vitali (1996: 414, 514–515). 324 chapter 12 of a strong central power the Ngari kingdoms continued to contest regional supremacy. War might disrupt the sacred site, as it did at the end of the 15th century,68 but there were now significant groups of renunciates from the various Kargyu sects established in the Tise region, along with those of other traditions. There are even references to a Bönpo creating problems there for the Phagmogrupa in the second quarter of the 15th century.69 But having established their religious dominance in the region and enjoying royal patronage and presumably the offerings and respect of pilgrims, the Kargyu began to take increasing advan- tage of new print technologies to celebrate both their beliefs and the heroes of their lineage tradition. The Kargyu embraced songs as a means of spreading their teachings among non-literate peoples but they also appealed to the literate strata of society through textual composition, and from the time of Gampopa followers of the Tilopa-Naropa-Marpa-Milarepa lineage wrote religious biographies (rnam thar) which were intended to “stimulate the devotion of present and subse- quent generations of practitioners, disciples, and lay followers.”70 They were also intended to reinforce their projection of a particular image concerning the Kargyu and Tise/Mapham and in the wider sense these biographies were a means of embedding the World religion in the local landscape. In the mid- 12th century Gampopa composed a short biography of the Indic holders of his lineage, Tilopa and Naropa, and his account of Marpa and Milarepa was also recorded.71 Of those mentioned above, a rnam thar of Marlungpa dates to 1241,72 and Jigten Gönpo’s rnam thar was another early text. But following Gtsang Smyon Heruka’s biography of Milarepa in 1488, accounts of Götsangpa (1503), Marpa (1505), Tilopa (c. 1523), Naropa (1523), and Pagmo Drupa (1552) were all composed within a comparatively short period.73 Kargyu pioneers became spiritual Heroes within this discourse, which bears many similarities with that surrounding Padmasambhava. They were portrayed as possessors of magical powers that enabled them to overcome the demonic local forces and convert the territorial deities—and thus the landscape in

68 Vitali (1996: 533). 69 Vitali (1996: 507). 70 Tiso (1994: 884–888). 71 Roberts (2007: 64). 72 Vitali (1996: 589); also see Quintman (2008: 375). 73 Smith (2001: 73–78). Klimburg-Salter (1997: 208) observes that regional artistic production also increased in the 15th and 16th centuries, suggesting a wider cultural movement presumably requiring patronage by the Guge kings. buddhacising the mountain 325 which they were manifest—into followers of the Buddha. Some local deities were understood to welcome the Buddhist teachings,74 but others had to be suppressed through ritual magic. In this construction, what had once been prosaic details of the pioneer’s lives took on cosmic significance and the lit- erature became history written by winners—winners with God on their side! Distinction between actual and visionary journeys became unnecessary, while ethical issues were resolved in the understanding that the ends justified the means because such Heroes were enlightened and their opponents karmically- challenged. The setting of these Heroes’ lives went beyond the worldly to that of the divine, for it was this that enabled them to convert the divine to the worldly, as they did in opening the circumambulation of Tise to ordinary pilgrims. Ulti- mately, being regarded as fully-realised beings, their visionary authority in such matters as the interpretation of Tise was more powerful than the precise logic of the Sakya Pandita’s position. For the authors of these texts, there was the claim to be the inheritors of that authority, and explicitly in such cases as Gtsang Heruka, the clear implication that the authors were of the same mettle. For their sect(s), there was the aura of inheritance. The Kargyu may have been temporarily eclipsed by the Sakyapas and eventually sidelined by the Gelukpas, and their fields of operation may have been the peripheries of the Tibetan world, but they claimed the centrality of divine sanction and cosmology. Studies of this literature have demonstrated that the biographical core of these accounts underwent a progressive hagiographical development in which considerable amounts of stock legendary material, and accounts of miraculous signs and portents, visions and prophecies, were attached to the core account in order to embellish the sacred associations and character of the subject’s life. These diverse accounts are often impossible to reconcile with each other or even with their earlier forms. Peter Roberts has defined stages in regard to the evolution of the narratives around Rechungpa and Milarepa that might be applied more widely to Kargyu rnam thars;

1 The earliest phase, from the twelfth to the thirteenth century: The appearance of localised narrative traditions that are independent of each other. 2 Thirteenth to fifteenth centuries: Independent traditions merge in var- ious ways to create new syncretic traditions.

74 See Vitali (1996: n. 670 407–408) where Madros appears to welcome the teachings, but see Vitali (1996; n. 743 445); also see Thakur (2012: 214). 326 chapter 12

3 The end of the fifteenth to the sixteenth centuries: What were to be- come the standard biographical or historical works are created from the earlier works. 4 The seventeenth century to the present day: The modern Tibetan tradi- tion of the last three centuries, which usually summarizes the standard works.75

Robert’s schema might be co-related to both the wider background of political processes as they affected the Kargyu sects and those aspects of their heritage that they shared, as well as to the Tise pilgrimage. In both cases there seems to be a consolidation process. In the first phase individual renouncers travelled to the region as part of their religious practice or carya phase and required only a minimum of patronage such as might be found amidst local peoples. Then in the 13–15th centuries, institutional support was developed as individual under- standings became those of a wider tradition and shared bodies of knowledge of the site developed. Those were represented by the body of literature emerg- ing in the 15–16th centuries, which drew on a range of cultural symbols and determinants to establish an indigenous canon of beliefs around the site. Those constructions were then established within wider Tibetan society. Subsequent embellishments consisted only of commentaries, refinements, reiterations of the established body of knowledge, and the occasional polemical defence of their position. This implies that a relatively fixed consensus of understanding concerning Tise was reached by Tibetans as late as the 16th century. In line with Pauline Belton’s conclusion that in regard to the accumulation of sacred elements around a pilgrimage site, “the more the better”,76 it was in the last phase of development that the complex was garnished with a multi- tude of myths and legends surrounding almost every feature of the landscape. Crucial to the development of the site as a Tibetan sacred place (rather than a local Buddhist tradition), was that the sacralisation process allowed space for the expression of diverse traditions reflecting the wider Tibetan cultural envi- ronment. Thus the Nyingma sect, as part of its claims to national prominence, could hold that a particular cave had been the site of Padmasambhava’s practice in the region and hence associate themselves with the sacred site.77 The result

75 Roberts (2007: 3). 76 “Sacredness may be acquired in many ways, the more the better”; Belton (1984: 24). 77 A simple sectarian chronological progression cannot be proposed, however. Cutler (1996: 72) cites the rnam thar of Zangs gling ma, a gterma text revealed by Myang ral nyi ma Od zer (1124–1192) as including Tise among the places where Padmasambhava meditated. Also see Huber & Rigzin (1999: 133) citing a biography of Padmasambhava, “[a]t white snowy buddhacising the mountain 327 was that what was once an exclusive sectarian claim became open once the region’s essential character had been determined and established. Nor should we forget that this sacralisation had economic benefits both in attracting pil- grims and in enhancing the stature of local religious specialists as local ritual performance and associated artistic expressions responded to the demands of the new constructions. This was in one sense a tribute to the ecumenical aspect of Tibetan Bud- dhism, but it was also an important part of the nationalisation of the site, with the necessity of broadening the appeal beyond the local in order to appeal to the wider understandings of a national audience. Thus just as Milarepa was transformed from an individual holder of Indic Buddhist lineages into a “Founding Father” of the Kargyu and thence to a Pan-Tibetan “Hero”,so too was Tise/Mapham transformed from a local site to a regional Buddhist sacred site and subsequently to a national sacred centre. The full ‘nationalisation’ of Tise may be dated to the visit there of the 1st/4th Panchen Lama (Blo bzang chos kyi rgyal mtshan; 1567–1662). A respected scho- lar and diplomatic peace-maker, he was a teacher and ally of the “Great Fifth” Dalai Lama (1617–1682) who awarded him the title Panchen (“Great Teacher”) that was later applied retrospectively to his three predecessors. Shortly after the death of the 4th Dalai Lama in January 1617 and the birth of his successor in October of that year,78 the Panchen was invited to Tholing by the Guge rulers. After three months he returned to his monastery of Tashilumpo, having made the circumambulation of Tise as he departed.79 No other Tibetan of his stature ever emulated this visit,80 but his ritual acknowledgement of its sanctity was a significant symbol of Tise’s acceptance into the Gelukpa understanding of Tibetan Buddhist sacred geography.

Ti-se he subdued the 28 deities of the lunar mansions; In the elephant secret cave he also concealed a treasure.” As noted, however, the Tise region was a site for Tantric meditation in the pre-phydar period, what is not established is that it was a specifically Buddhist sacred site. The early date of this Nyingma tradition seems to indicate the contested nature of the site in the 12th century, before the definitive moves by the Kargyupa. 78 Petech (2003: vol. ii; 569). 79 Petech (2003: vol. ii; 43); Karmay (2007: 144–145) records the Panchen’s wide-ranging religious views, which included knowledge of Dzogchen teachings as well as those of the Kargyu, Nyingma, etc., which were also popular in Ngari. 80 Another notable visitor was the 16th Karmapa Rigpe Dorji (1924–1981) who visited the site in 1947, making three circuits of the mountain and one of the lake; Bellezza (2005: n. 99 110). 328 chapter 12

During the 16th century, the ties between Ngari and central Tibet had be- come increasingly close and the Gelukpa presence in the west increasingly prominent.81 The Panchen’s visit to Guge was “the last moment of glory and prestige for the decaying kingdom.”82 The economic situation deteriorated and increased desertification of the region was another factor in its decline.83 Relations with Ladakh deteriorated into fighting by 1622 according to Tibetan sources and a short-lived truce collapsed partly, “because of the marauding activities of the [Drukpa] monks of the Manasarowar-Kailāsa hermitages.” (While this presumably reflects decreasing patronage of the renunciates in a declining economy, some links to the increasing militarisation of renunciation in India around that time are also possible.) Bhutan then became involved in these disturbances, sending troops into Guge in support of their Drukpa co-religionists and in 1630 Guge was invaded and conquered by Ladakh. The Ladakhi rulers were Drukpa and “coldly indiffer- ent (to say the least)” to the Gelukpa.84 They embraced the regional construc- tion of Tise by undertaking pilgrimages there, but any benefit to the Kargyu in Ngari was short-lived. Ladakhi hostility to the Gelukpa and the 5th Dalai Lama’s consolidation of the Tibetan world into an at-least-partly centralised state were factors in the despatch of a Mongol-Tibetan army to the west in 1679.85 This force defeated Ladakh and took the Guge-Purang region under central Tibetan rule, which lasted until the Chinese takeover in the 1950s.86 While there is evidence of subsequent competition for power and patron- age between the Tibetan Buddhist sects in Ngari, the Kargyu construction of Tise/Mapham as sacred Buddhist space was apparently accepted by the various sects with the notable exception of the Sakyapas.87 The Gelukpas’ acceptance of the sacrality of the site may be seen within the wider context of their polit- ical ascendency, with the need to reconcile the new population dictating that acceptance.88 This was also consistent with their tolerance of local aspects of

81 Petech (2003: 42–43); Petech (1997: 243–244). 82 Petech (1997: 247). 83 On regional climate change, see Bellezza (2002: 12–14). 84 See Petech (1977: 59, 139). 85 Petech (1947: 24–25) notes that the leader of the Tibetan forces carried out rituals at the sacred site en route to fight the Ladakhis. 86 Petech (1997: 247–248); also see Petech (1977: 42–44). 87 There was however, a Sakya monastery at Mapham and presumably some accommodation occurred at a local level. See note 94. 88 The District administrators made an annual circumambulation of Tise/Kailas; undated report (circa 1948) by British Political Officer, Captain R.K.M. Saker, “On the Future of buddhacising the mountain 329

Buddhism that did not challenge either their political position or the funda- mental criteria for acceptance of Buddhist orthodoxy. Two of the eight monas- teries that were later established around Lake Mapham came under Gelukpa rule,89 while their ultimate authority was represented by the largest regional monastery, Tholing, a branch monastery of Drepung (the largest monastery in Lhasa).

Vision and Origin

We have seen that Indian renunciates such as Swami Tapovan constructed sacred geography through informed visionary authority, with their revelations later amplified by mythology and recorded as text. Kargyu renouncers sim- ilarly read the landscape they traversed through the visionary perspectives of their faith. When they visited sites such as Jvalamukhi they knew that these places were also held sacred by the Śaivites. But in their understand- ing they held a higher truth, that the Śaivite deity had been subjugated by Heruka. Thus they interpreted the manifest landscape of this Indic site as Bud- dhist. Given the Tibetan interest in India as the birthplace of the Buddha, and the identification of sites in the western Himalayas and elsewhere as those of Budddhist mandalas, numerous Kargyu renunciates travelled to those places. While the proposed alchemical economy would have freed them from the need for local patronage, they were not operating outside of sectarian aims and identities. Toni Huber has located the significance of these renunciate journeys to India in the context of (re-)locating the 24 tantric pīṭhas of the Cakrasamvara tradition in the western Himalayas.90

the British Trade Agency”; (copy in the possession of the author, courtesy of Mrs A. Saker). 89 There are eight monasteries around the lake (a geomantic number in the context of the mandala). It is notable that none are or appear to have ever been, Bön (the last-named reflects the toponym of a mountain, not its ruling sect). The eight as given by Ricard (1994: n. 65 344–345) are, with their final sectarian affiliation in brackets: ser ra lung (Drigung), mnyes ’go (Sakya), khrus sgo (originally Druk, later Geluk), ’go tshugs (Druk), byi’u (Druk), gser gyi bya skyib (Drigung), glang sna (Druk), and bon ri (Geluk). Also see, Buffetrille (2000: 85–93). For traditional accounts of the geomantic-historical origins of these monasteries from the 1896 guide-book, see Huber & Rigzin (1999: 146–151). For a list of regional monasteries and nunneries, see Pranavananda (1984: 232). 90 Huber (2008: 101–108). 330 chapter 12

Among the most significant of these journeys was that of the famous Göt- sangpa Gonpo Dorje,91 the Drukpa Kargyu renunciate renowned as the “open- er” of the Kailas circumambulation. Götsangapa was at Tise in 1214–1217 then travelled from the alchemical centre of Tīrthapuri via the sacred Lahauli moun- tain of Gandhala,92 to Chamba via the Kugti pass and on to Jvalamukhi.93 Götsangpa’s disciple Orgyenpa Rinchen Pal (1229/30–1309ce), followed his teacher’s path to the west. Having studied under Götsangpa he moved to Kailas-Manasarovar at his master’s suggestion in 1253/4,94 and after briefly serving as rajaguru to the king of Rudok he travelled to Triloknath and Kulu.95 At Jvalamukhi, in co-operation with Śaivite specialists, Orgyenpa translated alchemical works such as those dealing with the preparation of mercury.96 Both he and Götsangpa were masters of esoteric religious technologies that blurred sectarian boundaries, such as the Cakrasamvara cycle, the Six Yogas of Naropa,97 and those Orgyenpa famously received from a ḍākinī in Uddiyana. Followers of Tibetan renunciate traditions continued to visit Chamba through the medieval period, although with the pīṭha tradition established new motivations developed for their journeys. Peter Schwieger has located the journeys of Taksang Repa Ngawang Gyatso (sTag tshang ras pa Ngag dbang rgya mtsho: 1574–1651), a key figure in Ladakhi Buddhist history, in the desire to return to the routes of the now-legendary Kargyu pioneers, whose biogra- phies were then circulating. Thus he followed the earlier route and concerns of Orgyenpa.98 Born into the Sakya-associated Khon family, Taksang’s career followed the Kargyu model of broad-based esoteric learning. Originally in receipt of Nying- ma initiations, he studied practices such as gterma and the Six Yogas of Naropa

91 Stutchbury (1999: n. 11 176) notes that there are fourteen biographies (rnam thar) of Götsangpa! 92 Stutchbury (1999: 102–103; for sites associated with him in Karza, see 155–156). 93 Huber (2008: 102–102); Thakur (2001: 47, 48). 94 Li (2011: 152) notes that Orgyenpa’s biographies give no details of his visit to Kailas but relying on the ‘timeless’ image of the site none-the-less presumes “he must have done rounds of circumambulation”, an assumption that seems unwarranted. 95 Vitali (1996: 421, n. 703 472, 566); also see Huber (2008: 103–104); Crystal Mirror 6 (Journal of the Tibetan Nyingma Meditation Centre. Berkeley: Dharma publishing) 1984; 440. 96 Gene Smith cites a text on the preparation of mercury drawn from the Rasasiddhiśāstra attributed to Vyāḍipāda which was translated into Tibetan by Orgyenpa and Śrī Naren- drabhadra and a similar work he translated with one Śivadāśa; Smith (2001: 189–190). 97 Stutchbury (1999: 156). 98 Schwieger (1996). buddhacising the mountain 331 and was later initiated into the cult of Cakrasamvara. Taksang Repa also under- took cemetery practice and after becoming a follower of the leading Drukpa Kargyu teacher of the time he first reached Kailas in 1598. Journeys—real or visionary—to Tsari and the Chinese sacred mountain sites of Omei shan and Wu-t’ai shan followed before, in 1613, he was sent to the western Himalayan pitha sites via the Sutlej route from the Kailas region to Kinnaur, Jvalamukhi, and then Chamba. After a year in Lahaul, he visited Uddiyana, Kashmir, and Zanskar before reaching Ladakh late in 1616.99 While such journeys could have been unique, these accounts probably represent only a fraction—the most influential fraction—of the renunciates that made such journeys.100 Tise was thus part of much wider sacralised landscape conceptions.

A Bhutanese Kailas

The Bhutanese association with Kailas developed in the 17th century as a result of close religious ties between the ruling stratas of Bhutan and Ladakh. The Ladakhi kings were followers of the Drukpa Kargyu which in the late 16th century fractured in an incarnation dispute. Zhabrung Nawang Namgyal (1594–1651?), one candidate for the new incarnation of the sect’s hierarch, fled Tibet to what is now Bhutan. That polity then comprised a series of small principalities whose peoples were spiritually loyal to the Druk and Nyingma sects. Nawang Namgyal proved a stabilising figure who settled and united the land under his spiritual and political leadership. An exceptional personality who is considered the founder of modern Bhutan, he gained a series of victories over invading Tibetan forces and was recognised as the head of the southern branch of the Druk (the lho ’brug).101 The Ladakhis under their great ruler Sengge Namgyal (1570–1642), had diplo- matically maintained ties with both factions of the Druk, many of whose renun- ciates were practicing in Guge-Purang. As we have seen, with the Kailas region in a disturbed state by the 1620s there were major clashes between the renun- ciates and forces loyal to the Guge kings, whose primary allegiance was to the Geluk sect. The Druk community gained the support of the Zhabdrung, who sent raiders from Bhutan to attack royal interests in Guge.102 In 1630 the weak-

99 Schwieger (1996); also see Huber (2008: 104–108, 176–178). 100 On which see, Huber (2008: 100–104, 175–179); Templeman (1997); Schwieger (1996); Shastri (2002): and Shastri (2009; 10–12). 101 On the history of Bhutan, see Phuntsho (2013). 102 On which see Petech (1977: esp., 42–44). 332 chapter 12 ened kingdom was conquered by Sengge Namgyal’s forces from the west, and came under Ladakhi rule for half a century. Ladakh and Bhutan remained allies, not least against Tibet, and the brother of Sengge Namgyal, one Prince Standzin (Ladakhi: Bstan ’dzin), travelled to Bhutan where he became the Governor of Wangdi district. Sengge Namgyal then offered the Zhabdrung a number of monasteries in the Tise region, “as a mark of respect.”103 The most significant of these was Darchen, the starting point for the Kailas circumambulation, and following the gift of these monas- teries an official Bhutanese representative was posted there in 1678.104 After their defeat by the Tibetans in the war of 1681–1683, Ladakh’s rule over Guge-Purang was ended in 1684 by the treaty of Temisgang (Gtingmo sgang). Tibet accepted Bhutan’s religious position however, and Bhutan retained con- trol over its monastic holdings in Ngari until the Chinese take-over in 1959. But although Kagyu Wangchuk (d. 1864), later Bhutan’s 48th Desi (‘secular regent’) served there,105 Bhutanese today remember that posting to those remote insti- tutions was almost a punishment and one that was delegated to others if at all possible.106 A low-ranking Bhutanese official was certainly stationed there in the 1920s when friction arose between Tibetans and Bhutanese culminating in a murder at Darchen. A British Political Officer, Frederick Williamson, acted as an intermediary in this dispute.107

Mountains as Frontiers

There are indications that in early Tibetan cultural understandings mountains were, in the absence of demarcated borders, territorial determinants mark-

103 Bray (2003: 192) lists the following regional monasteries as gifted to the Bhutanese; Gnyen- po’i-ri-rdzong; Bri-ra-phug; Rdzu-’phrul-phug; Ge-rdzong; Bya-skyibs; Ye-ri-dgon-phug; Gad-ser; So-mo-rgyu; Shi-ha-ra; and Dar-chen Bla-brang-dgon; the latter the most signifi- cant of these. Also see Mehra (1968). 104 Aris (1979: 249). Pranavananda (1984: n. 1 125) notes that a statue of Nawang Namgyal in Darchen was taken by Sikh pilgrims to be of Guru Nanak because of Nawang Namgyal’s beard! 105 Phuntso (2013: 429). 106 Discussion with Dr. Karma Ura, Centre for Bhutan Studies, Thimpu, September 2003; also see Wakefield (1966: 58–59). Satchidananda (1984: 108, photo on 111) refers to an unidentifiable “king and queen” as representatives of the Bhutanese government, who spent the summers there. 107 See Oriental and India Office Library, l/p&s/12/4163–1165, report of F.W. Williamson, 14 December 1932. buddhacising the mountain 333 ing the boundaries of political formations. I have previously suggested, in dis- cussing their role in the Zhang-zhung polity,108 that particular mountains rep- resented central identities of sacred power and acted as “protectors” of local cultural spheres, not least their frontiers. The powers of the deified mountain (yullha) could be drawn on through ritual by human intermediaries seeking protection and, significantly, could be transferred to a new location if the group migrated.109 This raises the possibility that local identities and conceptions sur- rounding a mountain such as Tise may, in a process associated with migration, replicate or be a development of those earlier attached to other mountains. Among the ritual equipment of the ‘spirit-mediums’ that intercede with the territorial mountain deities are rites for the suppression of enemies (con- ceived as demons) on the borders of the mountain gods’ territorial range.110 An understanding of these mountains as territorial markers and frontier protectors whose powers could be harnessed by the appropriate ritual thus seems feasible, particularly given that a mountain can “move” with its peoples. In that these sacred mountains could represent their surrounding clan territory there were clear political implications to their control, as discussed by Samten Karmay in the context of a cult of nine yullha. He observes that these mountains, the identity of which varied with the political circumstances of the Yarlung kings’ expansion were “mostly located in the territories that were conquered and that constituted the first nucleus of the Tibetan state.”111 The control of mountains may thus be contested over time by political as well as religious tendencies for mountains exist as symbolic and actual locations for competing ideologies— religious, political, social, and so on. Following these understandings, two sets of mountain frontiers might be identified. The first is the Kailas mountains of the modern Indian Himalayas that represent the westerly and southern frontier markers of a late Zhang- zhung phase. Each mountain had associations with that cultural frontier. Dar- dic culture also includes the sacred mountain model,112 indicating that the con-

108 See McKay (2011). In addition we may note parallels from Mongolia; see Birtalan (1998); also see Humphrey (1995: n. 10 159–160) which records that on the “Mongolian-Siberian border from Lake Xovsgol to the Xangai Range, a distance of some 1,000km., there are six mountains called Burin Khan”, suggesting parallels to the Kailas mountains of India; also see Vreeland (1957: 199). Historically, a mountain might serve as both a centre and as a frontier marker. 109 See Buffetrille (1996). 110 See Bellezza (2005: 320–321); also see Bjerken (2004: 36); Martin (1994: 41, n. 153 42). 111 Karmay (1996: esp., 62, 65). 112 Tucci (1977: 26–28); see however, Dollfus (1996). 334 chapter 12 cept would have been understood on the Dardic, or Indo-Iranian frontier with Zhang-zhung in which Kaplas and Manimahesh Kailas’s are situated, while Kinnaur Kailas was probably within Zhang-zhung. Sri Kailas is also located on the Indo-Tibetan/Zhang-zhung frontier, while even the Adhi Kailas is situated at the head of the Darma valley, where the language and culture indicate his- torical links with a Tibetan/Zhang-zhung Bön culture.113 The second set that might be proposed as Tibetan frontier mountains is that of the principal gnas ri mountains of present-day Tibet (as identified by Toni Huber114). These are peaks that have been subject to the full process of Buddhacisation to the point where their dominant identity is articulated by a World-religion (Buddhist or Bön). These mountains are Tise (Kailas), Targo, Nyanchen Tanglha, Amnye Machen, Kawa Karpo, Kongpo Bönri, Tsari and Labchi. While Nyanchen Tanglha marks an internal frontier between the Zhang-zhung and Yarlung principalities, the others surround the Tibetan cul- tural region and might be conceived of as archaic frontiers of the extent of “Zhang-zhung” or later Tibetan cultural influence or migration over time. That these mountains have been Buddhacised (or “Bön-icised” in the case of Kong- po), may thus be seen as an aspect of the Tibetanisation of their region’s identity after the absorption of its local kingdoms into a wider Tibetan state, perhaps connected to Chinese geomantic concepts and territorial expansion processes.

Translating Sacred Mountains into Western Understanding

Western constructions of Tise/Kailas as a sacred centre are overwhelmingly reliant on Indic understandings and references to its sacrality in the Tibetan world are generally brief and superficial. There are a number of reasons for this. Firstly the Western understanding of Tibetan language, religion, and cul- ture lagged far behind that of their understanding of the Indic world. Serious efforts to understand the Tibetan world only gained wide currency in the late 1800s, more than a century after the discovery of links between Sanskrit and the European languages had stimulated a major effort both in British India and in Europe—not least Germany—to explore those links. Only a handful of West- ern scholars learned the Tibetan language before the post-1959 exile period and their pioneering focus was understandably not on sacred mountains. In addi-

113 Personal correspondence, Dr. Yael Bentor, 8 May 2006; Darma is a Tibeto-Burman lan- guage, on which see Krishan (2001). 114 Huber (1999: 5). buddhacising the mountain 335 tion, Tibetan informants were few in comparison to the easy access European scholars enjoyed to the Brahmanical class in India, and those who did assist Western scholars were central Tibetans such as Sir Charles Bell’s informant Achuk Tsering or exiles such as L.A. Waddell’s Darjeeling monks. The existence of Tibetan guidebooks to Tise only became known in the West as a result of Tucci’s research in the 1930s and their translation and analysis took place much later.115 Thus while Kailas was being constructed as a sacred centre in Western understanding there was very little information available in European languages about the Tibetan beliefs concerning the site. The key agents in the Western construction of Kailas did not read or even have access to Tibetan texts referencing the region, and they did not have any knowledge of the historical processes and religious understandings represented by the Cakrasamvara mandala.

115 It is not my intention to discuss the Tibetan guidebooks to Tise as a genre. See in this context, Buffetrille (1997) & (2000); Huber (1999); Decleer (2000); Macdonald (1985). Six Tibetan guidebooks are known, as follows:

1. Gangs ri mtshog gsum gyi dkar chags, a circa 13th century guide by a Bönpo siddha, Ye shes rgyal mtshan (a version of which was printed at Dolanji in 1973). On which see, Ramble, (1999: esp., 17–18). Ramble states that the text is in the form of answers to the Bönpo’s questions about the site by 12 (“all more-or-less well known”) Bönpo siddhas. A translation of a section of the text pertaining to Ye shes rgyal mtshan’s spiritual experiences at the mountain is found in Loseries-Leick (1998: 159–160). 2. gNas chen Ti se dang mtsho ma pham bcas kyi gnas yig bskal ldan thar lam ’dren pa’i lcags kyu, an 18th century guide by Don rgyud bstan ’dzin, referred to by Ehrhard (2001: 279–285); not otherwise known. 3. ’dzam gling gangs Ti se dkar chag tshangs dbyangs yid ’phrog ’dod, written in two versions, 1844 and 1847 by the Bönpo Dkar ru Grub dbang Bstan ‘dzin rin chen (1801-?). On which see Norbu & Prats (1989); also see Cutler (1996: 125–197). 4. Gangs ri chen po Tise dang mtsho chen ma dros pa bcas kyi sngon byung gi lo rgyus mdor bsdus su brjod pa’i rab byed shel dkar me long [short title—Gangs ri’i gnas bshad shel dkar me long] a Buddhist guide by dKon mchog bsTan ’dzin Chos kyi Blo gros ’Phrin las rNam gyal (1869–1906), the 34th hierarch (gdan rabs) of the Drigung Kargyu, who visited Kailas in 1894–1896. On which see, Huber & Rigzin (1999: 125–153); also see, Filibeck (1988). 5. Gangs dkar Ti se’i gnas kyi dus chen rta lo ’khor chen la rten ’brel, written by the Buddhist Chos dbying rdo rje (1949–1999) and published in 1990 in Bod ljongs nang bstan. On which see Buffetrille (2000: 17–99). 6. gNas chen gangs ri mtsho gsum chu bo bzhi dang bcas pa gtan la dbab pa lung don snang bar byed pa’i me long by Ngag dbang ’phrin las. Buffetrille (1998: 25) identifies this as found in a collection published in Dharamsala entitled “Tibetan Guides to Places of Pilgrimage.” 336 chapter 12

In addition, the 19th century Western construction of Buddhism operated within the paradigm of the living faith representing a catastrophic decline into popular superstition and primitive, “priest bound” religious ritual from a “pure philosophical doctrine” of the historical Buddha. Tibetans did not believe that the Buddha lived on Tise, something which would have been accessible to Western minds. Rather their most prominent belief was that the mountain was the residence of the fearsome Tantric deity Demchok Heruka, along with his heavenly consort and a range of dakinis and cemetery-dwelling Bhairava demons. Whether in the minds of late 19th and early 20th century Protestant scholars or pious proto-New Agers, this was a belief that was simply too alien, too “Other”, and too incompatible with even the most enlightened of understandings of Buddhism then current. Indeed it seemed to represent a prime example of the presumed decline of Buddhism into popular superstition and primitive magic. Thus even today popular Western articulations of the Tibetan understanding of the site struggle to embrace or explain Buddhism’s relationship to such esoteric figures. (Indeed, one may detect a tendency for the site to be promoted as one for Buddhist practice in general, rather than specific deity-centred worship, a situation that seems to mirror that of the pre- or early phyidar era.) Furthermore, the Western understanding of Tise/Kailas emerged during the “Great Game”, the period in which the region was part of contested territory between British Indian, Russian, and Chinese empires. Tibetan perspectives were of little concern to the political struggle and the possibility existed in the minds of British Indian strategic planners that the Kailas region—indeed the south if not all of Tibet—could be taken into some kind of protected or annexed status by the British empire.116 With those political imperatives paramount,

116 See, for example, National Archives of India, Foreign Department, 1905, Secret e, February, 1398–1445, report of Louis Dane, 21 May 1904 (i.e.: during the Younghusband mission). In this report Dane, the Indian Foreign Secretary, states that he was concerned by Russian influence in Khotan, and suggests strengthening western Tibet by extending the British India frontier up to Kun Lun range—i.e.; annexing Ngari! The reply dated 13 June 1904 by Secretary of State Ampthil states that the British government would “have a fit” if he proposed anything of the kind. In response Dane immediately proposed a trade mart be established at Gartok, which proposal was accepted. (A file note dated 4 November 1904 states this was approved by Colonel Younghusband.) Dane may have been using a common bureaucratic strategy of making an extreme proposal in order to obtain by compromise a less extreme aim, but there were other initiatives on these lines at that time. On this and the history of the British Indian trade agency at Gartok, see McKay (1997: 158–165, also see, 29–42). buddhacising the mountain 337 it was only the economic possibilities of the Kailas pilgrimage that were a concern to the British, along with the need to properly map the region and gain a knowledge of possible invasion routes and exploitable natural resources there. The only aspect of sacred geography that was important in this context was as a justifying factor in Indic expansion; that is, if Kailas was a sacred Indic centre, then its absorption into India carried a certain cultural logic that would assist the potential process of taking it over. That it was also sacred to the Tibetans was not knowledge that in this context it was beneficial to bring out. That the mountain was the residence of an (in their understanding) obscure Tantric deity was of no importance compared to the Indic construction with its claim to ancient roots articulated in Sanskrit texts. Thus in the construction of the modern understanding of Kailas, Tibetan voices were either not heard or were simply downplayed. There is no evidence that at the dawn of the 20th century, Tise enjoyed a status notably higher than any other Buddhacised mountain in the Tibetan world. It did attract pilgrims from throughout the Tibetan Buddhist lands (which some other Buddhacised mountains such as Amnye Machen and Kawa Karpo did not117), but it could not compete with the sanctity of Lhasa in terms of attracting pilgrims. Nor could its patterns of ritual worship compare with, for example, the great state-sponsored circumambulation of the sacred mountain of Tsari. There the apparatus of the Tibetan state negotiated passage for 15–20,000 pilgrims through hostile tribal territory, and the Dalai Lama’s robes led the procession in a symbolic articulation of his presence at the head of the ritual.118 Yet the 19th century did see the production of both Buddhist and Bön guide- books to Tise, suggesting a certain impetus to the worship at the site in that era. Whether this represented a dynamic aspect to the period, a response to moder- nity, or an attempt to preserve what was fading is difficult to ascertain. This was a dynamic period in Tibetan Buddhism in other regions, with the ecumenical Rimed (ris med) tendency prominent, but sectarian elements seem to have pre- dominated in Ngari and it is difficult to isolate indigenous agencies in the early 20th century processes by which Tise Kailas rose to wider prominence in the wider world.

117 I am indebted to Katia Buffetrille for this distinction, demonstrating the gradients involved in the processes by which yullha (or other?) mountains were transformed into Buddhist (or Bön) mountains. 118 On this see Huber (1999: esp., 128–152); also see Huber (1997: 225). 338 chapter 12

There is a relationship between weak or absent state structures and the strength of a local mountain deity cult,119 and the remoteness of the region from events in Lhasa, Delhi, and other regional centres may have stimulated the development of the pilgrimage and its role in Tibetan society as it did in the Western imagination. Western and Indian reports on the region in the late 19th and early 20th centuries acknowledge the pilgrimage, but they also consistently indicate that the area had become a neglected periphery of the Tibetan state with corrupt officials, decayed institutions, spiritually stagnant monasteries, and an ever-present threat from bandits. To an extent this seems a realistic portrait. The widely acknowledged presence of bandits is characteristic of regions where central control is weak or absent and economic processes are either in abeyance or insufficient for social development. The system which existed whereby central Tibetan aristocrats were ap- pointed to the governorships of the region for short periods in which they had to recoup the money they had paid to obtain the position and then obtain as much profit as they could on top of that, encouraged poor government and cor- rupt activities. Only by taxes on trade and other enterprises could the governors obtain profit, for any excess income from the pilgrimage was remitted to the Bhutanese authorities controlling the religious establishments around Kailas and it was therefore apparently of no great financial benefit to the Lhasa gov- ernment. In addition, there is little evidence in the way of significant monastic or renunciate individuals, new movements or influential commentaries to sug- gest a dynamic Buddhist monastic body existed in Ngari by this time. What was dynamic was the emergence of a (new?) Bön claim to the sacred complex, as will be seen in the next chapter.

119 Karmay (1996: 70); also see Gingrich (1996: 237–239). chapter 13 Zhang-Zhung, Bön, and the Mountain

Introduction

Just as modern Buddhism and Hinduism claim Tise/Kailas as a sacred centre so too do the Bönpo, who identify themselves as followers of a religion that predates Buddhism in Tibet. Modern Bönpo believe that their religion pre- dominated in the Zhang-zhung kingdom before it was conquered by Yarlung dynasty Tibet, and they identify Tise as the ‘soul-mountain’ of Zhang-zhung. As we have seen, Zhang-zhung was a historical state whose territory included Tise and Mapham.1 But in discussing a Bön Tise history we are confronted by the fact that Bön is not a “rationalised” system in Weberian terms, one that is “where people seek a consistency and encompassing logic to their world view.” Bön is a system that is “mythologised”,2 with conceptions of history and world views that are meaningful only in contexts fundamentally incompatible with those of scientific modernity or critical historical approaches.3 While there is a Bön canon, the faith has no central authority, fixed religious centre, or single agreed pantheon, and Bön texts are not necessarily systemised, internally cohe- sive, or intended to be compatible with other accounts. Thus in analysing Bön sources, as with the Puranas, no single text can be considered foundational, entirely representative, or pre-eminent in authority. And while we may detect a lineal tendency to move from the singular to the embellished, we cannot con- clude that any one reference reveals a truth otherwise concealed. What we are dealing with is a compilation of bodies of knowledge taken from individuals, families, clans, lineages, regions, and schools of thought. Thus there is no sin-

1 Mapang for Mapham is common, particularly in Bön texts. While apparently synonomous, see Haarh (1969: 371) who cites a Dunhuang text where Mapang may mean “mother[’s] blood … flowing at a child’s birth.” 2 Holmberg (2011: 186–187); the quotation is in regard to the Tamang system. 3 ‘Bön modernism’ is currently a contradiction in terms, with authority remaining vested in elements outside of time and space; see for example, Norbu (2009). This is an entirely ahistor- ical work that cites non-Tibetan sources only in the translator’s introduction. Such ‘Nativist’ works are usually the product of political circumstances, here the construction by a refugee community of an imagined Golden Age in compensation for the psychological assault of dis- placement and exile. And in this case to an extent it is also a response to demands—Western and otherwise—for exotic images of ‘Lost Kingdoms’, ‘Concealed Wisdom’, etc.

© koninklijke brill nv, leiden, 2015 | doi: 10.1163/9789004306189_016 340 chapter 13 gle Bön Tise to define in such timeless constructions as the ‘soul mountain of Zhang-zhung’. Indeed there is no single Bön religion. The meaning of the term shifts over time and space, and escapes easy definition. The modern World-religion of Bön claims continuities with the spiritual world of the Tibetan plateau in millenniums long passed, but the veracity of the claim is highly debatable. Indeed, there is an intense and on-going debate over Bön origins and the processes shaping its modern identity. While we are not concerned here to reconcile those issues, an awareness of this context is essential in shaping our conclusions, which are necessarily tentative. But as Sherry Ortner put it; “[i]t seems to me that a tentative reconstruction, which might be modified by later evidence, is more useful than none at all.”4 Thus, after briefly outlining the relevant historiography of Bön, we consider the evidence for a sacred Tise centre in its claimed heartland of Zhang-zhung and discuss the evolution of modern understandings of a Bön Tise as their historical centre.

Situating Bön

Popular understandings of Bön and of a Bön Tise are still shaped by the ‘ani- mist, shamanist, pre-Buddhist religion of Tibet’ model which defined Western understandings of the faith until the late 20th century. But in recent decades there have been considerable advances both in the study of Bön and in its self- projection. Bön in its modern form is actually a distinct and sophisticated belief system that shares many conceptual and structural elements with Buddhism.5 These elements are shared to the extent that in certain contexts Bön is accepted as a form of Buddhism, albeit looking to a different historical founder, Shenrab (gShen rab), who Bönpo consider long predates the historical Buddha. They understand Shenrab as having brought Bön teachings to Zhang-zhung from the land of ‘Tazig’ (sTag gzig), his homeland to the west (and thus do not regard Bön as indigenous to Tibet). According to Bön self-projections Shenrab’s teach-

4 Ortner (1992: 25). 5 Essential to an overall understanding of the process by which we have come to understand Bön are: Kværne (2000); Martin (1997: 263–305); and Bjerken (1998: 92–107) (Bjerken’s clas- sification of Samten Karmay as a “nativist” cannot however, be supported given that their defining characteristic is a refusal to engage in critical enquiry or comparative textual analy- sis, both of which are features of Karmay’s work.) On Bön, see in particular however, Blezer (esp., 2011; and forthcoming 3 vols.). bön and the mountain 341 ings flourished in pre-Buddhist Zhang-zhung, only to be persecuted and greatly weakened after the Tibetan take-over of that kingdom. The problem with this Bön ‘history’ is that the critical study of its textual sources has demonstrated that Bön in its modern form cannot be identified before the mid-10th century emergence of texts self-consciously identifying themselves as Bön. As Henk Blezer states;

The history of Bon master narratives about their founder, geographical and cultural location of origin, and the history of ideas of its major doc- trines (the “three pillars of Bon”), in the form in which they are accessible now, all roughly converge in the 10th–11th [century] a.d. or later.6

From around that time Bön was apparently systemised as a distinct entity in response to the prevailing religio-political developments of the phyidar (as discussed in preceding chapters). This process must have been an on-going one, but emerges into the historical record most clearly with the personality of Gshenchen Luga (gShen chen Klu dga’: 996–1035),7 who can be acknowledged as the systemiser, indeed even founder, of modern Bön. His disciples headed some of the principal Bön teaching lineages and while there were earlier such discoveries, it was his finding of a number of texts in 1017 that signalled the ‘later propogation’ of Bön.8 The Bönpo, like the Nyingma, rely to a large extent on such texts from the Terma (gter ma: ‘Treasure’) tradition. These are relics including texts under- stood to have been hidden in an earlier period of suppression in order to be revealed at the appropriate time, something commonly believed to occur in accord with prophecy, revelation and vision. Always controversial, such texts became a distinctly Tibetan phenomena. Initially some were discovered acci- dently,9 but increasingly they were found by or ‘revealed to’, charismatic ‘Trea- sure Finders’ (gter ton). This corpus of texts included ‘mind-treasures’, received by Treasure Finders as revelation and first written down by them. Reconciling this corpus historically is difficult. Much of the Bön material is late, intertex- tual with Buddhist works, or comprising cycles of myths that have clearly been reconstituted in different contexts.

6 Blezer (2011a: 156). 7 On whom see Martin (2001). 8 Karmay (1975: 183–187). Also see, for biographical references to significant Bön teachers etc., Achard (2004: 245–276). 9 Kværne (1975: 33–35). 342 chapter 13

This later textual material is also collectively difficult to reconcile with the earlier Dunhuang material, problematising the Bön claim to continuities with the earlier religious world of the Tibetan plateau. In seeking to ascertain wheth- er the term Bön is appropriate to describe that world in both the pre- and post-10th century periods, enquiries thus centre on the identification of earlier elements (including rituals, genealogies, and cycles of myths), and the extent to which these elements emerge—or do not emerge—in some form in later systemised Bön.10 The question also arises as to whether elements of other faiths were drawn upon to form systemised Bön. Among the many proposed influences are other regional belief systems such as Mithraism,11 Jainism,12 and Hinduism,13 all of which have had their proponents. But while the latter shares the strong rit- ual basis of Bön and has had some impact on the Bön pantheon, more weight has been given to David Snellgrove’s theory that Bön represents a poorly- understood form of Buddhism transmitted from Central Asia to Zhang-zhung prior to the accepted introduction of Buddhism to Tibet (in the snga dar).14 This would allow that Bön’s apparent reformulation on Buddhist lines did not represent a radical ideological shift. There is, however, another Bön discourse that has emerged in post-exile Tibetan society, a ‘Nativist’ tendency that regards Buddhist elements of Bön as something of a corruption of the original faith preached by Shenrab. It identi- fies that ‘original’ Bön as centred around three elements; Drung, (‘narratives’) Deu, (‘symbolic languages’) and Bön (‘rituals’).15 In the wider context this iden- tification is located in the promotion of Bön as a New Age religion; exotic, time-

10 On the identification of elements of continuity, implying an indigenous or even ‘proto- Bön’, see for example, Karmay (1983); (1996a). 11 Kuznetsov (1975: 113–114). 12 The Sakya intellectual Tagshang Lotsapa (Stag tshang Lo tsā ba, b. 1405), refers to the ’Phywa Gshen ‘vehicle of Bon’ as deriving from Jainism, although as Martin (2001: 145) points out, “this is the only time anyone has suggested it is related to Jainism.” 13 “Shangshung [sic], embracing Kailāśa, sacred mount of the Hindus, may once have had a religion largely borrowed from Hinduism.”; Stein (1972: 236). Riddles seem an early shared element that might repay enquiry, though there are also Biblical precedents for these. 14 Snellgrove (1987: 326, 390–391, 400–405, 473). I have previously observed (McKay: 2003: vol. 1; 36) that the theory was advanced without supporting evidence. While the Kushan empire must have had some contact with the western regions of Tibet, essential features of Buddhist transmission are absent from Zhang-zhung, most notably monasticism and the associated vinaya, both inherent to Buddhism and its transmission from the earliest period. 15 On which, see Norbu (1995). bön and the mountain 343 less, and in harmony with the Earth. But in distinguishing itself from Buddhism, and advancing the idea of a ‘Golden Age’ of Bön under a mighty Zhang-zhung empire,16 the Nativist tendency relies almost entirely on the uncritical expo- sition of very much later Bön texts, and it is unwilling to engage with critical scholarship, particularly from the Western tradition. (Here it is in stark con- trast to both Buddhist and Hindu modernism, which actively investigate areas in which their beliefs may be reconcilable with science.) In addition to textual approaches to the study of these issues, we have some- what different perspectives offered by archaeological enquiries into Zhang- zhung,17 notably John Bellezza’s extensive “surface archaeology” and recording of human traces in western and northwestern Tibet.18 This work has demon- strated that these now barren and under-populated areas supported much larger populations at some period(s) in the past. But Bellezza’s association of these remains with a Zhang-zhung empire, or unified culture, is controversial, for much of the material remains more probably relate to the Guge-Purang period rather than to Zhang-zhung. Controversy over Bellezza’s work has, how- ever, partly overshadowed the fact that his studies of regional rock art add to the growing body of evidence suggesting the region was in communication with the early cultures of northern Central Asia and Siberia. In regard to Tise and Lake Mapham, we seek, therefore, to determine wheth- er the complex sacred geographies advanced by post-10th century Bön are man- ifestations of earlier sacralities or part of a territorial claim that must be located in the context of phyidar and post-phyidar era landscape contestations.19

Imagining Zhang-zhung Sacrality

Despite the extensive genealogies, pantheons, and sacred histories claimed by much later Bön texts there is actually no evidence of the sacrality of Tise or even Mapham that is either contemporary with the Zhang-zhung empire or recorded for a considerable period after its conquest by Yarlung dynasty

16 A construction that implicitly requires a distinct Zhang-zhang language, enquiries into which are on-going; see for example, Nagano and LaPolla (2001), and esp., van Driem (2001). 17 On archaeology, see Aldenderfer and Yinong (2004). 18 See the various works of John Bellezza listed in the bibliography. 19 This strategy by Nativist religious tendencies is hardly unique to Tibet: c.f. the architects of state Shinto in Japan and their, “elaborately imagined pre-Buddhist cultural and religious landscape that never was”; Maguire (2011: 541). 344 chapter 13

Tibet. This may reflect the wider shortage of sources for the pre-phyidar period, none-the-less, it is a point that must be emphasised. All of the evidence we have for the sacrality of the Tise region in the Zhang-zhung period is at least 400 years later and fully systemised narratives of place are much later still. This is not to discount the possibility of oral traditions retaining bodies of knowledge over that time-span, but we know that the processes of recording such traditions do not simply involve a restatement of the original idiom in a new medium. Rather the new forms are framed by the imperatives of the recording period. Thus, when accounts of pre-mid-7th century Zhang-zhung culture were recorded in the 11/12th century (and later), that recording process was governed in the selection of relevant information to be articulated by newly prevailing imperatives (many no doubt unknown to us now). What seems the earliest reference to Tise/Mapham in Bön literature is in an early biography of Shenrab (the mDo ‘dus), provisionally dated by Kalsang Gurung to the second half of the 11th century.20 This work describes Shenrab as spending a year at Mapang (Mapham) teaching the Klu, before being given a welcome feast on “the summit of Mount Tise [by] the deity Ma sangs.”21 Yet this account is not part of a systemised Bön narrative of a Tise centred Zhang-zhung and the resident deity is not that of later accounts. Those who described Zhang-zhung culture may not have had much reliable information to proceed with. As Christopher Beckwith has pointed out in regard to the collapse of the Yarlung dynasty;

After the [Yarlung] empire collapsed in the mid-ninth century …, nearly all genuine knowledge about it died too. During the little-known post- imperial period of disunity the Tibetans created a new, Buddhist-centered culture … with … a new, almost completely a-historical account of the Tibetan Empire based on their new socio-religious agendas. The most important of [which] was the promotion of Buddhism and a Buddhist interpretation of everything, including history.22

As Beckwith also notes, we have only a very limited and exclusively Buddhist selection of sources that are unquestionably contemporary with the Yarlung dynasty (principally inscriptions and the Old Tibetan Annals: pt1288). Neither in this corpus nor in the Dunhuang material is there evidence for a sacred

20 Gurung (2011); also see Blezer (2010b). 21 Blezer (2011a: 132). On the Ma sangs, a class of deities, often groups of brothers, and frequently including mountain deities, see Nebesky-Wojkowitz (1975: 224). 22 Beckwith (2009: 221). bön and the mountain 345

Zhang-zhung Bön Tise. In the Old Tibetan Chronicles we have passing refer- ences to Mapham, but in contexts that do not confirm its sacrality. In her well-known cryptic poetry cited in the Chronicles, the wife of the last king of Zhang-zhung (and the sister of the Yarlung Tibetan king), makes no reference to the complex and there are no clear statements of its sacrality. There is per- haps a hint of local belief in her acknowledgment that the capital Khyunglung is claimed by others to be “gold and gem” when viewed from the ‘inside’ per- spective; a similar understanding is later applied to Tise.23 But the Queen is said to have been absent from her castle fishing at Mapham when one envoy arrived! This hardly suggests sacrality. (It might imply that the modern belief in the medicinal value of the fish in the lake was held then, but more probably carries a literary or symbolic meaning now lost.) The situation is similar in regard to Tise. The take-up of Buddhism in the Yarlung period included the adoption of the Meru-centred cosmology from Indic culture and similar understandings informed Bön cosmologies. As in Indic culture the cosmic mountain then served as a metaphor, the king was “firmer than the power of Mount Meru”,for example.24 But there was absolutely no suggestion that Tise was then associated with Meru or with any other construction, and there is no mention of it in a political context even as a symbol of Zhang-zhung or its regions. While many Bön texts do appear to recycle earlier material, they are highly problematic as historical sources. This is not only the case in regard to Tise, but even in regard to a Zhang-zhung heartland of Bön. There is for example, strong evidence that Bön origins lie not in or to the west of Tibet as their texts claim, but around the Kongpo Bönri area southeast of Lhasa25 (and perhaps even further to the northwest towards the sources of the Dri-chu/Yangtse river.) In this context it is certainly notable that the overwhelming majority of Bön monasteries were established not in western Tibet but in Kham and Amdo and that their leading personalities derived from these regions, not Ngari.26 (While there is an established northern route from Amdo to Ngari—by-passing central Tibet—this seems of greatest significant in a very early period of migra- tions.27)

23 Uray (2003: 248). 24 Kapstein (2000: 620); also see Walter (2009: 277). 25 Blezer (2011a); Karmay (1992). 26 On which see Karmay & Nagano (2003); also see Jackson (1978: 202). In discussing Bön in the region considered to have been part of Zhang-zhung, Jackson notes that in the 11th/12th century, Bönpo teachers considered their religion was new to the area. 27 Denwood (2005: 31–40, esp., 35). 346 chapter 13

None-the-less, if the later sources describing a Tise-centred Bön kingdom of Zhang-zhung are an invention, there certainly was a Zhang-zhung king- dom and no doubt there existed an array of religious beliefs and practices there.28 What seems most probable, if only because similar understandings were almost universal in pre-modern societies and throughout Asia, is that the religious world of Zhang-zhung involved localised territorial sacralities and deities of place. To transact with these powers there would have existed spe- cialists in ritual, including sacrifices, and probably some form of shamanic performers. As in the Indic world there were royal priests,29 and there is (late) evidence that funeral specialists were an elite group involved in some form of royal cult, although this may largely be a reflection of elite sources.30 No doubt there were other practitioners ministering to non-elite groups who have left no traces in the records. There are references in the Dunhuang sources to types of ritual practitioners termed Bön and Gshen (gShen), who appear in religious contexts—sacrifices, healing, funerals, and so on.31 Given that later Bön claims continuity with these activities, scholars such as Karmay cite these references to conclude that, “the assertion that the Bon religion never existed before the eleventh century or it was cut off from the early Bon tradition is totally misleading to say the least.”32 Many scholars are content—in the absence of an established alternative—to use the term ‘Bön’ to embrace these earlier manifestations of religious activity in the region despite their tenuous links to the later form of the religion. This usage has the benefit of simplicity if not of precision; without the name ‘Bön’ we simply do not have a specific term for pre-Buddhist practices on the Tibetan plateau.

28 This was probably also the case with the Yarlung court. The Maṇi bka’ ‘bum (60.147r) a c. 12th century gter ma text that recycles earlier material, describes in almost humorous fashion the existence of four competing traditions at court: Buddhist, Bön, Tantric, and indigenous (Haarh: 1967; 385). 29 Tucci (1955: 199–200) refers to them as Gshen gnan; Choephel (1978: 27) as sKu-bon; and Norbu (1995: xvi) as sku gshen. 30 The appearance of ‘Bön’ as a term associated with funerals remains common in the Himalayas; see, for example Schneiderman (2002: 245). In modern Sikkim, however, “all death rituals … remain strictly the domain of the lamas, as only the supra-worldly deities of Buddhism are now thought to be of help to the deceased, for the worldly gods of bon are only useful in dealing with this life’s worldly problems.”; Balikci (2008: 274). 31 For an extensive examination of this material, see Arianne Macdonald (1971); and the critique by Stein in McKeown (2010). 32 Karmay & Nagano (2002: viii). bön and the mountain 347

This issue is evidence of historical processes, not simply one that derives from the Western academic obsession with labeling. Buddhism was very much concerned with articulating its specific identity, and within Buddhism the var- ious sects displayed similar concerns. They not only distinguished themselves through unique lineages and doctrines but emphasised separateness through such means as distinct dress, as did the later Bönpo with their blue robes. Famously the Bönpo also distinguished themselves through the reverse dis- play of circumambulation, the reverse swastika and so on. Such distinctions, actual and symbolic, were statements important to the long-term institutional survival of these faiths. Yet there is no definitive contemporary evidence of a religion in Zhang- zhung calling itself Bön or promoting any such distinct identity. This suggests that Zhang-zhung did not have a systemised religion and associated institu- tionalised priestly class,33 and if this was the case then in their absence clan or charismatic spiritual authority presumably predominated. This openness meant that there was no reason to exclude influences either from wandering renunciates or clerical ambassadors from the systemised traditions of neigh- bouring realms, be they Hindu or Taoist, Persian, Indic, or anything else. And as in the unsystemised early Indic realm it is probable that there were regional specialisations and developments, hence perhaps Zhang-zhung’s reputation for funeral specialists (like Uddiyana’s for necromancers and Kongpo’s for poi- soners!)34 It appears that naming itself was part of the Bön systemisation process, for which Buddhism provided the model the Bönpo followed, although instances of influence in the other direction are also found. Both religions informed each other. Local sacralities, however, were a key area of contestation as symbols of local power. But while those sacralities dominated the understanding of landscape and its distinct features—particular lakes, mountains, and so on— their character was of purely local concern. Thus such contestations are rarely prominent in Buddhist sources unless they serve some larger symbolic or liter- ary purpose. Rather than classifying those unique elements of the local religious world under such rubrics as “Folk” or “Nameless Religion”,35 with the implication of

33 Contra the argument of Arianne Macdonald (1971). 34 Tucci (1980: 227) recognised the importance of locality as sources for different traditions, and the extent to which, “in all the important centres the traditions predominating locally have had attached to them new, more or less arbitrary connections with particular powers which have gradually claimed and found general recognition …” 35 A division that is naturally manifest in Tibetan Buddhist discourse with its divisions of 348 chapter 13 lesser efficacy, importance, or appeal to sophisticated social stratas, it may be more valuable—and historical—to acknowledge them as distinctive features of wider regional cultures. We may then emphasise that the introduction of World-religions involved the subjugation of those local elements and powers as part of wider political strategies and imperatives. If we were to isolate the key process in the religious history of the western Himalayas, for example, without accepting the elite perspective of a World-religion, then the overarching nar- rative would surely concern the subjugation of local territorial deities by the forces of World-religions—Buddhism, Hinduism, and Bön. Michael Witzel has demonstrated the possibilities of a regional approach in his identification of the distinct historical characteristics of the local religion of the Hindukush.36 Beckwith has similarly advanced an understanding of a Central Eurasian Cultural Complex37 and that this culture extended to Zhang- zhung seems probable. Certainly much of the rock-art revealed by Bellezza and others seems to point in the direction of Central Asian influences, albeit that he concludes from the absence of foreign epigraphy for the period that the Zhang- zhung region was “not open to major cultural intrusions” for the period from c. 100bce–650ce.38 But we may similarly identify a Western Himalayan Cultural Complex in the regions discussed in this work, one situated across the Kumaon, Garwhal, Kin- naur, and Chamba regions of modern India and impacting on or even extending into, the Tibetan realm of Ngari/Guge-Purang/Zhang-zhung. It represents an earlier layer of culture in what Vitali, following Walter, terms the “Indo Iranian

Buddhist practice (lha chos) and local or pre-Buddhist practice (mi chos). But these terms inherently manifest power imbalances; thus I avoid them even as heuristic devices. 36 “Kalash Religion”; available @ http://www.people.fas.harvard.edu/~witzel/mwpage.htm, most recently accessed May 2015. (This is an extract from Witzel [2004].) Also see Lecom- te-Tilouine (2009: 14–15) who distinguishes the Nepal and central Himalayan region as one in which the religious dynamic is mediumism, involving direct manifestation of, and the consulting of the gods, in contrast to the ordinary Hindu model in which the Brahmans consult the gods. 37 Beckwith identifies the central feature of this as comitatus, “an oathsworn guard corp” around a charismatic leader; Beckwith (2009: 223). Tibet as part of a Central Eurasian Cultural Complex is also a theme of Walter (2009). 38 Bellezza (2010: 71). I have previously argued in favour of following Samuel’s use of the concept of a cohesive and unified cultural tradition or ‘cultural complex’ to describe Zhang-zhung, a usage sometimes also favoured by Bellezza, see McKay (2012); and see Samuel (2000). Samuel’s stimulating, while obviously somewhat speculative article raises a number of important and interesting historical and theoretical points. Such an approach enables us to avoid perspectives overly influenced by the modern Nation-state system. bön and the mountain 349 borderlands.”39 Zhang-zhung, which seemingly contained elements of these and other cultures, may be best classified as a cultural frontier, open to differ- ent influences at various times and including diverse population groups whose disparate traditions were never fully reconciled. With sacred lakes common across the region, the sacrality of Mapham in the context of a Western Himalayan Cultural Complex seems unproblematic. But while Lake Mapham had sacred status in local sacred geographies in Zhang- zhung there is no evidence that this lake was more sacred or significant than a multitude of other features of the landscape and there is nothing in the sources to directly connect it to a pre-systemised Bön. The sacrality of Tise is certainly not axiomatic. Although there are vague associations of sacrality, particularly funereal, with the snowy mountains in general (that might reflect earlier beliefs), there seems no tradition of spe- cific ‘sacred mountains’ in Western Himalayan or Indic culture. As we have seen, Michael Walter even problematises the existence of sacred mountains in Tibetan culture before the imposition of the Meru model as part of the phyi- dar,40 and there is no incontrovertible contemporary or near-contemporary evidence that Tise was the ‘soul mountain’ of Zhang-zhung. There is only pre- sumption based on much later practices and texts that are both influenced by

39 See Vitali (1996: n. 269 188 and passim); Vitali credits Michael Walter with suggesting the term but see note 40. Iranian influence may well have reached the region, perhaps as a result of the suppression of the cult of popular deities under the Achaemenid King Xerxes (486–465bce). But Naga lakes are not, to my knowledge, known in Iranian belief systems. For a discussion of possible Iranian doctrinal influence, see Kværne (1987); also see Templeman (2002). I am indebted to Katia Buffretrille (email communication, July 2012), for pointing out that this description of the characteristics of the whcc could equally apply to the Amdo region and clearly this is matter for further investigation. It is not perhaps, a question of unique identifying aspects of a culture, rather the survival of that culture as a distinct phenomena that distinguishes it from that of regions where a World-religion holds more-or-less exclusive sway today, with local elements expressed only through that lens. 40 Stein (1972: 236) in proposing Bön as “largely borrowed from Hinduism”, does note that this would mean the assimilation of “Indo-Iranian elements” (such as sacred mountains?) in Zhang-zhung before the adoption of Buddhism. Similarly, if Bön was as Snellgrove proposed, a poorly understood form of Buddhism, that too could have brought such elements into Zhang-zhung. The possibility that the Turkic belief in sacred mountains could also have acted to introduce this belief in the pre-phyidar period might also be considered, not least in the light of Beckwith and Walter’s Central Eurasian Cultural Complex model. 350 chapter 13

Indic ideas and part of a wider contested construction of sacred geography vis- à-vis Buddhist agencies. Even if Tise was the abode of a local deity we cannot simply extrapolate that sacrality to a central position as the ‘soul mountain of Zhang-zhung.’ Zhang- Zhung was not a single unified entity but a trifederacy, comprising (shifting) “inner”, “middle”, and “outer” regions. If mountain sacrality then existed, we might expect each of these regions to have its sacred mountain and it is of inter- est that Bellezza cites a tripartite textual scheme (gNas chen gangs ri mthso gsum) in which Tise/Mapham is one of three dyadic sites along with Nyanchen thanglha/Namtso, and Targo41/Dangra. Bellezza also cites a distinct Bön sacred geographical triad comprising two mountains, Tise and “Pos-ri ngad-ldan”,along with the Lake Mapham. He states however, that it is another triad, Kongpo Bonri, Tise and Targo, “that best defines the sacred geographical basis of Bon and its cult of mountains”, while there exists yet another triad comprising Tise, Nyanchen thanglha and Amnye Machen, a grouping also known to the Buddhists and termed “Ti thang spom gsum.”42 As he suggests, the latter combinations would reflect the expansion of the Tibetan empire and the absorption of different regions under Yarlung rule. This process whereby political expansion meant incorporating both polit- ical and religious forms in the conquered region was, as noted, identified by Samten Karmay.Discussing a Dunhuang text concerning nine mountain deities and their associated local feudal lords, he stated that, “if one local chief annexes neighbouring territory he seems also to adopt the local deity of the annexed ter- ritory for propitiation in order to safeguard his annexation.”43 This does imply the possibility of a process in which Tise/Ma-pham accrued enhanced status in a series of expansions of power and influence of the polity around it; first as the dyadic sacred centre of a local community, then of the outer division of Zhang-zhung,44 of greater Zhang Zhung, and so on.45

41 Bellezza (1997: 293) states that, “rTa rgo”, a “range of nine glaciated peaks” [c.f. the “nine black mountains”?], was the bla ri (“soul-mountain”) of the Zhang-zhung kings and thus as significant as Tise to the Bönpo. On rTa rgo (Targo) see Berglie (1980). 42 Bellezza (1997: n’s 1–2 2–3, 16, 24–25). 43 Karmay (1996: 59–76, quotation from 63); also see, for the example of Orgyenpa subduing the gzhi bdag or territorial deity of Rudok as a religio-political act, Vitali (1996: n. 965 566. It is unclear to me if this gzhi bdag is, or could be, Gekhod; see note 45 below). 44 The three divisions were fixed neither in time or space; see, for example, the discussion in Dagkar (1997: passim). 45 That process might also explain how the deity Gekhod, whose origins seem to be as a local bön and the mountain 351

As in the case of Indic and Buddhist histories, the evidence indicates that until the second millennium we cannot elevate beyond the local the status of Tise/Kailas or its dyadic lake(s). While there were renunciates practicing in the region by the end of the first millennium ce., it is only in the second millennium that the origins of a great sacred mountain pilgrimage site can be found.

Ölmolungring

We have noted indications that the Buddhist ‘Other’ in the early phyidar period was the esoteric Tantric movement, indicating that Bön did not then represent a systemised opposition to the interests represented by King Yeshe Ö. But Bön soon became part of a network of competing sectarian identities on the Tibetan plateau and critiques of Bön on various levels of doctrine and practice became significant in the 11–14th century period. While there is much we do not know about this time frame, it appears that various factions among the landed aristocracy were engaged in a complex power struggle of which sectarian competitions were the most visible manifestation in our sources. The aristocratic factions formed alliances with charismatic religious leaders whose followers became loyal supporters of their patrons, with the monasteries that formed around the sects becoming regional power centres of the alliance. At the same time, much of the wider Buddhist programme was aimed at the conversion of landscape through the subjugation of local deities by charis- matic Buddhist ritualists in the tradition of Padmasambhava, notably in the western realms, Rinchen Zangpo, the ‘Great Translator’. The impulse for this process was probably the suppression of local political forces in addition to the Buddhacisation of landscape and society.46 While the historicity of individual agents or events in this process may be questionable, it seems highly unlikely that the entire process was a fiction manifest in a literary genre. Ritual perfor- mance directed at these ends must have been extant in that period, as it has been in the modern period (as seen in Chapter Eight concerning the activities of Khyung-sprul in Kinnaur in the early 20th century).

yullha associated with Rudok (see, for example, Vitali 1996: 384, 566), came to be installed at Tise; i.e., as a result of a take-over of the region by a Rudok-centred polity. 46 Clearly there were political imperatives surrounding the systemisation of Bön and as Samuel (2002: 13–14) pointed out, the Buddhist emphasis on the subjugation of landscape may reflect, “a time when the point was not, or at least not only, to suppress non-Buddhist deities so as to make the land of Tibet safe for the Dharma, but rather to suppress the non-Buddhist deities in order to subdue those who supported them.” 352 chapter 13

In this complex religio-political maelstrom, Bön was ultimately manoeuvred from the centres of power by Buddhism. Given its claims to antiquity it is tempting to conclude that Bön represented forces oppositional to the Indic faith and drew support from conservative elements through the articulation of ‘traditional’ plateau culture. But it is equally possible that their patrons were not central Tibetans, that their constructions were not as convincing, their political skills lacking, and so on. We simply lack evidence for firm conclusions. But it is particularly interesting that Blezer has found that;

… tropes of far-western origins of Bon, in regions such as Zhang zhung and ‘Ta zig’ … become thematic relatively late in Bon identity discourse. In fact, the earliest self-consciously Bon narratives (e.g. in the mDo ’dus and the Klu ’bum) … by and large ignore ‘Ta zig’ and Zhang zhung, which later became so intricately involved with Bon identity.47

Instead of the “grand come-from-the-west story” of Shenrab introducing Bön from Tazig to Zhang-zhung and its subsequent centrality there, the earliest accounts of Shenrab and Bön ritual activity arise in literature concerning south- west central Tibet.48 Yet it was far western origin that became a fundamental tenet of the Bön faith. Significantly, that founding mythology may have pre- dated the key Buddhist mythology of the contestation for control over the sacralisation of the landscape around Tise—their account of Milarepa’s vic- tory over the Bön sage Naro Bönchung in their magical contest over Tise. The Milarepa tale thus seems to be a counter-claim—one that had the great bene- fits of simplicity and dramatic narrative! The apparently earlier Bön myth was of a complexity that seems both extra- ordinary and unnecessary—except that it must have been necessary in some sense we cannot fully understand today. The myth varied somewhat49—this emerging Bön was doubtless still very much a collection of bodies of knowl- edge and there is no evidence of systematic collation at that time. But the construction centred on the statement that Shenrab had been born in Ölmol- ungring (’Ol-mo lung-ring), a place situated in Tazig, and that he travelled to Zhang-zhung (and briefly to Tibet in search of stolen horses). In Zhang-zhung he transmitted the Bön doctrine of his birthplace to the local priests (of unspec- ified religious identity).

47 Blezer (2011: 157–158). 48 Blezer (2011: 158, 167). 49 See, for example Martin (1999: 260), who notes a 12th century source that splits Tazig into two entities. bön and the mountain 353

The location of Tazig was not specified beyond its western direction and Bön scholars and pilgrims, as well as modern academics, have advanced var- ious suggestions, but whether Persia is meant or somewhere in the vicinity of modern Tajikstan as the name might suggest, remains unclear. Ölmolun- gring was said however, to be a part of this world (in Tazig), but also a heavenly realm. Descriptions of it actually bore an uncanny resemblance to the sacred complex of Tise/Mapham and its set of four river sources. The land of Ölmolun- gring was dominated by a central mountain, the crystal monolith Yungdrung gutse (gYung drung dgu brtsegs: ‘pile of nine swastikas’). Four rivers flowed from its base or from the lake below it. Their identifications varied somewhat with the source. By one account they were the Nara in the east flowing from the mouth of a lion-shaped rock, the Paksu in the north flowing from the mouth of a horse-shaped rock, the Kyimshang in the west flowing from the mouth of a peacock, and the Sindhu in the south flowing from the mouth of an ele- phant.50 The influence of Abhidharma cosmology is clear, but the early date of this claim does allow that the Bönpo, as part of their systemisation, may have asserted spiritual authority over the Tise/Mapham complex before sim- ilar Buddhist initiatives and that Kargyu claims thus were a later response to the Bön. But an early source (the mDo ’dus), which makes no mention of Tazig as a geographical entity, refers to a Lake Madros (i.e.; the same name as that of the Klu of Mapham) situated between this crystal peak and another,51 and describes four more lakes as existing in the four cardinal directions (as in the Chinese model of the central mountain). This source also has the trope of four rivers descending from animal mouths (not rocks resembling them). These are located in the cardinal directions; here the sets are Gangka, east, elephant;

50 Karmay (1975: 171–175); also see, however, the revised version of this article (Karmay; 1998: 105–106), where the order is given as Kyimshang-east-horse: Pakshu-north-lion; Ganga-west-elephant; and, Nara(dza)-south-peacock! The point is, surely, that while there are of course, only four cardinal directions, and the animals vary only slightly, there is an almost infinite number of variations around the general theme of four rivers, their directions, and their associated animals. This is an ideal schema, not a geographical account. (Huber [1999: 51], notes a similar ideal of four lakes and four rivers at Tsari.) No standard, or ‘original’ set is canonical in Hindu, or Bön (see, for example, Blezer 2011a: 125) sources and the version preserved in the Abdhidharma literature is not necessarily followed in Tibetan sources. See the discussion in Vitali (1996: n. 6 92), but unlike Vitali I do not believe these combinations are reconcilable in geographical terms. 51 This other peak, “sPos-ri ngad-ldan”,as noted, is identified by Bellezza as part of the sacred geographical triad of gNas chen ri mtsho gsum. It is not otherwise prominent, but see Karmay and Nagano (2003: 241). 354 chapter 13

Sindu, south, bull; Pagsu west, horse; and, Sita, north, peacock. Testifying to the many strands of construction, the text also has a contradictory arrangement, but it does specifically distinguish Ölmolungring from Tise.52 Karmay suggested that Ölmolungring represented the memory of the Tise complex, which location had been forgotten by the time the Bön systemised their faith.53 “Confident that there could be nothing ordinary about the origins of Bon”, he states, the later systemisers located Ölmolungring in an realm beyond Tibet.54 Other interpretations are possible; Ölmolungring may have drawn on the heavenly Indic model of Meru, whether as a manifestation of a much earlier element of Bön as a “poorly understood form of Buddhism” or as part of the later Tantric or phyidar Buddhist versions of the Indic model. The numerous correspondences between Tise and the Muzh Tagh Ata mountain in the Pamirs (noted in Chapter 10) might also be responsible for the confusion. But Dan Martin has pointed to the Bru clan as possible agents of transmission of this schema. The Bru, who became powerful supporters of the teachings of Shenrab, migrated westwards into western Tibet and Tsang in the late 9th /early 10th century.55 If the early Ölmolungring seems to have a worldly location somewhere to the west of the former Zhang-zhung realm it became, from at least the 13th century, an increasingly other-worldly location. It was, as Martin describes;

a paradise both on and beyond the earth, an exotic country, a place of religious origins, a visionary landscape … a quite original holy place rooted in the sacred biography of Lord Shenrab … a place interwoven with the sacred time of Bon’s founding moments.56

Thus we might conclude that if the Indic Kailas was a process of transforming the heavenly cosmology into the worldly, the Bön Kailas was a process of transforming the worldly into the heavenly.

52 Martin (1999: 263, 265). 53 Karmay (1972: xxx–xxxi). 54 “Ti-se and its vicinity is considered to be the holy land of Bon, yet this region is too ordinary to be the birthplace of gShen-rab since one can visit that place. This tangibility is unsatisfactory to the mind of the believer who prefers a mystical, unknown land that exists only on a spiritual level.”; Karmay (1975: 175). 55 Martin (1999: 268). 56 Martin (1999: 279); also see Karmay (1972: xxviii). bön and the mountain 355

A Modern Bön Tise

While descriptions of Ölmolungring may be the earliest specific Bön associa- tions of the Tise landscape with their faith, our knowledge of the emergence of systemised Bön is insufficient to properly contextualise these references. The appearance of a Bön pilgrimage guide to Tise by the 13th century esoteric prac- tioner YesheGyaltsen (Yeshes rgyal mtshan) indicates that they were part of the contested spiritual landscape of Guge-Purang in the post-phyidar period, but a Bön presence in the region was a shadow of that claimed for the Zhang-zhung period. The Guge-Purang Genealogies do contain an account of the suppression of Bön under King Yeshe Ö, with its followers driven into a house that was burned down and their texts being collected and thrown in a river. But the account, which occurs only in Buddhist sources, is much later and may be a fiction not least because Yeshe Ö was himself the holder of at least one Bön lineage. Vitali (allowing an earlier form of the faith), concludes that Bön had been largely driven from the region during the time of the Yarlung take-over and the subsequent suppression of Bön that is said to have occurred in the 8th century under the Yarlung King Trisong Detsen. There is evidence for regional clans adopting Buddhism in that period, with the only significant Bön presence being the renunciate followers of the (Dzogchen) Zhang-zhung snyan rgyud lineage57—whose doctrines to an extent overlapped with Buddhist tendencies. Otherwise there were only “infrequent or isolated cases of Bon-po presence” in Guge-Purang,58 and that situation prevailed there until at least the late 18th century. While Bön was prominent in eastern Tibet during these centuries (and among the nomads of the northern steppes), and it maintained a monastic basis in central Tibet despite a lack of powerful patronage there,59 it did not have an institutional or other prominent basis in its claimed heartland for nearly a millennium, if at all. In the late 18th century, however, the Rimed (‘universalist’), a movement or at least a more prominent tendency towards ecumenicalism, emerged in eastern Tibet particularly among the leading figures of the Nyingma sect.60

57 On which, see Blezer (2010). 58 Vitali (1996: passim, esp., 110, 140–141, 192, 214–215, 221–225, 410, quotation from 141). Also see Jackson (1978: 195–227, esp. 200–203), which indicates Bön missionary activities in that region, again connected with the Zhang-zhung snyan (b)rgyud transmissions. See Kind (2012: 226) for the 16th century visit by one Nyima Gyaltsen. 59 Karmay (1975: 186). 60 On which see Smith (2001: 227–272); Samuel (1993: esp., 532–543). 356 chapter 13

This was to be an extremely influential tendency in the development of Tibetan Buddhism; indeed Geoffrey Samuel concludes that, “Tibetan Buddhism today outside the Gelukpa order is largely a product of the Rimed movement.”61 One aspect of the movement was its embrace of the synthesis of traditions including Dzogchen and other Bön teachings. Of Bön teachers, Shardza Tashi Gyaltsen (1859–1934/562) was the most prominent figure among a number of charismatic eastern Tibetans who represent a new phase of Bön additional to that followed by more conservative tendencies at Menri, the central Tibetan institutional centre of the faith. The appearance of a new account of Tise, written in 1844 (the ’Dzam gling gangs rgyal Tise’i dkar chag), marks the first indication of a renewed interest in the Tise region within Bön. Its author was the Bönpo Karu Druwang (1801– 1861), whose authority derived from a vision of the site he claimed to have had as a five-year old. Charles Ramble describes the guidebook as “a tour de force of systemisation … [that] develops [Tise] into a veritable citadel of his religion.”63 It is actually this text that first presents the fully developed legends of a mighty Zhang-zhung Bön empire.64 Critiquing Buddhist accounts of the site it locates Bön origins in the distant past, describing how a sage (named k‘ri-lde ‘od-po), from Tazig took on the form of a vulture and flew to the palace of the Zhang- zhung king located on the side of Tise. He and his disciples then established a series of communities totalling 16,000 practitioners whose presence in the region lasted 2,500 years.65 This account of Tise has several points of interest to us. Notably, in the usual account of the four rivers and their association with the horse, lion, elephant, and peacock, the southern river is named as the Karnali (rma-bya k’a-’bab).66 We may detect here the influence of scientific geography, and it seems that amidst the traditional there is a concession to the growing awareness that the source(s) of the Ganges were to the south of the Himalayan watershed. But four lakes are enumerated: “Gur-rgyal”, “Ma-pang” [Mapham], “La-ngag” and “Gung-c’u dṄul-mo”, which are said to emanate from four cosmic eggs (a common Bön cosmological concept). Lake Mapham is also said to be the abode of two (female) Klu, Mapang [i.e.: Madros?] and the otherwise unknown Zichen

61 Samuel (2001: 537). 62 On whom see Achard (2008); Gorvine (2006); also see Thar (2002: 155–172). 63 Ramble (1999: esp., 17–21). 64 Blezer (2011a: 128). 65 Norbu & Prats (1989: 53–56, 117). In that the 1896 guide-book to the site was written by the 34th Drigung Kargyu hierarch, that may have been a response to the Bön text. 66 Norbu & Prats (1989: 115). bön and the mountain 357

(gzi-c’en). Interestingly Rakas Tal seems to have no inauspiciousness, here it is “the medicinal lake.”67 It is not entirely clear what agencies this work represents. What led the Bönpo to a new emphasis, perhaps even a new discovery of the Tise region is nowhere stated. There are occasional references to individual Bön pilgrims to Tise in this period,68 but the only indication that a wider promulgation of ‘sacred homeland’ mythology was underway might be that several prominent iconographical depictions of Ölmolungring date to the 19th century.69 But Karu Druwang does seem to be a key figure here for he is as Ramble stated, “the most important architect of what we may, for want of a better term, refer to as ‘Zhangzhung nationalism.’” The guidebook of Karu Druwang (who was born in the Hor region of Kham), was part of a wider impact he made on the western Himalayas as he undertook a series of pilgrimages throughout the region. In the last decade of his life, he returned to take up the abbotship of his home monastery of Norbu Ling (also known as rTing ngu).70 Another important figure, also born in Hor like Shardza Tashi Gyaltsen and Karu Druwang, was Spa Nyima Bumsal (sPa Nyi ma ’bum gsal: 1854–c. 192171), who became the abbot of Spatshang, the largest Bön monastery in the region.72 Spatshang had been founded in 1847 by a member of the Spa family (a prelate of the king of Hor), whose once influential position in western Tibet had declined.73 After some members of the clan had migrated east to the Hor

67 Norbu & Prats (1989: 12–16, 112). 68 See for example, Karmay & Yasuhiko Nagano (2003: 146). 69 On which see, Buffetrille (2009: 313–326); also see Karmay & Yasuhiko Nagano (2003: 172), this painting is undated here, however. 70 Ramble (2008: 481–502, esp., 481, 483–487). Ramble’s article, drawing on Karu Druwang’s biography, corrects a number of errors in Norbu & Prats (1989), notably their statement that he was born in the vicinity of Tise. Kind (2007: 208) records that, perhaps ironically, a disciple of Karu Druwang opened the pilgrimage to ’Jag ’dul in Dolpo; “a place of pilgrimage as great as Mount Kailash … so [to obtain all its spiritual benefits] one does not actually need to go to Kailash.” On this mountain see Kind (2012). 71 Nyima Bumsal’s birth year is usually given as 1854, but Karmay & Nagano (2003: 133), give the date as 1825. This appears a typographical error for they state that he died aged 67; i.e. c. 1892. But he met and was the ‘root guru’ of Khyung-sprul, who was born in 1897. Alay (2011: n. 63 216) gives his date of decease as 1922. Thus his dates seem to be 1854–1922. (I do not have access to his biography.) 72 Lhagyal (2000: 460). On the medical traditions of Nyima Bumsal and his monastery, including their modern connections with the Tise region, see Schrempf (2007: 110–111). 73 Vitali (1996: n. 217 162) refers to an ancestor of the Spa clan who spent three years 358 chapter 13 country, they were able to re-establish their position there and Nyima Bum- sal was a key figure, linking or at least known to most of the individuals behind the dynamic regional developments of the period.74 Thus in interpreting the renewed interest in Tise by the Bönpo, we must take into account the inter- ests and traditions of the Spa clan and their supporters in Hor, whose ancestral interests were in western Tibet. The tempestuous political climate of 19th cen- tury Hor, situated on the frontiers of China and Tibet, may well have stimulated that westward gaze. Bönpo interests in Tise then became most explicit in the figure of Khyung- sprul (discussed in Chapter Eight in the context of his activities in Kinnaur). He was again from Hor and born into a politically well-connected and pow- erful nomad family. Khyung-sprul met Nyima Bumsal, who became his ‘root lama’, when he was still a young boy and after he became a monk it was Nyima Bumsal who suggested that Khyung-sprul make his first major pilgrim- age to Ngari rather than to Tsari or central Tibet. This was because, his guru stated;

Although this was the region where Bon had originated, nowadays not even the name of Bon existed there [and] what’s more, I have heard that Bon po pilgrims face great difficulties in that region. For this reason I would be most pleased if they were to receive some support.75

Khyung-sprul’s first journey however, was not directly to the Tise region. De- parting his home in 1919, he took full monastic vows at the Menri monastery in central Tibet then travelled to Bhutan where he stayed some seven months under the patronage of Bhutan’s first king, Urgyen Wangchuk. He then travelled to the Buddhist pilgrimage centres of India, Nepal (where he remained for two years), and then again to India where he travelled via Delhi and Rewalsar to Rampur (Kinnaur). He spent just two months in Rampur, staying with patrons,

meditating in the vicinity of Tise; also see, Karmay (2001: 10), apparently concerning the same individual, who “went to a crystal cave on Mount Kailāsa and, having performed ascetic practices for three years, achieved union with his three tutelary deities.” 74 Karmay & Nagano (2003: 131–136). Also see this work (2003: 40) re “bDe chen sgang”, the hermitage (near Tingri) established by the Spa clan in 1040; also see, re their first monastery established in 992, Vitali (1996: n. 322 225). Achard (2008: 136–138) notes that Nyima Bumsal had given advanced teachings to Shardza Tashi Gyaltsen (1859–1935), the author of the Legs bshad mdzod, edited and translated by Samten G. Karmay (1991/1972). 75 Kværne (1998: 73; quotation from his biography); also see Alay (2011: 225); Blezer (2007); Millard (2009). bön and the mountain 359 preaching and studying astrology, before heading on to Tholing.76 While he cul- tivated a number of powerful persons on this journey, it seems exploratory and of little lasting impact. In 1930 however, after circumambulating Tise, Khyung-sprul made a second pilgrimage to the Buddhist holy sites of India. He then settled in Kinnaur for four years under the patronage of Padam Singh, the last raja of Bushahr. It was then, as noted, that he carried out the series of rituals that bound 16 Kinnauri yullhas to an oath to serve as Bön protectors,77 the kind of ritual landscape ‘taming’ activity associated with great Himalayan religious figures since at least the time of Padmasambhava. In 1935 Khyung-sprul returned to western Tibet where he settled at the Gurugem (gu ru gyam) monastery78 he had founded near what he believed was the old Zhang-zhung capital.79 Bön now had established, or in the emic view re-established, its western Tibetan centre. Khyung-sprul is a complex figure. He was a Bönpo, yet he went on pilgrimage to the holy places of Buddhism in India. His patrons included Hindu and Bud- dhist monarchs and religious hierarchs and he did not necessarily always iden- tify himself to others as a Bönpo. That may have been part of the wider Rimed ecumenicalism or a deliberate political strategy (and he was clearly a diplo- mat). Khyung-sprul was intent on locating ancient Bön landscapes through visionary processes, yet he was also a modernist. He took trains, used the Indo- Tibetan postal service, and quickly recognised the advantages of modern print- ing methods for producing religious texts. He was thus a very similar figure to other modernist visionaries such as Swami Tapovan and Kailas savant Swami Pranavananda, who he may have met in the early 1930s.80

76 Kværne (1998: 72–76). 77 Kværne (1998: 77–79). Millard (2009: 153) notes that “binding” of local deities was a major part of his activities, with 40 pages of his biography devoted to this. 78 On which see Karmay & Nagano (2003: 240–241). 79 On which issue see, Blezer (2007: 75–112), (2011a). Blezer, tracing the development of references to Khyunglung dngul mkhar (‘The Silver Castle of the Garuda Valley’ although as he points out, a less romantic translation as ‘sandy fort of the Garuda Valley’ is equally possible), finds that the earliest mentions of this site seem to locate it much closer to Tise. Also see, on the mythological location of the capital on the inner circuit around Tise at the time of Gshenrab’s visit to “rGyang grags”, Bellezza (2002: 62–63). There is an inner path around Tise, traditionally taken only after 108 circuits of the outer path. 80 Pranavananda (1984: 104), also see Blezer (2007: 84). While Pranavananda visited the monastery and refers to a Tibetan who must be Khyung-sprul, it is not clear if they actually met. The monastery is also described by Govinda (1984: 221–226). 360 chapter 13

Khyung-sprul’s biography makes it clear that he “made conscious and coor- dinated efforts to (re)open ‘Zhang Zhung’ and (re)establish Bon to what he believed to be its rightful place in the area.”81 Just as Taksang Repa’s journey had been to the lands he saw as those of Kargyu and indeed Tibetan Buddhist ori- gins,82 and just as Swami Tapovan considered himself in the land of the ancient rsis, Khyung-sprul’s was a journey of faith. These men all believed themselves to be spiritual archaeologists, ‘Treasure Finders’ uncovering the historical her- itage of their faith and restoring it to its rightful pre-eminence. They drew on the authority of tradition, a tradition that empowered the visionary aspects of their travels and which was understood to transcend normal time and space. It was that visionary authority that enabled Khyung-sprul to (re)locate the ‘Silver Castle of the Garuda Valley’ that he understood as the old Zhang-zhung capital and to establish Gurugem monastery there,83 creating a Bön centre from which the faith could spread its power over the landscape. What was once imagined, was now real.

Reflections

Systemised Bön incorporated many strands of local belief and identity—Garu- das (Tib: khyung),84 genealogies, pantheons, and so on—and assimilated them into a system reflecting its diverse origins and traditions. There is much we do not know about this process, but as the disjunction between the apparent early centrality of Mapham and systemised Bön’s centering of Tise demon- strates, the transformation from local to World-religion was a fragmented pro- cess involving complex re-imaginings and negotiated boundaries of knowledge and power.85

81 Blezer (2007: 86). 82 Schwieger (1996). 83 It seems to have been Khyung-sprul who informed Giuseppi Tucci as to the site of Khyung lung dngul mkhar; Blezer (2007: 83). 84 On Khyung, see Bellezza (1997: n. 69 80–83); also see Nebesky-Wojkowitz (1975: 256–258). 85 This transformation of local to World-religion sacralities was not unique to Tibet. Valerie Hansen (1990) notes the contemporary 12–13th century movement from local to regional deity worship under the Southern Song dynasty as a result of increased mobility among cultivators. The process by which there was, as Ramble put it (Ramble: 1998; 141), “a displacement of the village gods away from the surrounding territory and into a more abstract realm of divinities unrelated to places”, was stimulated by Buddhist, Hindu and the systemised Bön faith. bön and the mountain 361

The Bön contestation with Buddhism for authority over local sacralities demonstrates this complexity.The Ölmolungring construction seems a Bön ini- tiative rather than a response to Buddhacisation, for it apparently predates the Milarepa mythology. But competing narratives were part of the contestation process by which both religions ultimately sacralised the entire landscape. The result was that as Bellezza states; “Tibet is overlaid with an infrastructure of … gzhi bdag (‘master of the locality’), representing decentralised spiritual com- mand of the land, and a network of tribal and clan reference points.”86 While the Milarepa mythology and the historically evident movement of the Kargyu into Ngari were part of a Buddhist take-over of local sacralities, sys- temised Bön advanced its own claims, not least the powerful claim to historical precedence. Zhang-zhung, half-forgotten, already legendary, was a convenient peg on which to hang a complex mythology of origin that was probably with- out significant historical basis. A millennium later, those claims would resonate with the New Age demand for timeless antiquity, lost kingdoms, and place- based spirituality. What is certain is that the later construction by Khyung-sprul and those who inspired him was, once again, by strangers to the region. Just as the key Hindu agents in the construction of Kailas were from South India, and the Buddhists who centralised Tise were from Ü-Tsang, the key agents in the Bön construction were from Amdo and particularly the Hor region of Kham. Again we see the suppression of local voices, the lost Nagas.

Postscript: A Saivite Response?

The 11th–12th centuries landscape contestations were part of a wider process of identity formation that resulted in the emergence of sectarian divisions in Tibetan religion. What became the dominant understanding of the landscape, Tise as the mandala of Demchok, was established by the Kargyupa with the sup- port and patronage of local elites. Their conceptual model—a deity dwelling on a mountain at the centre of a field of power radiating out over a sacralised geography—was pliable and infinitely replicable, and it may have influenced Indic formulations of Kailas mountains. The modern understanding of a Hindu Kailas—the mountain as the abode of Śiva—was a comparatively late development in Indic thought, and it is notable that the Śiva-Kailas equation is unique in Indic culture. Major Hindu

86 Bellezza (1997: 41); also see for the wider context, Karmay (1996). 362 chapter 13 gods are otherwise beyond worldly territoriality. No other significant deity in Pan-Indian Hinduism is consistently identified with a defined earthly location. There are local claims and associations, not least those of Śiva with Varanasi and other sites, but the dominant cultural depiction of the domestic Śiva is that of the deity at Kailas, an idealised setting that is both earthly and heavenly. This Śiva at Kailas concept has no affinity to Tibetan concepts of a mountain deity, an understanding for which there are no close parallels within Sanskritic culture. Śiva is not understood as a local protective deity or worshipped in that class of rituals associated with such deities and there is no sense that he is man- ifest in the mountain itself, it is simply his abode. The statement represents not simply the fixing of the deity’s residence as part of an increasingly detailed nar- rative emerging over time, but rather a conceptual model, a set of signifiers that denote a wider claim to landscape sacrality and which constructs a Kailas- centred sacred geography. As a model, this Kailas could be relocated onto other landscapes, creating replicant sacred geographies, just as the Kargyu mandala model served as a device to articulate claims to authority over various land- scapes. The development of the Kailas model was late. David Chandler cites a Cam- bodian cult to Śiva as a mountain deity as early as the 5th century, with Śiva said to live on a Mount Mo-Tam (modern Ba Phnom?), a Pasupata site.87 But that mountain was not called Kailas, that identification is later than the Brahman- ical Tantric culture that spread to southeast Asia.88 As we have seen, various texts contain contrary information as to Śiva’s residence until well into the second millennium. But by around the 11th century we have texts such as the Vīṇāśikha Tantra which locate Śiva and his consort on Kailas,89 and by the early modern period that understanding is predominant. We saw that Indic renunciates were present in Ngari by the 13th century and probably earlier, and were thus witness to that period of complex con- testations over the landscape sacrality of the Kailas region. They were in no position to contest the Kargyu sacred geography construction, indeed (to gain patronage) most probably manifested a Buddhist identity or at least a Tantric character beyond religious boundaries. But as we have seen, the Śaivite renun- ciates are the leading candidates for having been the agents who identified Kailas-Manasarovar as the abode of Śiva, and subsequently deployed the broad concept of the Demchok mandala—a deity at the centre of a sacred realm—at

87 Chandler (2000:14–15). 88 Samuel (2008: 304–309). 89 Samuel (2008: 305–306). bön and the mountain 363 sites such as Manimahesh, just as the Kargyu deployed their mandala at other Himalayan sites. Their Kailas abode-of-Śiva model would appear to be a con- ceptual borrowing from that model successfully advanced by the Kargyu. chapter 14 The Tise (Kailas) Deities

The Kailas Deities

Before we discuss the modern period, we leave our narrative to briefly consider the deities associated with Tise/Kailas. While it is commonly stated that Hindus consider Kailas the home of Śiva and Buddhists consider it home to the Tantric deity Demchok, its historical association with resident deities is far more com- plex. While deities such as Śiva, Demchok and the lake-dwelling Madros are discussed elsewhere, the following brief survey of some other members of the Kailas pantheon demonstrates the multiplicity of deities that have been asso- ciated with the site. Their variety provides further testament to the fact that the mountain was understood in many different ways by different groups and that different cycles of mythology, symbols, and identity markers were applied to the mountain at various times. As Tucci stated in discussing the cosmogonic mythology associated with Kailas-Manasarovar;

The simultaneous presence of various schemes in the same text is note- worthy …; their separate elements evidently came from different locali- ties, social groups or traditions. These cosmogonic, theogonic and genealogical ideas, differing from place to place, have remained alive as long as they have continued to be transmitted as tradition in particular families; the claims to divine origin contained within them naturally promoted their continuance. In other cases such family traditions were taken over by other family groups linked to them by kinship or otherwise. Such influences or fusions are clearly recognizable if these myths are submitted to a careful examination …… The many-layered nature and multiplicity of cosmogonic myths, which never fully fuse with each other into a single account … are transmitted onwards in a literature clearly influenced by local traditions.1

While the ‘Buddhacised’ Kailas with its Demchok mandala is comparable to other such sites where that device is central to the Buddhist reading of the mountain Tise as a yullha mountain is, as we have noted, more problematic.

1 Tucci (1970: 221).

© koninklijke brill nv, leiden, 2015 | doi: 10.1163/9789004306189_017 tise deities 365

Mountains such as Nyangchen Tangla have clear and distinct yullha and the patterns of worship of that deity are in line with models of that formation. But while a Tise lha btsan is acknowledged it is not prominent. It is not even res- ident on the mountain. The 1896 guide-book to Tise for example, locates the residence of that “fierce” deity on a minor peak to the west of the mountain.2 As will be seen below, the disparate deities and classes of deities associated with the mountain cannot be reconciled into any systematic schema or hierarchy. Nor do we seem to be dealing with lesser deities on the periphery of a man- dala (of Demchok). What seems most likely is that there were numerous groups that associated their deities with the mountain and that these discourses were not in close communication but co-existed or competed. While we are deal- ing with traditions that are in a large measure oral, and textual sources are overwhelmingly Buddhist or Bön, it is also notable that there is no deity associ- ated with the mountain whose antiquity is demonstrably pre-phyidar, nor any continuity between the deities mentioned in the early texts and those of later texts, when a Tise lha btsan does emerge. This can be seen as further evidence that Tise was not sacred in an earlier or ‘pre-Buddhist’, tradition and evidence that a mountain deity was only envisaged there after the mountain was con- structed as sacred in World-religion traditions. In what follows we discuss the Tise deities that are, at least textually, the most prominent.

Tise / Ti tse

Tise, the toponym most commonly used by Tibetans is, according to Tucci, “the aboriginal name of Kailāsa, perhaps this name is to be related with Te Se known in Tibetan demonology as one of the northern spirits (sa bdug: [i.e., territorial deities]).”3 Expanding on this, he noted that se was the name both of one of the (five or six) original tribes of Tibet and of a class of demons,4 and he suggested that Ti was “probably” to be related to the word for ‘water’ still surviving in Kinnauri5 (although the application of that to a mountain seems at some remove6). He also identified T’e se “the god of the year”, who headed a

2 Huber and Rigzin (1999: 131–133). 3 Tucci (1940: n. 3 41). 4 Kværne (1997: 395) states that “Se (sad) is … the [Zhang zhung] word for ‘god’.” 5 Tucci (1949: vol. ii 714). 6 If it were a river name, Ti ought to be the final element of the name (not the first): I am indebted to Dan Martin for this point. For a list of river names in Nepal containing the syllable ti, see Witzel (1993: 21). 366 chapter 14 circle of 360 lesser deities, as an early local god onto whom Chinese astrological ideas were grafted and who was later associated with the mountain.7 Tucci’s conclusions, particularly in regard to the ‘Ti = water’ equation, have generally been followed by later scholars.8 Yet we might problematise Tucci’s conclusion for the etymology of both Ti and se/tse is complex and not necessarily exclusive in application to landscape. Ramble refers to a local deity in northern Mustang who is associated with water and named “Ti-se dkar-po”.9 It is also possible that se/tse/tshe is cognate with the Tibetan rtse = ‘peak’, which gives us ‘Peak of the Waters’; i.e., the source of rivers. Bellezza, however, notes that ti in the Zhang-zhung language can mean ‘north’ (or ‘seaside’[!]);10 thus ‘Northern (mountain) of the Se’ is possible (or even, ‘Peak of [the cosmological?] sea’). Bellezza also refers to a Bön text where the mountain is termed Te sed; which may in Zhang-zhung equate Te = snow, and sed = colossal form; thus ‘Snow Giant’.11 The range of meanings that the term se may represent has been subject to some discussion, notably in an article by Ramble which demonstrates that in addition to the se class of divinities, the term serves as both an ethnonym and a toponym in northern Nepal. The term is still used as an ethnonym for two communities of Thakali speakers from the villages Te and Tshug, in the region probably once known as Se-rib. Ramble concludes that, “the Se people may well have been long established in the region of Western Tibet before moving into Nepal.”12 Thus the se in Tise could refer to a deity, a community, or their territory. But there are texts that describe Tise as being known under a number of different designations according to region,13 and similarly there are references to other deities resident there. In addition to those considered below, Tucci noted Srid pa’i sman (or; A Phyi gung rgyal) a female spirit, or sky-goddess who is one of the original nine sisters of creation born to a water goddess.14 Nebesky-Wojkowitz refers to three other categories of deities resident there;

7 Tucci (1949: vol. ii, 714, 724); also see, Tucci (1940: n. 3 41). 8 Karmay (2001: n. 2 xxix), states that Ti-tse/se is a Zhang-zhung term for ‘water’,but so too is ‘ting’ or ‘ting nam’,and Karmay also cites two texts where Ting-tse is used “as an alternative for Ri-rab (Sumeru)”. Also see Takeuchi (2011: 127). 9 Ramble (1997: n. 2 488). 10 Bellezza (1997: n. 3 16). 11 Bellezza ibid. 12 Ramble (1997: 510); also see Turin (2002: 260) who states that, “there is good reason to posit Se/sem/seṃ/sɨm as a pan-himalayan descriptive ethnonym for people of Tibetan stock.” 13 Bellezza (1997: n. 3 16). 14 Tucci (1980: 216, 219). tise deities 367 those ruling the 28 lunar mansions (Skt. nakṣatra; Tib. rGyu skar nyi shur rtsa brgyad); the 13 Bon deities of the mGur lha bcu gsum class [sometimes ’Gur lha; cf. Gurla Mandhata?] and also a Tise lha btsan.15 The btsan are a class of powerful deity described by Tucci as;

tribal gods … located in a particular place, generally a mountain … over which they were considered to preside … the all-powerful spirit under whose sway other demons and entities were supposed to be.16

The identity of the btsan ruling over Tise/Kailas however, may be contested, indeed so unformulated is its identity that Vitali takes the view that refer- ences to him should be understood as representing “the local establishment.”17 Nebesky-Wojkowitz’s “Ti se lha btsan” is resident on Kailas, yet seems not fully integrated with the “Gangs-ri lha-btsan” whom Bellezza identifies as the pro- tective deity of (Gangs) Tise, for Gangs-ri’s “main abode” is on a peak to the west of Tise.18

15 Nebesky-Wojkowitz (1975: 223). The 1896 guide to the sacred site also refers to it as the abode of Seng gdong dkar mo (see Filibeck [1988: 67]). Bellezza (2005: 224) identifies this figure as the consort of “gshen rNgam-pa lce-rings”, associated with the Targo deities. (Nebesky-Wojkowitz [1998] has a reference to “Seng gdong dkar mo” in his index, but not the text!) 16 Tucci (1949: vol. 2, 727). 17 Vitali (1996: n. 685 414). Humphrey (1995: n. 11 160) notes that in Mongolia Buddhist lamas often conflated various minor spirits into Buddhist categories, with the result that the name and appearance of the mountain deity may be lost but it is still worshipped; this may be the case here. 18 Nebesky-Wojkowitz (1998: 223–224) does not give a source for his information. Bellezza (2005: 245–259, esp. 249–250, n. 466 330–331) notes the deity “Gang ri lha btsan” (e.g. [presumably] Ti se lha btsan) who lives on a peak to the west of Kailas with a small lake; the deity’s consort is “klu-mo mdzes-ldan (‘Beautiful Female Klu/Naga’)”, who “perhaps” dwells in Mapham. Bellezza considers this deity may have been a yullha but “there does not appear to be any literature devoted to him.” Bellezza (2005: Plate 7, between 488–489) shows a mask of Gang-ri lha-btsan that has survived the Cultural Revolution and is kept at Chos-ku monastery. I am advised by Katia Buffetrille (email correspondence, August 2012), concerning the guide-book (gnas yig) of Choyang Dorje (see Buffetrille [2000: 29, 57]), where a “Gangs ri lha btsan” is referred to as a zhing skyongs (‘Protector of the Field’). Like gzhi dbag, this term refers to an overlapping category. She notes that she found a thangka of this deity in Purang gompa in which the deity was depicted as “white and very peaceful”, in sharp contrast to the usual ferocious appearance of protective deities. 368 chapter 14

The 19th century Bön guide to the sacred site refers to a ruling deity Tise in the context of Shenrab’s descent to Zhang-zhung. This is said to have angered its gods (lha rigs), who unite under their leader, Tise. In what seems a post- unification of Tibet aspect to this myth, Tise and his forces are joined by those of fellow mountain gods, Nangchen Tangla and Amnye Machen (although all are subdued by the Bön founder.)19 This Tise deity is also referred to there in such contexts as appearing to Kargyu pioneer Dorje Ghuya Sangpa with an offering of gold that the Kargyupa used to build a hermitage.20 Bellezza cites three locally-collected offering texts to the lha btsan of Tise, but these are heavily overlaid with Tantric Buddhist ideas and motifs and are of no great antiquity.21 It is also interesting—particularly given the Bön claim to represent earlier sacralities at this site, their claimed ‘soul-mountain’—that Tise lha btsan does not seem to be assimilated into Bön in any way.22 In the 13th century Bön guide-book (dkar-chag) to Tise, author Yeshe Gyaltsen is informed in a meditational vision that the mountain is identical with the “highest site (gnas mchod) of the Body Mandala of [the bodhisattva] Samantabhadra.”23 Yeshe Gyaltsen’s guide-book also states that the original name of Tise was “Gangs-gnyan ya-bag sha-ra”.24 This name recurs in Bön texts, in for example, the context of their defense against the Sakya Pandita’s critique of the equation of Tise to the sacred mountain of Buddhist cosmology, with a 13th century Bön text (the bsGrags byang),25 locating the ‘real’ Tise in space.26 But we cannot be certain if this toponym related to a yullha, was part of an entirely distinct cycle of mythology, or was an invented claim that failed to be accepted.

19 See Tucci (1980: 240); Bellezza (2005: 47–49). They cite the Ti se’i dkar chag by Karu Druwang. 20 Filibeck (1988: 75). 21 Bellezza (2005: 250–261). Bellezza notes that one text, a Drukpa Kargyu gsol-kha, refers to Bhutan’s Zhabdrung Nawang Namgyal and thus must be dated after the first half of the 17th century. The other two texts, which seem to have informed the Kargyu gsol-kha, are associated with the Nyingma and are somewhat earlier. 22 Bellezza (2005: 250) observes that there is no Bön literature on him. Even the modern Bön ‘Nativist’ Namkhai Norbu (1995, 2009) does not refer to him. 23 Loseries-Leick (1998: 160). On this text also see Ramble (1999: 17–18). Huber (1999: 29) notes that this primordial deity in its bodhisattva form was used in Chinese conceptions of sacred mountains. The bodhisattva was associated with the great Sino-Tibetan sacred mountain Omei Shan, in Sichuan; see Robson (2009: 53). 24 Ramble (1999: 17). 25 Lhagyal (2000: 434); also see Martin (2001: 46–47). 26 Dagkar (1997: n. 9 3). It is also mentioned in the 19th century Bön dkar chag of Tise; see Loseries-Leick (1998: 161). tise deities 369

We are in the presence here of different mythological cycles, culturally related, but available for application to whichever deity is appropriate. What seems important here is not that we can isolate one meaning and judge it correct, but that the terms encode a number of meanings (something dis- cussed more fully in the section on toponyms in our Conclusions). These terms can thus be understood in different ways by different peoples and lineages, and each understanding might contain surviving fragments of earlier cycles of myths but also serve as a basis for further mythological development as part of either distinguishing or integrative strategies. Ultimately, as we have observed in the case of Śiva, contradictions are conceptually resolved in the understand- ing that deities manifest, or are present, throughout time and space, space both worldly and other-worldly.

Gekhod and the Bön Deities

Specifically Bön understandings of Tise as the home of deities also indicate the existence of distinct bodies of knowledge, both systemised and unsystemised, and the multiple uses of mythological and symbolic devices in the understand- ing of the deities. While we have seen that the earliest Bön reference to a deity on Tise is to one of the ma sangs gods and some Bön sources recognise a Tise as the deity of the mountain,27 the mythology is actually more developed in regard to the residence there of the Gekhod deities.28 Gekhod is one of the five main protective, or tutelary deities (yi dam) of the Bön,29 and is said to have descended onto Tise in the form of a yak, which disappeared into the mountain. Karmay states that the “central deity that was worshipped by the people of Zhang-zhung was sKu-bla Ge-khod whose residence was Mount Ti se.”30 Gekhod has 360 subordinate deities,31 and is thus reminiscent of Tucci’s “T’e se”, ‘the god of the year’. But there are also 360 divinatory spirits (bdud lha) known to the Bön.32 That number recurs frequently, Bönpo claim their doc-

27 Karmay (2001: n. 2 xxix). 28 On the Gekhod class of deities, see Achard (2004: 249); Bellezza (2005: n. 197 399–400); Kvaerne (1995: 80–86); also see, Karmay (1972: 196–198); Norbu (1995: esp., 262–263, n. 87 286). 29 ‘The Five Excellent Ones of the gSas Citadel’ (gSas mkhar mchog-lgna); Karmay (1975: 197). 30 Karmay (1975: 18); also see Kværne (1995: 80–84). 31 Karmay (1975: 198). 32 Tucci (1980: 234) Karmay (1975: 198). 370 chapter 14 trines were originally written in 360 languages,33 a number of other mountain deities have a retinue of 360,34 while in Himachal Pradesh 360 village gods are believed to assemble at Kulu for an annual festival.35 The number 360 thus has various associations according to different religious understandings. Like 13, 84, and 108 (among others), it is an auspicious or symbolic rather than necessarily a precise or actual number. Tucci’s association of Tise and a deity with a retinue of 360 lesser deities was not mentioned in Réné Nebesky-Wojkowitz’s classic study of Tibetan deities, although he cited Tucci’s work. Nebesky-Wojkowitz did however, identify this model as common to two other great mountain deities of Tibet, Nyanchen Tangla and Amnye Machen.36 The application of this schema to mountains at the extreme east and west of the Tibetan plateau as well as to the more central Nyanchen Tangla seems to suggest, or is intended to imply, a certain cultural unity across the plateau, and might thus be dated as post phyidar, and even to unification under the 5th Dalai Lama. While others might be identified,37 we shall note finally that Per Kværne refers to another Bön deity “intimately associated with … Mount Ti Se” and said to reside on the summit of Mount Meru; this is Nyipangtse (Nyi pang sad), the “Guardian of the (rdzogs chen) Doctrine.” Nyipangtse is subsidiary to the Gekhod ruler and is something of a composite, appearing in different forms under different names which include Tsangpa (Tsangs pa), the Brahma of the Indian Buddhists.38 Further evidence for influence from neighbouring cultures—and the mobility of tribes and their deities as well as what seem processes of cultural fusion—is the figure of Ge khod dbang thang dkarpo, the yullha of Rudok and the sacred mountain of the Mongol or Central Asia Turkic tribes, whose forces once occupied Rudok.39

33 Tucci (1980: 217). 34 Nebesky-Wojkowitz (1975: 207–209). 35 Sharma (1990: 139). 36 Nebesky-Wojkowitz (1975: 205–213, 223–224). 37 Loseries-Leick (1998: 161) for example, refers to a 14/15th century Terma text (the Tshe dbang gnas brgyad kyi bstod pa), in which the Bön protector deity Tsewang Rigzin (Tse dbang rig ‘dzin; [on whom see Bellezza: 1997; n. 42 414]) resides on the peak of Ti se. 38 Kværne (2007: 395–400); also see Bellezza (1997: n. 77 84). Kværne identifies Me-ri as the tutelary deity of the Zhang zhung sNayn rgyud lineage of Dzogchen, on whom also see Loseries-Leick (1998: n. 46 164). Nathan Cutler identified Nypangtse as the Bön equivalent of ‘Ti se lha btsan’. See on this and other relevant Bön deities, Cutler (1996: 177–178, 201–214). 39 Bellezza (1997: n. 134 144); also see Vitali (1996: 384). tise deities 371

Kvaerne distinguishes two overlapping forms of Gekhod, Welchen and Meri, as being closely associated with Zhang-zhung. The many forms and attributes of their names seem non-Tibetan, while the entourage of Welchen includes another important deity, Ati Muwer, understood as the sage who first taught the Welchen Gekhod cult and whose body may be identified as the mountain itself. With each deity having its own entourage, as well as consort, there is evidence of the exercise of considerable scope in the expression of a vast array of local deities and those of various traditions, tribes, and regions.40

Kubera/Kuvera

The tendency to amalgamate the characteristics of local deities into those of a single deity also seems manifest in Kubera, the deity most closely associated with Kailas in the earliest Indic texts. The etymology of his name is uncertain,41 but he is known in both Hinduism and Buddhism, and worshipped by the Bön under the name Nor bdag chen po kubera.42 Kubera has attracted a colourful mythology within these traditions and while depicted in various forms he is most commonly pot-bellied, and is said to be fond of drinking. There are various lists of his family members, most notably he is the half-brother of the demon Ravana.43 Kubera first appears in the Śatapatha Brāhmaṇa (xiii. 4.3.10) where he is the leader of a band of thieves but he later becomes chief of the raksasas and yaksas; minor demi-gods or demons associated with fertility and with water. His ascent culminates in his becoming the god of wealth and treasure. He is also known by his patronymic Vaiśravaṇa, and these are separate deities in some formulations. Kubera is, furthermore, identified as one of the guardians of the four directions in the Buddhist tradition. He guards the northern quarter,44 along with his position as god of wealth, where he apparently absorbed Nor (lha), an earlier such deity in the Tibetan tradition.45

40 Kværne (1995: 80–86); also see Bellezza (2005: 145, 249–256); Cutler (1996: 160). 41 For a recent discussion of the etymology of Kubera; see Parpola (2004–2005: 146). 42 Nebesky-Wojkowitz (1975: 81). 43 On Kubera particularly in the Hindu tradition, see Sutherland (1991, esp., 61–68); Misra, (1981: 59–71); Dowson (1928: 173–174). On Kubera in the Buddhist tradition, see Nebesky- Wojkowitz (1975: 68–81); Getty (1914: 139–141). 44 Although he is absent from the earliest formation of this schema in Yajurveda 1.8.7: cited in Sutherland (1991: 65, n. 116 180). 45 Nebesky-Wojkowitz (1975: 69). 372 chapter 14

Though Kubera’s worship continues, in temples at Badrinath and Chamba for example, his role as the primary resident of Kailas in the Indic tradition faded away with the rise to prominence there of Śiva.

Towards an Understanding of Pantheon

This multiplicity of deities associated with the mountain might be understood via the underlying concept of a mandala, with the supreme deity being sur- rounded by his entourage of lesser deities. But more relevant is the cultural principle of henotheism. This requires the acknowledgement of whichever deity is being addressed at that time as supreme, and it may be that the tex- tually recorded suggestions as to the apparently supreme association of some of these deities with Tise/Kailas are a polite and ritual fiction that we need not interrogate. Historically, in the Hindu tradition, Kailas was Kubera’s mountain until it became Śiva’s, while for Buddhists Kailas was the centre of the Demchok mandala. In the pre-Buddhist and non-Sanskritic local tradition(s) it may not have been sacred at all, with Mapham/Manasarovar being the sacral centre. Only in the Bön tradition is the identity of its central deity a complex issue. Do the Bön deities assimilate a number of indigenous and/or local traditions or do they represent the application of a wider Bön cosmology to a specific landscape? Are they thus to be contextualised within landscape and cultural process contestations? Or might they represent Bön identity discourse, with increasing complexities and pantheons representing the development of Bön and its ahistorical and actually comparatively recent constructions of a Tise centred Bön Zhang-zhung? Clearly cycles of myths concerning the divine associations of the mountain exist in local traditions, traditions that were not absorbed into or claimed by Bön, and the culturally dominant agency of Buddhism has overlaid many of these associations. Yet the absence of any contemporary evidence for a pre-phyidar local deity, a yullha /gzhi dbag / zhing skyongs etc., allied to the apparent absence of any myth of conversion beyond the general understanding that local deities were subjugated under Buddhist control by Padmasambhava or his successors, only strengthens our conclusion that the mountain was probably not sacred until the phyidar period. The toponym Tise does serve to imply ancient sacralities, but the safest conclusion is that it is primarily an topographical identifier to which various sacralities could be attached, and that a lha btsan of that name is a late addition in accordance with wider cultural understandings held by one section of the population. section 4 Modern Histories

chapter 15 The European Construction of Kailas-Manasarovar

Introduction

The modern Kailas-Manasarovar attracts pilgrims from all over the world. But the process by which it came to be known and understood as a significant spiritual site by cultures outside of Asia was, until the early 20th century, a slow and lengthy one. In the wider sense the process of knowing can be traced to the ancient Greeks, particularly Ptolemy, whose accounts continued to underpin European geographical knowledge of the lands beyond the Indus down to the 19th century. But the first specific mention of Tibet came in Arab sources1 and as Mughal rule drew India into the Islamic world there was a growing awareness of the Himalayan land of ‘Thibbet’, particularly as the source of the medicinal spice, musk.2 The first certain reference to Manasarovar in European writings was by the Jesuit, Antonio de Monserrate, who attended the Mughal court in 1580.3 He referred to “yogis—who visit many territories but tell many lies and mix legends with facts”, and Himalayan priests describing a plateau over the mountains where, “[o]n the banks of a certain lake there—which the local people call lake Mansaruor—a certain tribe inhabits a very old city.”4 But de Monserrate’s report circulated only in Jesuit circles. It remained unpublished and was for- gotten until the early 20th century, as were most of the Jesuit and Capuchin reports on their Tibetan travels in the 17th and 18th centuries.

1 On which see Dunlop (1973). 2 On which see Akasoy, Burnett and Yoeli-Tlalim (2011). 3 Sven Hedin credited the first historical mention of Manasarovar as being a 1553 report by Mirza Haider, a general of the Khan of Yarkand. He referred to a fortress called Luk-u-Labuk beside an unnamed lake, which according to Hedin, “must have been Manasarvar [sic].” No fort is otherwise recorded there, however (although a Buddhist monastery might have been intended). Unless there is more to Haider’s report (which I have not sighted) than Hedin references, the identification seems questionable for this is a region of many lakes. Unable to confirm Monserrate’s reference, Hedin credited the first European account of the lake to an account by the trader Johan van Twist dated to 1638; Hedin (1917: volume i; 156). See however, Lahiri (2005:82) concerning maps from this period showing a central Asian lake as the source of 4 rivers. 4 Quoted in Allen (1982: 36); also see Wessels (1924: n. 3 46).

© koninklijke brill nv, leiden, 2015 | doi: 10.1163/9789004306189_018 376 chapter 15

European knowledge of the Himalayas rapidly increased after the establish- ment of their trading centres in India in the early 17th century, with occasional mentions of Manasarovar in the context of the origins of trade goods such as borax and musk. But there were no first-hand accounts of the site or any mention of the lake or the as-yet-unmentioned mountain having any spiritual connotations. Even the Portugese Jesuit Antonio de Andrade (1580–1634), who travelled from Badrinath to Tsaprang over the Mana-la in 1624–1625 and was the first European observer in the region, did not mention Kailas or the lakes.5 Not until 1715 did the first European, Italian Jesuit Ippolito Desideri, report sighting Kailas. He stated that the region was held sacred by the Tibetans, “on account of a certain Urghien [i.e.: Padmasambhava] who is the founder of the religion professed in Tibet.”6 He then described seeing;

an enormously high mountain, very wide in circumference, its summit hidden among the clouds, covered with perpetual snow and ice, and most terrible on account of the icy cold. In a cave of that mountain, according to legend, there lived the above-mentioned Urghien in absolute retirement and uninterrupted meditation. Not only do the Tibetans visit the cave, where they invariably leave some presents, but with very great inconve- nience to themselves they make the round of the whole mountain, an occupation of some days, by which they gain what I might call great indul- gences.

Desideri called the mountain “Ngari Niongar” and Manasarovar “Retoa”.7 Curi- ously he did not mention Rakas Tal but reported that Retoa was said to be the source of the Ganges and that, “this lake is greatly venerated by these supersti- tious people, and so from time to time they gather there on pilgrimage and cir- cumambulate it with great devotion …”8 While incorrect regarding the Ganges, Desideri did correctly identify the Indus and Brahmaputra rivers but as with de Monserrate’s reference his writings, though known to Sven Hedin, remained unpublished until the 20th century.9 By that time they were little more than a historical curiosity and had no real impact on the formulation of a European understanding of the region.

5 On Andrade and the Jesuits; see Wessels (1924). 6 Quoted in Allen (1982: 51). On Desideri, see Bargiacchi (2008); Sweet & Zwilling (2010). 7 This toponym is not otherwise known. 8 Sweet & Zwilling (2010: 170). 9 Allen (1982: 51–55). european construction 377

During the 1700s, Europe’s relationship with Mughal India underwent a major change. The power of the East India Company gradually increased until in 1757 the Mughal Emperor was forced to accept the Company’s rule over Bengal. From that base British power expanded across the sub-continent. The eic sought to locate and develop new markets, but also needed to protect the territory they controlled. Both factors brought the need for intelligence concerning their neighbours, not least those to the north. Apparently unaware of the Jesuits’ accounts of Tibet, the eic began to build up their own body of information on the Himalayan realms.10 In the British understanding, following Francis Bacon, that knowledge was power.11 Geographical information was at the centre of the early phase of the Euro- pean knowledge-building process. With the rise of scientific mapping it be- came possible to precisely locate every geographical feature and site of human habitation on a map—something of obvious military value—and the blank spaces on maps were rapidly filled in as a result of two interweaving and often inter-related imperatives, the political and the intellectual. Enquiries arising from political imperatives centred on information of military and diplomatic significance. Implicit and sometimes explicit in these reports was the possi- ble future need to move military forces through the territory at issue. Thus the reports of government officials concerning previously unknown territory identified the nature of routes between significant places and sources of water, fodder, and supplies. They also described the structures of power in the region, outlining the individual character, associations, and sympathies of the local elites in order to identify individuals or groups with whom they might develop alliances. It was characteristic of the political information gathering process that the earliest accounts of an area were obtained from or by local informants. Sub- sequently responsibility for compiling bodies of knowledge was assumed by Europeans, who collated information in forms appropriate for transmission among Europeans.12 That archive was then available for use by the civil and military officials concerned with that territory. Significantly for our enquiry this process eliminated indigenous concepts of sacred geography, which were of no practical value in the new bodies of scientific knowledge.

10 Notably through the journeys of the East India Company emissaries George Bogle and Samuel Turner, who were able to visit Shigatse in 1774–1775 and 1783–1784 respectively. 11 There is a considerable body of literature concerning the European knowledge-gathering project in India, which demonstrates the close relationship between power and knowl- edge in the colonial context. The foundational text of this approach is Said (1978). 12 On this process, see Dirks (1993). 378 chapter 15

But complementing this political knowledge project was a contemporary spirit of enquiry. While this often coalesced with the political project it pro- ceeded from rather different imperatives. Scientific advances allied to the Euro- pean Enlightenment had bought to the forefront the belief that the entire world could be categorised, classified, and thus understood in scientific terms. The founding of organisations such as the Royal Geographical Society (in 1830), drew on and stimulated the process of knowledge-gathering and systematic classification.13 The result of this spirit of enquiry was a body of writings which focussed on particular scientific aspects of different regions. Their geology, botany, climate, and so on, were recorded and ethnographical studies then undertaken that did allow some consideration of local belief systems. The contemporary spirit of enquiry also legitimised and often lionised the figure of the ‘explorer’.Individuals who travelled to territories outside European knowledge could gain both fortune and popular fame. This cult of the explorer was at its peak during the late 19th and early 20th century, with individuals such Richard Burton and David Livingstone gaining considerable renown. The unknown Himalayan regions were a particularly popular target for such explor- ers. Given the resources necessary for exploration and the scientific and intel- lectual frameworks in which geographical information had to be located in order to be accepted by the knowledge-collating bodies, it is not surprising that most of the explorers came from the same upper and upper middle-class background as government officials. They also travelled under the auspices of societies and organisations that were either government controlled or which generally co-operated with government. Thus, in many cases, the individuals involved in the process of exploration—explorers, government officials, and related authorities—were personally known to each other.14

13 The rgs (and other societies that emerged later, such as the Royal Central Asian Society) maintained close relationships with colonial administrators, many of whom served the society after their retirement. Among those mentioned elsewhere in this work who served as rgs President were Colonel Francis Younghusband, Lord Curzon, and Sir Richard Strachey, along with other famous ‘India hands’ such as Sir Henry Rawlinson, Sir Clements Markham, and Sir Thomas Holditch. For an amusing description of the mileau of the rgs, see Meyer and Brysac (1999: 311–315). 14 A classic example being the friendship between explorer and imperial officer Sir Francis Younghusband and 1899–1905 Viceroy of India, Lord Curzon. The famous Botanist Sir Joseph Hooker, a friend of Charles Darwin, is another example. In 1849 he explored Sikkim with Dr. Archibald Campbell, the Superintendent of Darjeeling. On the connections of the Orientalist Professor Giuseppi Tucci, see McKay (2015). european construction 379

While individuals could and did operate outside of this sphere, the require- ment that knowledge should be expressed within established frameworks if it was to be accepted as authoritative by political and cultural authorities meant that those who aspired to that acceptance worked within those frameworks of knowledge. So individual explorers validated their achievements by scientific observations, meticulously recording heights, temperatures, and other such data. As the Orientalist Richard Burton put it, “travel in unknown lands is a mere waste of time unless the traveller has suitable instruments and the skill to use them.”15 Explorers who disregarded such conventions, or appealed pri- marily to alternate—generally spiritually orientated, although also literary— audiences, were considered unreliable or inauthentic by authorities such as the Royal Geographical Society, and thus went unrewarded and often unacknowl- edged by such bodies. A nexus of governments, churches, universities, learned societies, publish- ers and newspapers provided support and stimulation for the knowledge- gathering process. But these bodies were not a unified force. While they inform- ed each other, each had their particular interests and responded to particular demands which did not necessarily coincide with those of other elements of the nexus. Missionaries, for example, were liable to be concerned with tribal groups that were of little concern to government, while publishers and news- papers required an element of human interest and emotion that was irrelevant to universities and learned societies (although in line with Enlightenment sen- timents some brief aesthetic judgement of landscape and culture was normally expressed in their reports). But while the disparate bodies of knowledge were not properly collated by any authority, the breadth of European exploration meant that significant bodies of knowledge relevant to particular authorities were produced in great quantity during the long 19th century. As a result, places such as Tibet that were largely unknown or unmentioned in the late 1700s had become well-known and even centralised in popular, academic, and political discourse by the late 1800s. The appeal of Tibet to the mystically and spiritually inclined is well-known, and has a long history. In essence it involved an understanding of Tibet as a realm in which the other-worldly was manifest in various forms and where spiritual wisdom lost to the West was preserved for the seeker after esoteric ‘truths’. That understanding was enhanced by the closing of Tibet to outsiders after 1792,16 which greatly hindered the European project of gathering empirical

15 Quoted in Deasy (1901: 1–2). 16 On which see Englehardt (2002: 229–246). 380 chapter 15 knowledge but left considerable space for imaginal constructions that only enhanced the lure of the ‘Forbidden Land’. In the sense that its closed frontiers prevented the development of empir- ical knowledge of Tibet keeping pace with that of other regions, Sir Francis Younghusband was correct in referring to it as, “the one great mystery that the 19th century has bequeathed to the 20th.”17 None-the-less, while the post-1792 restrictions all-but prevented access to central Tibet and particularly the capi- tal Lhasa, western Tibet remained relatively accessible. There were numerous routes across the passes from India and it was impossible for the Tibetans to prevent all Europeans from entering that region, which was under-populated and poorly-guarded. Consequently western Tibet was better known to 19th cen- tury Europeans than any other part of the country. Just three Europeans visited Lhasa during the 19th century,18 and none of them produced authoritative reports on the city. But more than a dozen trav- elled to western Tibet, almost all of whom were from the British imperial officer class. While the colonial government officially imposed a ban on any Euro- peans entering Tibet from British India, in practice it turned a blind-eye to such expeditions and even rewarded those who returned with useful informa- tion. A number of its officials crossed the frontier on shooting expeditions and other officers—as well as ‘native’ surveyors (pandits)—were actually sent by government on intelligence-gathering missions to western Tibet.19 The result was that the greater Kailas-Manasarovar region was both mapped and exten- sively explored by the end of the 19th century. The exploration of western Tibet is not our focus here. What is important is the form of knowledge that the British Government of India and its supporting structures gathered about Kailas-Manasarovar and the uses to which it was put. Knowledge imperatives did not concern spiritual or mythological matters, which were not considered relevant to scientific enquiry. The spiritual status of Kailas-Manasarovar was of no concern to 19th century colonial government.

Early European Discovery of the Sacred Spaces

What seems the first mention of the sacred associations of Kailas-Manasarovar in colonial sources came at the beginning of the 19th century, when the British

17 Cited in Waddell (1906: 1). 18 These were the Englishman, Thomas Manning (1811), and the Lazarist missionaries Huc and Gabet (1846). 19 McKay (2003a: esp., 637–638). european construction 381 were endeavouring to discern the geography of the upper Himalayas. (At the time of the compilation of Major J. Rennell’s famous 1782 map of India the location of the Ganges source was still unknown.) Captain Francis Wilford, in an 1801 article in Asiatick Researches entitled “On Mount Caucasus”, referred to Manasarovar and described how “Mahadeva [i.e.; Śiva] generally lived in Scythia, in which is situated the peak of Caliasa, the abode of Mahá-deva [sic].”20 ‘Caucasus’ at that time was used to refer to the general region of the Pamir range. The French physician Francois Bernier, who accompanied Mughal Emperor Aurangzeb to Kashmir in 1664–1665, had for example, reported that the source of the Indus river was in a mountain range called “Caucase” [sic].21 Wilford seems to have equated the region with northern Afghanistan for he also mentions Bamian and thus probably had access to the body of knowledge whereby that region was the site of Manasarovar. Captain Wilford (1761?-1822), a Bengal Engineers officer who had served as assistant to the Surveyor General before joining the staff of Benares Sanskrit College in 1794, was the first European, and perhaps the first person, to attempt to interpret the geographical references in the Puranas. As with so many sub- sequent scholars, this led him to conclude that they embraced a vast geograph- ical area and in this case that they confirmed a link between Druidic Britain, Ancient Eygpt, and the early Brahmans! Wilford discovered however, that his pandit has been creating new Puranas to accord with his theory and most of Wilford’s work was subsequently discredited, although his ideas occasionally resurface in obscure esoteric literature.22 The first British traveller to actually visit the Kailas-Manasarovar region was William Moorcroft (1767–1825), a veterinarian with the East India Company who was charged with improving the quality of the Company’s cavalry horses through the acquisition of Central Asian breeding stock. In 1812 Moorcroft travelled to Manasarovar disguised as an Indian pilgrim in company with the Anglo-Indian adventurer Hyder Jung Hearsey, who as we saw in Chapter 9 had been involved in exploring the Ganges source a few years earlier.23 Moorcoft’s reports, primarily concerned with trading issues, contained the required litany of precise scientific observations and also displayed an early

20 Wilford (1801: 507). 21 Hedin (1917: 204). 22 See for example, http://www.philaletheians.co.uk. On Wilford and his forged Puranic sources, see Leask (2000: 204–225); Trautman (1997: 89–93). 23 Moorcroft was also accompanied by a renunciate guide, one Swami Harivallabh who had visited the lake Manasarovar in 1796; www.swaminarayan.org/essays/2005/2204.htm; accessed 26 March 2007. 382 chapter 15 concern with Russian influence in the area. But references to the sacred nature of the region were entirely incidental to his main concerns. Manasarovar was mentioned as considered by Hindus to be “the most sacred of all places of worship, founded probably on the difficulty of access to it.”24 But otherwise Moorcroft only noted that he observed the ritual worship of Kailas (a sems offer- ing) by two Tibetan pilgrims, and that it was also “supposed to be the favourite residence of Mahadeva [i.e. Śiva].”25 Moorcroft’s companion Hearsey was sim- ilarly reticent in his diary, observing only that “Kylass Mountain appears very plain, distant and detached.”26 Both men did however, discuss the rumoured channel between Manasarovar and Rakas Tal, which they failed to locate. A more comprehensive account of the region was not obtained until the journey by Henry Strachey in 1846, with no Europeans known to have visited the area in the interval. Although he travelled in token disguise as an Indian, Strachey was a well-connected officer of the Bengal Infantry and the fact that he filed an official report on his journey suggests it was inspired by the British need for further information on the region following the General Zorowar Singh Kahluria’s invasion of western Tibet in 1841. Strachey’s report was a typical mixture of geographical, trade, and political information, but again had little to say on spiritual matters.27 Strachey was clearly familiar with the available literature on the region, including Arabic references, and was aware that it was of some religious status. His report noted that his Bhutia companions bathed at Manasarovar “by way of Dharm[a]” (i.e., for religious reasons), and that “[a] few of the Gorkhas visit Gangri [Kailas] on pilgrimage …”.28 But while he was informed by a new landscape aesthetic that appreciated the beauty of wilderness and thus, in contrast to Hearsey, acknowledged the “majestic beauty” of Kailas,29 Strachey was not concerned with its religious aspects. Strachey was concerned with the hydrological features of the Kailas region, concluding that following the work of Moorcroft and himself, allied to the use of trustworthy local informants, “the geography of the lakes is fixed in the rough,

24 Moorcroft (1816: 463). 25 Moorcroft (1816: 463). Moorcroft and Hearsey also used the then common term ‘Kailas mountains’ in its general sense, equating to ‘snowy mountains’. 26 The unpublished diary is quoted and discussed in Snelling (1990: 415–421, quotation from 418). 27 Strachey (1848). This report contains a wealth of detail on trade and administration in Ngari. 28 Strachey (1848: 63, 84). 29 Strachey (1848: 62). european construction 383 beyond all reasonable doubt.”30 Although he was correct in this conclusion, later travellers questioned his findings and generated controversies that were to fester over the ensuing century, playing a significant part in the scientific desire to explore the region further. But Strachey does mention one important fact in regard to the history of Kailas-Manasarovar. His report begins by describing an encounter with two renouncers he met in Askot (near Almora) at the beginning of his journey. Both men had just returned from a visit to Manasarovar. The first of them, “an intelligent, smart and decent looking person”, was suspected of being a spy and apprehended by the Tibetans. He was refused permission to circuit Kailas, allowed to carry out his ritual worship at the lake only under surveillance, and was then required to depart Tibet immediately. The second, Strachey wrote, was a “Jogi, black with dirt, and half fool; he accordingly met a better reception than the Sunyási and was allowed to extend his pilgrimage to Kailás without hindrance.” But even this ‘Jogi’ was forced to wait a year before entering Tibet, as the ban on all foreigners had even been applied to renouncers in the wake of Zorowar Singh’s invasion. Strachey also records meeting him again on returning to India, at which time the “Jogi … propounded his own ideas about the lake and mountain, which were silly and superstitious.”31 Thus did Strachey dismiss, unfortunately, information that is central to our enquiry.32 But Strachey then goes on to record that these two sannyasis were “said to be the only two who have succeeded in reaching Mánasarowár, via Byáns [approx: Kumaon], during the last two years.”33 So just two Indians, both renunciates, are said to have travelled to Kailas from Kumaon during 1845–1846. Clearly Zorowar Singh’s campaign had been followed by a tightening of the border, and the scarcity of renunciates visiting Kailas could date to 1841. But while we cannot

30 Strachey (1848: 63). 31 Strachey (1848: 87). 32 After Moorcroft, no 19th century European travellers seem to have taken account of the renunciates’ knowledge of Kailas. Although the East India Company had used renunciates to obtain information in the 18th century, armed conflicts with the sannyasis in the 1770–1790s coloured British opinions as to their character and in the 19th century they were increasingly mistrusted. See for example, oioc, e/4/799 (p. 143), report of 18 January 1849, concerning the grant of 576 Rupees by local officers to a “Fakir who undertook to supply information connected with Thibet whither he was proceeding on a pilgrimage”; the Company doubted “that dependence can be placed on the Faqueers veracity and we have many doubts as to the propriety of this expenditure.” The Company used renunciates in military capacities however, on which see Pinch (2006: esp., 104–147). 33 Strachey (1848: 2). 384 chapter 15 extrapolate from these figures, they seem curiously low if Kailas-Manasarovar was then of any great significance to Indic culture.

In the decades following Henry Strachey’s visit, numerous other Europeans crossed the passes from India to the areas around Kailas. Notable among them were Strachey’s brother Richard, the German geographers Adolf and Robert Schlagintweit, and Richard Burton’s later companion in Africa, John Speke. But while they were at least aware that the Kailas-Manasarovar region attracted pilgrims from Tibet and India, none of these travellers considered their visiting Kailas-Manasarovar significant enough to be worthy of any extended comment in the publications they left. That the sacred sites of western Tibet were of no immediate importance to the colonial state is further indicated by the fact that none of the remarkable spying missions of their pandits were directed by Government specifically towards Kailas-Manasarovar.34 But studies of the classical texts of Indian religions were developing within academia and the ics officer Edwin Atkinson detailed a number of aspects of Himalayan beliefs in his writings. Popular interest in the Himalayas was also growing, and the sacred region was briefly mentioned by Andrew Wilson in an account of his 1873 travels in the western Himalayas, Kashmir and Afghanistan. Entitled The Abode of Snow, it reached a wide audience and has been frequently reprinted. Wilson’s previous work included a popular account of General Gor- don’s campaign in the Taiping rebellion and while paying attention to scientific description, Abode of Snow provided the reader with numerous adventures, racy ethnographic descriptions, and opinions on matters religious and polit- ical.35 Like Strachey, Wilson described meeting a renunciate, in this case one travel- ling from Zanskar to Lahual, “accompanied by one attendant and with nothing but his loin-cloth, a brass drinking pot, and a little parched grain.” (When Wil- son refused to give him any aid, the renunciate understandably cursed him.36) Wilson’s understanding of Kailas however, was far less informed than Stra- chey’s. Describing the sunrise view from the cantonment outside Mussoorie, he writes of seeing;

the Jumnotri and Gangotri peaks, the peaks of Badrinath and of the Hindú Kailas; the source of the mightly sacred rivers; the very centre of the

34 On the pandits see in particular, Waller (1990). 35 Wilson (1876). 36 Wilson (1876: 272). european construction 385

Himálaya; the Himmel, or heaven of the Teuton Aryans as well as of Hindú mythology. Mount Meru itself may be regarded as raising there its golden front against the sapphire sky; the Kailas, or “Seat of Happiness”, is the cœlus of the Latins; and there is the fitting unapproachable abode of Brahma and of his attendant gods, Gandharvas and Rishis.37

The Tibetan Kailas cannot however, be seen from this viewpoint, and it appears from later references that Wilson is referring to a range of mountains rather than a particular mountain. He states that the “great cluster of mountains called the Tibetan Kailas (…) well deserves to be called the centre of the world”38 on account of its elevation and river sources, and in describing his travels up the Sutlej he refers to the Raldang (Kinnaur) Kailas as “one portion of the great Indian Kailas, or abode of the gods.”39 That abode he attributes to all Hindu deities but particularly to Siva, citing an 1853 translation of Kalidasa’s poem Birth of the War God.40 Wilson did play a minor part in beginning the process that would culminate in the crowning of Kailas as a major spiritual centre in popular imagination. But his brief and confused mentions of the peak suggest he had only a superficial knowledge drawn from a handful of scattered sources. He was actually far more expansive about a mountain on the Shipki-la border east of Puh on the Sutlej route, which he called “Lío Porgyúl.” “Even more remarkably than Kailas”, he wrote, “this gigantic mountain suggested an inaccessible dwelling-place of the gods …”41 So prior to the 1890s the sacred status of Kailas-Manasarovar had attracted only passing interest in the West. Explorers in the region had devoted consider- able attention to determining the location of the sources of India’s great rivers, the existence or otherwise of a channel from Manasarovar to Rakas Tal, and the correct toponyms for the various mountain ranges. But they had devoted very little attention to the sacred nature of the region. At the end of the 19th century, western Tibet gained greater prominence when it became the setting for the first popular account of the region, a best- selling travel book by Henry Savage Landor entitled In the Forbidden Land.42 This was a vainglorious account of the author’s travels in 1897, which ended

37 Wilson (1876: 41). 38 Wilson (1876: 174). 39 Wilson (1876: 107). 40 Wilson (1876: 258). 41 Wilson (1876: 148–149). 42 Landor (1899). 386 chapter 15 with his arrest and expulsion from Tibet. So exaggerated is Landor’s account of his adventures that much of it is clearly fictional and his geographical claims were swiftly dismissed by others better-informed. What is significant to our enquiry is that although it drew attention to western Tibet, Landor’s account of his travels refers only very briefly to the sacred nature of Kailas-Manasarovar. Despite his work being serialised in the Daily Mail before being published as a book, the populist Landor did not glamorise or mythologise the area as might have been expected if it was a sacred site in the Western understanding. This indicates that at that time the region still had no particular mystique in the European imagination for him to build on. On the contrary, Landor described the Kailas peak not in spiritual terms or within the wilderness aesthetic, but as hostile and forbidding, “uncomfortably angular … intensely unpicturesque.”43 For Landor, the chief interest in the region was as a remote setting for tales of adventure. The main association that might draw readers still lay in the geographical controversies surrounding the region, otherwise it was merely a suitably remote setting for his own fantasies of exploration as a symbol of racial superiority. Even in Landor’s work however, there are indications that Kailas was a site visited by renunciates. Landor followed Strachey’s route and he too encoun- tered in Askot an individual recently returned from Kailas, a long-haired, ash- covered renunciate. Landor describes how, “I saw the tall gaunt figure of a man emerging from a cloud of smoke … He seemed half stupefied and had very lit- tle to say for himself.”44 Landor it seems, had encountered a cannabis-smoking renunciate of the type common to the Himalayan region today. We may conclude therefore, that at the beginning of the 20th century, En- glish language works did not subscribe to Kailas as a place of religious signif- icance (and there were no Western networks of believers in its sanctity). The last work to which that applied was Captain Cecil Rawling’s The Great Plateau (1905). Rawling had crossed into western Tibet in 1902 and surveyed there in 1903 before joining the Younghusband mission to Tibet in 1903–1904. After its completion Rawling was sent with three other officers to survey the route to the planned British Trade Agency at Gartok, which passed by Manasarovar.45 Rawling’s account of the region again acknowledged only in passing the sacred status of the site. After briefly noting the enthusiasm with which his Hindu ser-

43 Landor (1899: vol. 1: 247). 44 Landor (1899: vol. 1; 30–31). 45 Rawling’s career subsequently took him away from Tibet and he was killed in wwi; on which see Davis (2011: 97). european construction 387 vants collected water from the lake, he devoted a page to describing the beauty of the Kailas peak, concluding that on account of that appearance;

No wonder, then, that this spot is believed by Hindus and Mahomedans alike to be the home of all the gods, that of the waters of its lake they drink, and that in its unexplored caverns they dwell; to them it is the Holy Mountain, and the most sacred spot on earth, a pilgrimage to which ensures both sanctity and renown.46

Rawling clearly erred in citing Kailas as sacred to Muslims, but he does give its sacrality to Hindus an elevated status. He also made one very prescient comment, that “Kailas … and Manasarowar … will in time become to Hindus as important a place of pilgrimage as Mecca is to Mahomedans.”47 Whether this suggests that he—and therefore others who served with Younghusband and were privy to British India’s future plans for Tibet—was aware of how the site was very soon to be transformed, is uncertain. But within a year, another British official was to publish a work largely devoted to Kailas-Manasarovar. That book would form the basis for the modern understanding of the site, and initiate a process that would lead to its becoming a major international attraction.

Charles Sherring and the Construction of a Modern Kailas

After Younghusband’s forces had invaded Tibet and fought their way to Lhasa they established a Trade Agency at Gartok.48 This was one of three agen- cies they created under the 1904 Anglo-Tibetan Convention, the others being located in central and southern Tibet. Proximity to the Kailas-Manasarovar region does not seem to have been a factor in the creation of the Agency. Arti- cle ii of the 1904 Convention, which allowed British subjects (who included Indians), free access to the Trade Agencies was intended to foster trade and there is no record of consideration being given to the possibility that it might also facilitate pilgrimage to Kailas. Significantly, there is also no evidence that there was any demand from British Indian subjects that they should be allowed to travel to Kailas. The two

46 Rawling (1905: 203). 47 Rawling (1905: 247). 48 On this ‘mission’ (/invasion), see the special issue of Inner Asia 14.1 (2012) entitled “The Younghusband ‘Mission’ to Tibet”. Also see Fleming (1984). On Younghusband, see French (1994). 388 chapter 15 categories of Indians who wished to visit the region were usually permitted to do so, for while closing their frontiers to outsiders in 1792 the Tibetans had continued to allow Indian traders from the Himalayan border regions their traditional access to markets in western Tibet. If satisfied as to their genuine nature and intent (and as noted they were particularly suspicious in the wake of Zorawar Singh’s invasion), they had also allowed renunciates to visit the sacred sites there. But the question of ordinary Indians visiting Kailas does not seem to have arisen. Article ii however, could be, and soon was, interpreted as allowing all Indians to cross the Indo-Tibetan border and visit Kailas. Despite its name, the Gartok Trade Agency was originally intended to func- tion primarily as a listening post to monitor any attempts by the Russians to exert influence in western Tibet, and the original intention was to station a British Indian Political (i.e.; diplomatic) Officer there. But the Government of India vetoed that for reasons of economy. Instead they placed the Agency under the control of the Government of the United Provinces, which provided a local government officer, Thakur Jai Chand,49 to serve as the Trade Agent. As the title was intended to suggest, that officer was charged with facilitating and encour- aging cross-border trade, although this was and remained insufficient to justify the cost of maintaining such a remote posting.50 There was in fact, very little for the Agent to do. Assisting Indian pilgrims was not officially a part of his duties, but Jai Chand and his successors did occasionally report on the number of pilgrims to the region, probably because they had very little else to report on! Under the imperial system, agencies under local control were subject to inspection by British officers. As the Gartok post had been placed under the control of local government it was the Deputy Commissioner at Almora who was instructed to make the first such inspection in 1905. The Almora dc at the time was Charles Sherring (1868–1940), who was educated at Westmin- ster and Cambridge before he joined the Indian Civil Service and arrived in India in 1899. Charles was the son of the Rev. Mathew Sherring of the London Missionary Society, who had served in Benares during the 1857–1858 rebellion and written histories of the sacred city, the Indian Christian church during the rebellion, and an encyclopaedia of Hindu tribes and castes.51 The Rev. Sherring had also been associated with various learned institutions in Benares including the Sanskrit College established by the British and he was sympathetic to Hindu

49 Jai Chand was from a Kumaon land-owning family and had previously served in Lahaul. 50 For details of the establishment and history of the Gartok Trade Agency, and a list of the men who served there, see McKay (1997: 158–165, 233). 51 Sherring (1858); (1868); (1872–1879). european construction 389 modernist reform movements. His son Charles, a young and ambitious officer, thus had a family background likely to have given him a far greater knowledge of Hinduism than was usual among ics recruits (although he does not appear to have read Sanskrit). Among the instructions Sherring was given prior to his departure to Gartok was that he should report on “measures required for developing the trade and pilgrim traffic between the United Provinces and Western Tibet.”52 As he would have discussed these instructions with his superiors prior to their being issued, this suggests that either Sherring himself suggested investigating the issue of pilgrimage, or that other elements of the imperial government did foresee some benefit from an increase in pilgrimage traffic to Kailas.53 While it was known that renunciates visited Kailas-Manasarovar, there was no economic or obvious political benefit to Government from their travels and thus no point in further encouraging them. But the British had long been aware of the economic role of pilgrimage in Indian society and Sherring or his superiors realised that if ordinary pilgrims travelled to Kailas they would contribute to local economies and development along the route to the border. Ordinary pilgrims however, were unlikely to undertake pilgrimages through undeveloped and unsafe territory. Thus after inspecting the Trade Agency and travelling past Kailas, Sherring returned to Almora and filed a report in which he recommended that roads on the pilgrimage route be improved in order that, “as elsewhere the devotee will be the pioneer of trade.”54

52 oioc, l/p&s/182–1677, report of Charles Sherring, 3 September 1905: I have not sighted any subsequent instructions to inspecting officers that mentioned investigating the pil- grimage. 53 The colonial goverment’s attitude to pilgrimage in India varied with time and circum- stance. In general, Christian influence opposed any government support for Hindu or Muslim pilgrimage, while public health considerations over the spread of epidemic dis- eases tended to favour regulation and restriction. But wider considerations of British relations with the Islamic world favoured support for the Mecca pilgrimage and after the 1857–1858 ‘Indian Mutiny’ government was extremely cautious in regard to any interven- tions liable to upset local religious sensibilities. In the region with which we are concerned pilgrimage tended to be encouraged as a means of stimulating the economy and to engen- der goodwill, as in the case of such initiatives as the early 19th century construction of safer bathing areas at Haridwar (which development was supported by local pandits), or the improvement of access to Kedarnath and Badrinath (which also served a strategic pur- pose). A proper study is called for, but see Kerr (2001) on the development of railways as a factor stimulating pilgrimage; see however, Ahuja (2003). 54 oioc, l/p&s/182–1677; also see oioc, l/p&s/7/179–1208, undated file note repeating this statement. 390 chapter 15

Sherring never returned to Tibet after his 1905 visit, but he became a key figure in the history of pilgrimage to Mount Kailas, indeed he may be identified as the primary agent in the construction of the modern understanding of the site. In 1906 he published a book entitled Western Tibet and the British Borderlands which was principally concerned with the sacred mountain. It opened with a photograph of Kailas as a frontispiece, and the Introduction stated that the region’s chief interest lay in its being the site of “the Heaven of Buddhists and Hindus … the abode of the great gods.”55 Superficially Western Tibet was a travel book describing Sherring’s journey to Gartok. But it was also a highly strategic text intended for a particular pur- pose, to stimulate pilgrimage to Kailas-Manasarovar. It contained a sophis- ticated continuation of his plea to the imperial government to assist in the development of the Kailas pilgrimage infrastructure, but most significantly it announced to an educated Indian audience that pilgrimage to Kailas was not a dangerous journey into the unknown. It was one that could now be under- taken within the sphere of British government influence and thus protection. Crucially Sherring’s text also proclaimed that this site was the most sacred place for Hindus to visit. In that sense it was, in its way, a British-authored version of a mahatmya! It brought Kailas to the centre, described the route to the sacred site, and outlined the blessings to be obtained from such a visit. Sherring presented an increase in Kailas pilgrimage from India as an inevita- ble consequence of the 1904 Treaty. While this increase could be relied upon to bring benefits to India in the form of economic development, it was prudent for government to plan ahead;

Whatever else has been accomplished by the Treaty of Lhasa, free ingress into Tibet is vouchsafed to every Hindu … A concession of this nature should appeal very strongly to every Hindu in the country, and, as soon as the facts become generally known through the length and breadth of India, we shall probably find an enormous number of pilgrims wending their ways to these sacred places … we may assert, without misgiving, that a very extensive pilgrim traffic is certain to spring up as soon as the Hindus of India realise that the way to Kailas and the Mansarowar Lake is no more difficult than the road to Badrinath. Further, this pilgrim traffic will be practically independent of the Tibetans, and it will be immaterial to the ordinary Hindu pilgrim whether the Tibetans regard us with affection or hostility. The holy places to be visited are only a short way across

55 See, for example, Sherring (1974: 30, 32, 36–44, & passim). european construction 391

the border—in fact, so short that food can be carried on the person sufficient for the journey—and no doubt in process of time rest-houses, so common in our own hills, will spring up to shelter the traveller from the weather.56

In pointing out that Indian pilgrims would spend most of their money within India, Sherring was using his bureaucratic skills to forestall government’s aver- sion to any unnecessary expense. His imperial officer readership would see the economic and political benefits of stimulating pilgrimage while retaining resources within Indian borders. Similarly in seeking government expenditure on road-building along the route to Tibet, he framed his appeal in terms of its ultimate economic gain;

Further, it may be remembered that trade takes time to develop, but for the pilgrim the only requisite is a good road, and when expenditure is being incurred, let us not forget that the devotee is sure to be the pioneer of trade.57

That trade Sherring implied, would be worthwhile. He noted that the leading Tibetan officials in the region had divided the customs tariffs of the various trade routes among themselves,58 and that;

[T]he Barkha Tarjum … has found it worth his while to buy over from the Daba Jongpen certain rights to collect dues in his territory, and the result is that part of the profits from the large mart at Gyanema now go into his pocket.59

The predicted increase in trade would, Sherring claimed, centre on wool (“there is no question that Mansarowar is the spot round which the whole of the wool trade of Western Tibet centres”), salt and borax. To save transport costs traders in these items would, he added, soon orientate towards the nearby Indian plains and its railway networks rather than far-off Lhasa.60

56 Sherring (1974: 145–146). 57 Sherring (1974: 159). 58 Sherring ibid. 59 Sherring ibid. The ‘Tarjum’ was the local headman; the Lhasa-appointed Jongpen (sic: Dzongpon literally ‘Master of the Dzong [fort/castle]’), equates to ‘Governor’. 60 Sherring (1974: 262–263). 392 chapter 15

Linking his appeal to government and to an educated Indian audience was Sherring’s modernist exposition of Hinduism.61 It was this construction that Western-educated and influenced Hindus (who were a comparatively pros- perous section of society), came to accept or to respond to, with Sherring’s appropriation of the authority for the definition of the Kailas pilgrimage typical of the period. The Hinduism that Sherring propounded included a strong divi- sion between Śaivites and Vaisnavites. While it was accepted that there were devout and moral Śaivites and that the deity himself might be seen in a good light, he was more commonly associated with the perceived debasement of Hindu thought. In contrast, Vaisnavites were seen as moral and orderly and it was believed that numbered among them were the majority of “respectable” Indians anxious to shed “later superstition” and return to “pure” original teach- ings. Thus Sherring wrote that;

Shiva is at all times a grim god, a god who inspires terror, and with whose worship there has been associated every form of debauchery, lust, and cruelty … to him is given a wife … goddess of death and blood, and in their abominable rites the Tantrik sect adore her with every form of obscenity, licence, and brutality. It is the Ling of Shiva which is worshipped through- out the length and breadth of India, and it is to this phallic worship that men and women flock everywhere—which cannot but have a degrading effect upon the morality of the nation.

In contrast;

The god Vishnu is the type of all that is best in Hinduism, and the worship- pers of this god number amongst them most of those who wish to throw off all the impurities and extravagances of the debased religion which are so common in India.

It was therefore, to the modern educated Indian, the Vedantist or the devout Vaisnavite familiar with the possibility of pilgrimage to Badrinath, that Sher- ring most appealed. The appeal contained an implicit economic aspect. The primarily Śaivite renunciates already visiting Kailas contributed little or noth- ing to the local economy and were in any case unlikely to read his book. His intended audience of English-literate middle and upper-class modernists how-

61 For an excellent example of Hindu history in the colonial understanding, see Sherring (1974: 79–80). european construction 393 ever, were of a class who would pay for local facilities and manpower, bringing revenue into the districts en route and ultimately, it was hoped, improve the economic condition of the whole community. Sherring thus emphasised two points throughout his book. Firstly he pro- claimed Mount Kailas as the ultimate pilgrimage for devout Hindus (and Bud- dhists) and more sanctified even than Badrinath and Kedarnath. Secondly he assured his readers that the journey to Kailas was no more difficult than that to those sites, and that the 1904 Treaty allowed ordinary Indians to travel to the Kailas region. “A concession of this nature”, he wrote, “should appeal very strongly to every Hindu in the country.”62 To his Indian readers, Sherring’s work carried considerable authority. Not only was he a member of India’s elite civil service and effectively the highest authority in Almora district, but he was also respected as a scholar, one able to quote from Hindu scriptures as well as his own. Further enhancing his authority was the fact that he had personally visited the sacred site. Thus there seemed no reason to doubt his eulogistic account of Kailas-Manasarovar’s sanctity, accessibility and beauty;

The view as one surveys this holy place, venerated alike by Buddhists and Hindus, is one of the most beautiful throughout the whole of this part of the country. The Mansarowar Lake, … and Rakas Tal, … make with their lovely dark blue a magnificent foreground to the range of the Kailas mountains at the back, while the holy Kailas Peak, Tise of the Tibetans, the Heaven of Hindus and Buddhists, fills the centre of the picture, full of majesty, a king of mountains, dominating the entire chain … The colour- ing of the rocks and the hue of the water, softened by the green of thou- sands and thousands of acres of verdant pasture-land, form a setting to the landscape which is indescribably charming, and although one misses the foliage of the forests, the colours are so exquisite in their brilliancy that they clothe the austerity of the mountains with a mantle that veils all their harshness. As one reaches the heaps of stones (each traveller should cast a stone on the crest according to universal custom) and sees the monu- ments (Chortens) erected by pious hands which mark the top of the pass, and the view bursts upon the sight, prayers and ejaculations break forth on all sides from the weary travellers, giving place later to a feeling of abso- lute contentment that they have now been blessed to see “what kings and many mighty men have desired to see and not seen.”63

62 Sherring (1974: 144). 63 Sherring (1974: 262). 394 chapter 15

Sherring’s book was the first European work to devote more than superficial reference to the religious significance of Kailas-Manasarovar and its appeal to the Indian reader was enhanced by his treatment of Indian religion, which even in its critiques was by the standard of the times, rational, balanced, and respect- ful. Its success also owed much to its readability. Sherring wrote in a style far less stilted than that generally favoured by government servants. He recounted colourful myths and legends surrounding the site and gave lengthy descriptions of local religious observances spiced with racy ethnographic portraits. In lend- ing the authority of a British imperial official to the understanding of the region as of the greatest sanctity his work was fundamental to the development of the modern understanding of Kailas-Manasarovar.64 But a critical reading of Sherring’s Western Tibet indicates the extent to which the text represented a new departure rather than illuminating existing understandings. Firstly, there was no canonical route to Kailas from India, but Sherring cites a description of the route through his district which he states is from a “careful and authoritative account … given in the Hindu scriptures.” The extract is taken however, from Atkinson’s Manasakhanda (the problematic nature of which we have seen in Chapter 6). Sherring makes extensive use of that text although he was aware of its probable recent composition (admitted, as noted, in Atkinson’s footnotes). He cites from Atkinson Śiva’s statement that “I dwell everywhere, but Himachal, [which Sherring specifically identifies as Kailas] is my peculiar seat …”.65 He also repeats the lengthier quotation (cited in Chapter 6) that begins “[h]e who thinks on Himachal …, is greater than he who performs all worship in Kashi”, and ends with the much-quoted lines “As the dew is dried up by the morning sun, so are the sins of mankind dried up at the sight of Himachal.”66 Oddly however, Sherring attributes the latter quote to the Ramayana rather than the Manasakhanda. This might have simply been an error, or evidence of his second-hand or superficial knowledge of the texts, but it could also be a deliberate attempt to strengthen the authority of the cita- tion, which was to be repeated by many later writers who relied upon Sherring’s authority.67 So Sherring’s “traditional” source was an Anglicised and compar- atively recent work of dubious authority and he conflated understandings of different mountains to advance the claims of Kailas.

64 This seems a prime example of the phenomena pointed out by Huber (2008: 36)—the “linkages … between the scholarly production of knowledge and the development of trends in religious practice.” 65 Sherring (1974: 48). 66 Sherring (1974: 37). 67 See for example, Singh (1982: 94). european construction 395

There are numerous other such constructions apparent in Sherring’s work. While acknowledging “a mingling of facts true of the country to the north of Cashmere with facts true of the country north of Kumaon”, Sherring clearly blurs the distinctions between the Garwhal-Kumaon Himalaya with its sacred peaks at Badrinath and Kedarnath, and the Kailas region. He also blends the identity of Tibetan and Indian pilgrims and their understandings of the region, and rarely provides sources for quotations and accounts he claims are taken from Indian texts. Sherring was also selective in his choice of material from those sources. Most significantly he did not mention a point that completely contradicted his central message; the fact that the Epic accounts of Kailas specifically state that the region was not open to ordinary pilgrims and could only be reached by the most spiritually advanced renunciates. There was also no canonical support for his grandiose central claim—the principal theme of the book—that, “there is no more sacred spot in all Hin- duism, or Tibetan Buddhism.”68 This is a significant point. Various texts and commentators have proposed such status for numerous sites, including Bena- res, Haridwar, Badrinath, Kedarnath, and Gangotri,69 but in the absence of any central authority within the Hindu world there is no agreed ‘most sacred place.’ That status in the understanding of any individual or religious community is influenced by regional origin and largely dependent on sectarian orientation. Similarly there is no ‘most sacred place’ in Tibetan Buddhism (as noted in Chap- ter 11). Sherring did not invent the sanctity of Kailas. But earlier European works describing the region had been far more circumspect; descriptions were largely empirical, local beliefs were briefly summarised, and the regions’ status was depicted as peripheral to the main regional secular and religious currents. This was consistent with the pattern of Indic textual references to the site which were from disparate bodies of knowledge with no overall consensus. While the centralisation of the region was actually a welcome departure in terms of academic enquiry—which was overwhelmingly focussed on urban religious perspectives—Sherring’s uncritical approach and explicit agenda greatly distorted indigenous understandings of the site, incidentally transform- ing the beliefs of a small sect of Saivite renunciates into those of Hindus in

68 Sherring (1974: 32). 69 See, for early examples of this tendency, Oakley (1905: 128) who states that, “Kedernath and Badrinath are the supreme objects of pilgrimage to devout Hindus.”: also see the reference by Pringle (1885: 739) “… the two most sacred spots to the Hindoo, viz, the source of the Ganges, or Gungootrie, and that of the Jumna, or Jumnootrie.” 396 chapter 15 general. His work elevated Kailas to a central status in Asian religious under- standing and in so doing it initiated a growing romanticisation of the region, which resulted in the modern understanding of that site being framed in largely mythical terms.70 The motives behind Sherring’s construction of Kailas-Manasarovar and the Government of India’s approval of his work are not difficult to ascertain. As noted, his superiors may even have initiated the idea, for opening Tibet to out- siders was part of their long-term regional ambition.71 But generating increas- ing tax revenue for the colonial state was a central duty of its district officers, something that was far easier in predominantly agrarian regions. In remote constituencies far from the patronage of the centre other approaches were needed. Exploiting forestry or mineral resources was one possibility, but that required reasonable infrastructure to allow the resources to be transferred to the plains. Developing pilgrimage was thus one of the few options open to Sher- ring. His book must be seen in that context. This is clear from his strong recommendation of the Lipu Lekh pass as the easiest route from the plains to the Kailas region, an itinerary which, he wrote, “combines directness with comfort.”[!]72 This route passed through the Almora district for which he was responsible and any increase in the pilgrim traffic there would bring increased revenue and employment opportunities to Sher- ring’s constituency.That would assist its development and bring that increasing prosperity which would reflect well on British administration, so Government were naturally keen to encourage such schemes, at least in principle. In addition however, the successful development of a region attracted the favourable attention of Government to the officer whose work brought about those improvements. Thus, while acting within the broad guidelines of British

70 The motives of those who developed that process are outside of the scope of this paper, but we may suggest, following Peter Bishop (1989), that this was in some ways an attempt to create a new fantasy of place in the Western imagination after the Younghusband mission had breached the isolation of the ‘Forbidden City’ of Lhasa and found it to be a prosaically human place. For an early example of this, see perhaps Rawling (1905: 129 & 151). He refers to “the sacred town of Rudok” being “as jealously guarded as Lhasa” Neither statement appears to have any factual basis. 71 On the possibility that the idea originated in higher circles concerned with long term planning of Indo-Tibetan relations, Pranavananda (1983: 5) cites Younghusband (source not given) as stating that; “Efforts should be made both in India and England to lead expeditions to the Himalayas to find out the best view-points of the mountains and make them known to the outside world. When those best spots would be discovered, they would be turned into and preserved as places of pilgrimage.” 72 Sherring (1974: 149). european construction 397 policy, Sherring was also advancing his own career prospects. And if his ini- tiative was within accepted parameters of action by frontier officials, more personal motives may also have been involved. Sherring probably realised the public demand for colourful accounts of the ‘Forbidden Land’ and hoped to emulate the sales gained by Landor’s book. Sherring’s influence on the Kailas pilgrimage has previously been noted by the late John Snelling, a Western Buddhist. In his well-researched popu- lar account of the pilgrimage entitled The Sacred Mountain, Snelling credits Sherring with being “the first of our Kailas travellers to be really awake to the extensive religious connotations of this rare quarter of the world.”73 But what Snelling found in Sherring’s work was actually the first modernist account of Kailas and it was thus the earliest work that confirmed Snelling’s modern pre- conceptions. Snelling credits Sherring with having been a cultured and scholarly individ- ual with a compassionate interest in the local people, “an example of the better type of British servant of India.”74 But the imperial archives reveal Sherring in a very different light, suggesting he was a man of staggering self-confidence, almost a caricature of the arrogant imperialist, who became accepted as an authority on western Tibet largely by enthusiastic self-promotion. His official report referred to the Tibetans as “ignorant barbarians”, and his attitude to fel- low imperial officials of lower rank was similarly disdainful and overbearing, as demonstrated by his bitter complaints over a fellow officer crossing the border for shikar. His faults were obviously apparent to his superiors; Sherring’s appli- cation to transfer from the ics to the (even) more prestigious Indian Foreign and Political Department was refused.75 He was soon transferred from Almora and in 1913 left the ics after what seems an unfulfilled career. Sherring’s book was, however, approved by his Government. As required by regulations governing civil servants, he submitted it for official censorship prior to publication. As the British were reluctant to reveal that they spied on Tibet, a number of references to the pandits’ journeys were removed before it was approved. Sherring then sent the resulting manuscript to the Government of India, confidently suggesting it should order 600 copies.76 In the event, after

73 Snelling (1990: 121). 74 Snelling ibid. 75 oioc, l/p&s/7/190–1240, Charles Sherring to Under-Secretary, Government of the United Provinces, 9 March 1906; National Archives of India, Foreign Department 1906, External b, March 154–157, various correspondence. 76 The Government of India did purchase books of which it approved; this acted as a means of subsidising specialist works. 398 chapter 15 lengthy enquiries to government departments from Burma to the Persian Gulf as to whether they wanted a copy of it, Government ordered just 30 copies.77 But in the absence of any published account of the Kailas region by an imperial official, that order was enough to ensure that Sherring’s book became the standard reference work for government officials. Western Tibet was published by Edward Arnold, whose books were exten- sively distributed in British India. It sold well in the public sphere with six printings of the original edition and it was frequently quoted by subsequent travellers in the region as well as imperial administrators, academic, and popu- lar writers on the subject, both European and Indian. Sherring’s construction of Kailas-Manasarovar thus became the basis for the predominant understanding of the sacrality of the region.

The Support of Sven Hedin

Sherring’s work was followed in 1909 by Three Years in Tibet, an account of the Japanese Zen monk Kawaguchi Ekai’s travels from western Nepal to Kailas- Manasarovar en route to Lhasa in 1900. Kawaguchi was an eccentric but learned individual who travelled to Lhasa posing as a Chinese monk in order to cir- cumvent the regulations excluding ‘foreigners’ (who included Japanese but not Chinese), from entering Tibet.78 Written in a simple and naïve style from a Bud- dhist practitioner’s perspective, Kawaguchi’s book did not have the impact of Sherring’s work and was more concerned with the Tibetan perspective than the Indic. But it did reinforce the image of Kailas as a sacred site and amplify it with colourful descriptions of the Tibetan pilgrims he travelled among. Kawaguchi did not actually write the lengthy (719 page) book; it was com- piled from interviews he gave to Japanese journalists and was originally seri- alised in 155 parts [sic!] in a Japanese newspaper. It was then published in Japanese in two volumes in March 1904. These were translated into English around that time and in May 1904, a copy of the translation reached the War Office in London. There its references to Russian weapons being transported to Lhasa were of considerable interest, for the Younghusband mission was then en route to Lhasa.79 (Sherring does not mention Kawaguchi’s work, and whether,

77 National Archives of India, Foreign Department 1906, External b, July 16–60, various correspondence, 1905–1906. 78 For an excellent account of Kawaguchi’s travels, see Berry (1990). 79 Berry (1990: 251–253); also see Berry (1995: 53): Fleming (1984: 267). european construction 399 as a provincial officer, he would have been supplied a copy of it by the Govern- ment of India is doubtful.80) Knowing of the numerous publications concerning Tibet that had appeared in the aftermath of the Younghusband mission, and also aware that Sven Hedin was about to publish “an excellent book of his travels in Tibet”, Kawaguchi was reluctant to sanction an English edition of his book. But he was convinced by famous Theosophist and Congress party co-founder Annie Besant that it was worth publishing and it appeared under the imprint of the Theosophical Society in Madras.81 In addition to his influence among Japanese readers, the choice of this English-language publisher ensured that Kawaguchi’s work found a readership among those who were particularly open to Asian religions and understandings of Tibet as a mystical land. Kawaguchi’s work thus widened the circle of persons informed of the great sanctity of Kailas-Manasarovar to include those of esoteric bent. It was, however, the works of the Swedish explorer Sven Hedin (1865–1952) which found the widest readership and which firmly entrenched the new con- struction of Kailas in the Western imagination. Hedin, inspired by the exploits of travellers such as his fellow Swede, polar explorer Baron Nordenskiöld, deter- mined from his boyhood that he would be a professional explorer. To succeed in his chosen profession he trained himself to withstand extremes of cold, as well as studying such arts as mapping and drawing. In time he sought out pow- erful men such as Younghusband and Curzon, as well as the Tsar of Russia and the Emir of Persia, to further his travels, and was able to make a series of lengthy journeys, most notably in our concern through western Tibet in 1906– 1908. With a mixture of “single-minded ruthlessness in the pursuit of his goals and craving for recognition and approbation”,82 Hedin succeeded in becoming perhaps the greatest of Europe’s Central Asian explorers. Hedin’s journeys bought him fame and due recognition83 particularly in Britain, where among his awards were an honorary knighthood, the Royal Geographical Society’s highest medals, and honorary doctorates from Oxford and Cambridge. But Hedin’s Nietzschean approach to travel and his some- times exaggerated or even unwarranted claims to new geographical discoveries increasingly alienated the British. His support for Germany in the First World

80 Coincidentally, however, at the time his work was published Kawaguchi was staying in the staff quarters at the Hindu College in Benares, where Sherring’s father had taught: Ekai (1979: viii). 81 Ekai (1979: vi–vii); also see Berry (1990: 267, 305–306). 82 Allen (1982: 192). 83 There is even a crater on the moon named after Hedin, located at 00 n-s; 75–80° west. 400 chapter 15

War marked a complete break with them and saw many of his honours with- drawn. His later support for Hitler further diminished his reputation and in consequence many of his achievements have been largely forgotten.84 At the peak of his fame however, Hedin enjoyed great celebrity, and as a pro- fessional explorer from a middle-class background must, he worked hard to promote himself. While he found generous financial patronage for his journeys from supporters such as Alfred Nobel, Hedin largely supported himself through his writings. These were often formulaic,85 and always verbose. Parts of his works were frequently combined or reprinted under different titles, and some- times rewritten for younger readers, making a comprehensive bibliography problematic. But he left a vast body of work—reportedly more than 30,000 pub- lished pages—and his writings were translated into many languages, including Chinese, Japanese, and even Yiddish. In the early decades of the 20th century, Hedin’s works enjoyed immense popularity. The demand for them in Germany for example, “seems to have been insatiable”, and in 1912, one million copies of his booklet warning of a Russian threat to Sweden were printed.86 British officials on the Indo-Tibetan frontier such as Younghusband and (Sir Charles) Bell, as well as the Military Intelligence Department, closely studied the Swede’s works and accepted them as authoratitive. Hedin was a man of great influence in both official and public spheres. After a failed attempt to reach Lhasa in 1901, the Swede returned to Tibet in 1906, intending to follow the course of the Brahmaputra from its source near Kailas eastwards across Tibet all the way to India. Curzon’s replacement as Viceroy of India, Lord Minto, was determined in the aftermath of the Younghus- band mission to maintain Tibet’s isolation and forbade him passage from India. But Hedin’s friend Younghusband was then serving as Resident in Kashmir. He ‘lost’ the telegram ordering him to prevent Hedin crossing the frontier, and the Swede was able to travel in Tibet for more than a year, including extensive exploration of the Kailas region.

84 For a biography of Hedin, see Kish (1984). Allen (1982: 193–237) provides a brief but balanced summary of Hedin’s life and achievements. 85 As Meyer & Brysac (1999: 316) write, “Hedin was to write nearly fifty books based on his own best-selling recipe; to a generous helping of descriptive writing, add a spattering of derring-dos, pepper with historical asides, baste liberally with meetings with the high and mighty, finally top off with one near-death experience. [sic]”. 86 http://www.silk-road.com/bibliography/hedinb3.html—A Sven Hedin Bibliography: com- piled by Daniel C. Waugh, (University of Washington, Seattle); accessed 9 May 2007. The booklet referred to is Hedin’s Ett varningsord af Sven Hedin. 1912 Stockholm: Bonniers. (Ein Warnungsruf. 1912. Leipzig: Brockhaus). european construction 401

Although he became the first European to undertake the parikrama of Kai- las, Hedin’s central concern in western Tibet was not with Kailas-Manasarovar as a sacred site, but rather to ascertain the precise geographical sources of the great rivers that arose there. He had however, closely studied the available sources not only of the geographical history of the region, about which he later wrote in overwhelming detail, but also its spiritual significance. Yet Hedin was not equipped with expertise in Indo-Tibetan languages, and thus had access to only a handful of local sources. He therefore seems to have relied primarily on the works of Atkinson and Sherring, while also drawing on Landor, Kawaguchi, Moorcroft, and Strachey. He may also have had access to unpublished colonial sources on the region through his contacts with Curzon and Younghusband, but as we have seen these were of little value in regard to spiritual issues. Although Hedin did critically analyse earlier geographical accounts of Kai- las-Manasarovar, he relied heavily on Sherring’s sacred construction in the absence of sources that could problematise it.87 With the commercial impera- tives of his writing he then used that construction of a remote sacred centre in western Tibet as a part of the appeal of his books. Not only were they describ- ing heroic adventures and geographical discoveries, they were also describing heroic landscapes and sacred worlds.88 In his classic account of western Tibet,Transhimalaya, published in 1909– 1913, Hedin quoted the lines from Atkinson’s Manasakhanda concerning “the dew … dried up by the morning sun”,and wrote lines that have become canoni- cal in the projection of Kailas-Manasarovar as a spiritual centre—lines that are sometimes now attributed to Hindu scripture!

There is no finer ring on earth than that which bears the names of Man- asarowar, Kailas and Gurla Mandatta. It is a turquoise set between two diamonds.89

The ageless sanctity of the region was frequently reasserted in his work. He described how for “thousands of years Siva’s guests have performed the round dance which was to lead them to an imaginary heaven”,90 and lauded the site as being;

87 While I have not checked Hedin’s entire works, he seems to acknowledge Sherring’s book only briefly; see Hedin (1913: 244). 88 “There were, of course, many imaginative landscapes produced at the turn of the century: Hedin’s heroic landscapes …”; Bishop (1989: 145). 89 Hedin (1913: 198–199). 90 Hedin (1913: 217). 402 chapter 15

Celebrated in grand hymns by the poets of remote antiquity, a dwelling- place of mighty gods, a mirror beneath the paradise of Brahma and the heaven of Siva, the goal of innumerable yearning pilgrims … the most wondrous lake on earth lies dreaming among the snow-clad summits of lofty mountains … As far back as traditions and legends carry us, Manasarowar has attracted the aspirations of men and their prayers. On its banks we tread ground which was already classic when Rome was founded.91

Hedin’s great contribution to the Kailas construction was to conflate Epic and Puranic understandings of the region into a timeless and static body of knowledge common to all Hindus. This supposed antiquity of pilgrimage to Kailas-Manasarovar was a recurring theme in Hedin’s work. Later, in 1917–1921, when he published a nine-volume account of his travels entitled SouthernTibet, the first volume, devoted to the sacred site, begins with the statement that;

To the Hindus and their ancestors South-Western Tibet has been known since thousands of years, and indeed we are here moving on ground which was classic when the foundation stone of Rome was laid.92

It was this emphasis on antiquity and the timelessness of the pilgrimage that was most characteristic of Hedin’s work. As Peter Bishop has demonstrated, this was part of a wider contemporary discourse which located Tibet outside of space and time, a conceptualisation that had political implications in the era of British colonial expansion. But by the time much of Hedin’s work was published it also partook of a nostalgia for a lost era, a pre-industrial spirituality, and a place beyond easy reach of the growing bands of European tourists.93 As Bishop observes, “[b]y the end of the [19th] century Tibet had become a fully formed sacred space for the West, with the axis mundi firmly estab- lished at the Potala in Lhasa.”94 Hedin had sought that centre in 1901, but with Younghusband’s ‘unveiling’ of Lhasa and the subsequent publication of several books describing the city, it ceased to be a major attraction to him. But in cen- tring Kailas-Manasarovar as a “fully formed sacred space for the West”, and in

91 Hedin (1913: 249). 92 In the first of the nine-volumes of Southern Tibet, Hedin discussed references to the site in classical, medieval, and modern Greek, Roman, and Arabian sources, as well as those from India, China and Tibet. 93 See in particular, Bishop (1989: esp., 149–190). 94 Bishop (1989: 185). european construction 403 relocating the imagined spiritual centre of Tibet from its capital to an exotic (and geographically significant) mountain-lake complex in one of its remote corners, Hedin revitalised the appeal of Tibet. His writings thus appealed not only to geographers and concerned officials but to a much wider readership. With the carnage of the First World War bringing a wide-spread doubt as to the value of industrial society vis-à-vis the pastoral,95 the appeal of his works only increased. As Bishop states;

The belief that Tibet contained meanings, and the urgent desire to partic- ipate in those meanings, was a constant background. Every travel account described an attempt to articulate these meanings, to locate them, to jour- ney towards them, and to embrace them.96

Hedin’s work was a classic manifestation of these tendencies, allied also to the standard 19th century use of scientific reportage. The latter—though its details might be disputed by specialists—gave Hedin’s work authority, as did his public recognition as an explorer. His interpretation of the meanings of Tibet and more precisely Kailas-Manasarovar, drew on that authority and fulfilled the demand of the reading public for sensational adventure, but also for vicarious meaning within their own conceptions of the world. In the context of the times it was a powerful message, and each strand informed the other, stimulating many later lines of enquiry, scientific and mystical.97 As with the earlier construction by Sherring, Hedin’s projection of the spir- itual nature and heritage of Kailas-Manasarovar may be easily deconstructed, for his preconceptions—and literary flourishes—are plainly apparent. Hedin was not a Sanskritist,98 and thus relied heavily on the Manasakhanda although aware of its doubtful provenance. In addition he used Sanskrit literature such as Kalidasa’s Megha-duta uncritically in English translation,99 and he also drew on the work of early Orientalists such as Lassen and Burnhouf, whose pioneer- ing translations and commentaries naturally lacked critical context. Thus, like

95 For an excellent study demonstrating the attractions of that landscape to those who had served in the trenches in wwi, see Davis (2011). 96 Bishop (1989: 188–189). 97 An open-minded but rational approach to the other-wordly images of Tibet was also characteristic of British officials posted there, with Younghusband’s explicit acceptance of the mystical being a powerful influence on that open-mindedness; McKay (1997: 207–208). 98 Correspondence from Hakan Wahlquist, Stockholm Ethnographic Museum, 20 May 1997. 99 The poem had been first translated into English in 1813 by the pioneering Sanskritist H.H. Wilson; Dodson (2013: 73–74). 404 chapter 15

Sherring, his mistakes are as likely to be errors of scholarship as slips of the pen, Manasarovar, for example, is referred to as “the lake […] created by the soul of Buddha”, rather than from the mind of Brahma as Sanskrit texts indicate.100 One consequence of the promotion of Kailas and the opening of a British Trade Agency at Gartok was that it encouraged Christian missionaries to be- lieve that they might be able to gain access to Tibet via its western regions. They understood from the works of Sherring and Hedin that the site was a cen- tre of Hinduism, which made it an immediate target for their activities. British India had always refused to allow the missionaries entry into Tibet, fearing their effect on the social system there, but the missionaries borrowed a tactic that had been used by the 19th century colonial government to gain access to Tibet—they used local employees. The most notable of these was the Christian convert Yunas Singh. His account of his journey to Kailas in 1917 reiterated the new image of the site. He spoke of “great numbers of Hindu worshippers” visit- ing Kailas-Manasarovar (which does not tally with the evidence of the Gartok Trade Agents), and followed the now common aesthetic in his explanation for the sanctity of the site; “[b]ecause of its beauty, no doubt, men have thought of worshipping it.”101 “To the Hindu”, Singh wrote, echoing Sherring and Hedin;

Kailas is heaven itself, and the pivot of the universe. It is the throne of Shiva and the abode of the gods in general. To visit it is to win the highest achievement of Hindu piety.102

Within this understanding, Christian missionaries continued to seek access to the region in order to confront its spiritual ‘Otherness’ (although largely without success). It was another Western spiritual movement, the New Age, that was to embrace the sanctity of Kailas and eventually bring about its transformation into a world, rather than regional, sacred site.

A British Colonial Kailas

The establishment of the British Trade Agency in Gartok was primarily due to the then Foreign Secretary of the British Government of India Louis Dane, who was closely involved in planning the Younghusband mission and a close friend

100 Hedin (1913: 198); the error is doubly ironic—Buddhism denies the existence of a soul. 101 Singh (2003: vol. 3, 507). 102 Singh (2003: 501), who also records an earlier visit to Manasarovar by Gerard Agnew of the London Missionary Society. european construction 405 of the key British officers involved.103 Dane had a particular interest in western Tibet and as we have seen had even considered extending the Indian frontier up to the Kunlun mountains (which would have brought Kailas-Manasarovar into British Indian territory).104 While it soon became apparent that the Agency was of very little political or diplomatic value, considerations of prestige demanded that it be maintained. While the agency was taken under imperial government authority in 1936 it remained a backwater, occasionally inspected by a British official. Relations with the local Tibetans were perhaps best summed up by one British officer who noted that the local administrator, while friendly, “gave me the impression that he would not be inconsolable if he never saw a white man again in his life.”105 While it is difficult to gauge the extent to which the disparate official and semi-official reports of the region were actually assimilated into the imperial system, or by any individual, the reports of Moorcroft, Strachey and other 19th century imperial agents were well-known to the frontier officers of the Raj by the early years of the 20th century. Thus when the Gartok Agency was opened those who controlled it were aware that Indian renunciates crossed into Tibet to worship at Kailas-Manasarovar. Yet this issue was not considered important enough for the right of those pilgrims to enter Tibet to be subject to any of the formal Anglo-Tibetan or Anglo-Chinese agreements concluded between 1890 and 1914. Britain and China agreed in 1908 that “natives of the Indian frontier … [were] at liberty to continue their trade, in accordance with the existing practice”, while the position of traders and their right of access to the Trade Agencies (or ‘marts’) in Tibet was covered in the 1893 Anglo-Chinese Trade Regulations and the 1914 Simla Convention. But the only agreement relating to pilgrims to Tibet came in the 1907 Anglo-Russian Convention, which was never officially shown to the Tibetans. Although Britain and Russia agreed that their (European) subjects would not visit Lhasa, Article Two of their agreement was designed to allow continued access to Tibet for Russian and British Indian Buddhists. It stated that Buddhists were allowed “into direct relations on strictly religious matters with the Dalai Lama and other representatives of Buddhism in Tibet.” No mention was made of the rights of Hindu pilgrims.106 Although several visiting British officers made the Kailas parikrama, British sources provide only a random series of vignettes concerning the pilgrim-

103 McKay (1997: esp., 158–165). 104 See note 116, chapter 12. 105 oioc, l/p&s/7/207–1873, report of W.S. Cassels, (A.C. Almora), 23 September 1907. 106 These treaties are most accessible in Richardson (1984). 406 chapter 15 age, for the colonial government and its officers had no particular interest in such matters. The pilgrimage served no strategic purpose, raised no inter- governmental issues, and was not a source of national revenue. Nor do the renunciates or the pilgrims seem to have availed themselves of the consular services of the Trade Agency or even visited it, for Gartok was not on the most common routes from India to Kailas. W.S. Cassels ics, visiting in 1907, was joined by three renunciates. He noted that they bathed in Manasarovar but only contemplated (darśan) the moun- tain. He reported there were around 150 Indian pilgrims a year, whom he describes as “fakirs” (i.e.: renunciates). But he noted one pilgrim, a tahsildar from Jaipur, who was also content to reach the plain of Gyanema (around Man- asarovar). The pilgrim told him that, “there were no religious devotees in Kailas learned in Hindu philosophy and religion and that nothing particular was to be gained by a visit to the place.”107 A 1926 report estimated the number of pilgrims from India at 100, a figure lower than that of 20 years earlier, although that was also an estimate. Interest- ingly a figure of 4–500 Tibetan pilgrims annually was given, a unique glimpse of its attraction (or lack thereof) in the traditional Tibetan world.108 But what appears to have been a crucial—if perhaps incidental—step towards encour- aging the pilgrimage came in 1928 at the recommendation of another visiting British official. Regulations concerning firearms licenses were eased for cross- border travellers, something which appears to have significantly affected the number of travellers to Kailas-Manasarovar.109 Banditry had always been a regional danger according to the accounts of travellers and pilgrims,110 and while traders were probably at greater risk, the prospect must have discouraged many householder pilgrims. Just as the greater security imposed by the British presence on the roads of India had encouraged trade and pilgrimage there, so too did it now encourage Indian pilgrims to Kailas-Manasarovar. In 1930 the Gartok Trade Agent reported that although the number of Tibetans (given as 600), had not notably increased, 730 pilgrims had visited from India (including 150 Ladakhis, presumably Buddhists).111

107 oioc, l/p&s/7/207–1873, report of W.S. Cassels, 23 September 1907. 108 oioc, l/p&s/12/4163–490, report of H. Ruttledge, (D.C. Almora), 10 November 1926. 109 oioc, l/p&s/12/4163, Government of India Foreign Department to Chief Secretary, Gov- ernment of the Punjab, 14 July 1931, approves Wakefield’s suggestions for easier gun li- cences for traders. 110 Swami Tapovan states that a robber he met in the 1920s informed him there were over 400 robbers on the route to Kailas at that time; Tapovan (2003: 231–232). 111 oioc, l/p&s/12/4163–5751, British Trade Agent Gartok to Superintendent Punjab Hill States, 1 September 1930. european construction 407

The number of Indian pilgrims subsequently remained around that num- ber, although 1931 was an anomaly as the Maharaja of Mysore visited with an entourage of around 700.112 The Trade Agency was transferred to the control of the new Indian government when the British left India, but was then taken over by China following the Sino-Indian accord of 1954. The last published fig- ures before the closing of the frontiers in 1959 are from 1951, when an estimated 600 pilgrims travelled from Almora to Kailas.113 By that time prominent renun- ciates were making proposals to the Tibetan authorities for the construction of rest houses (dharamsalas) at Manasarovar, but these came to nothing.114 With the Chinese take-over of Tibet in 1950, Indian pilgrims became closely moni- tored and the borders were sealed in 1959, closing off the pilgrimage until the 1980s.115 It is notable how few Westerners visited Gartok during the years that the Trade Agency existed there. Aside from occasional visiting officials on inspec- tion duty, the only prominent visitor was Professor Tucci. The British generally co-operated with the Tibetans in excluding all but British official visitors from Tibet. Tucci, however, befriended a number of influential British officials, not least the officer directly responsible for relations with Tibet, the Sikkim Political Officer (Sir) Basil Gould, who had brought in a wide range of political initia- tives designed to anchor Tibet more closely to India and to support the Tibetans against the Chinese. Gould noted that Tucci’s work “contributes in the direction of one of our main political aims, ie showing that Tibet is more closely allied to India than to China.”116 Tucci was thus allowed to travel in western Tibet and his pioneering scholarship was of great importance, not least in regard to the life of Rinchen Zangpo and the artistic heritage of the region.117 As a pioneer

112 oioc, l/p&s/12/4163–2202, British Trade Agent Gartok to Superintendent Punjab Hill States, 13 September 1931. 113 Bharati (1965: 96). 114 oioc, l/p&s/12/4163–7751, British Trade Agent Gartok to Superintendent Punjab Hill States, 9 September 1938, states that one Swami Nityananda-ji Saraswati, “promoter of Kailas-Manasarovar Yatra scheme at cost of around 10,000rps” requested permission to build 21 dharamsalas in Tibet. A letter formerly in the possession of the late Tashi Tobden ias (Gangtok), Swami Pranavananda to Rai Sahib Tobden Khazi, 18 June 1945, states that the local administrator had denied permission for a planned dharamsala, apparently a separate scheme. 115 The last published account from that era that I have located is from 1954. See Singh, (1955). 116 oioc, l/p&s/12/4247, Political Officer Sikkim to E.P. Donaldson, 6 April 1946. On Tucci in western Tibet, see Tucci and Ghersi (1935); also see Nalesini (2008: esp., 82–91); McKay (2015). 117 See Tucci (1988/1932). On Tucci see de Castro and Templeman (2015). 408 chapter 15 however, he relied on the records of the leading Gelukpa sect, and he was not concerned to critically examine the history of the pilgrimage, although he gath- ered manuscripts that would later assist in that process. Ultimately the presence of British representatives in western Tibet, and the imperial government’s acceptance that Indians needed to carry weapons in order to travel safely in western Tibet (a measure which required no govern- ment expenditure), created conditions which stimulated the growth of the pil- grimage. Once the site became accessible in relative safety, it became increas- ingly popular with a broad section of Hindu society, and from around 1930 onwards the pilgrimage became sufficiently popular to bring about the devel- opment of structures which mark a Hindu pilgrimage site. The modern under- standing of the popularity of the Kailas pilgrimage and of its being a journey undertaken by ‘ordinary’ or caste Hindus can only be dated from that period. A sacred space visited only by renunciates had been transformed by the impact of British imperial economic development aims into a Pan-Hindu pilgrimage site.

Conclusions

The fundamental contribution of early European accounts of Kailas-Manasa- rovar was the scientific mapping of the region and the establishment of its hydrological features, which as we have seen had significant implications for the sacred geography of both this region and the headwaters of the Ganges. The sacrality of the site was not the concern of those early reports. As late as 1905, even after the Younghusband mission, the Kailas-Manasarovar region and its sacred associations were known only to a small number of Europeans with a specialist interest in the region or in Indic textual sources. But within a decade following the publication of books by Sherring, Kawagu- chi and Hedin, the site had become widely known in Japan and the West as one of the most, if not the most, sacred places in the Asian religious world. In tak- ing disparate Indic and other Asian sources and creating a new framework in which to locate these earlier bodies of knowledge within Western understand- ing, these three authors took the authority to define the site from its holders— particularly the local people and the Kargyu and Śaivite renunciates—and relocated it in elite, or at least Western, hands. That they were able to do this was partly a result of the new print technologies, which made a book such as Sherring’s available not only all over India, but in Western libraries. Sherring’s regional claims thus became Pan-Indian and even world knowl- edge. european construction 409

We may also see this process in the context of state consolidation within modernity; the new construction of Kailas-Manasarovar became dominant knowledge because it was state-controlled. Other bodies of knowledge, wheth- er those of Jesuit missionaries or Śaivite renunciates, were side-lined as knowl- edge was concentrated in the hands of government and its supporting struc- tures. The new understanding of Kailas-Manasarovar and the process by which it arose was not, and has not been, subject to critical analysis, and so subsequent visitors to the region travelled with preconceptions deriving from the works of these writers. But under examination, it becomes clear that their construction did not represent a translation of Indo-Tibetan beliefs, or a balanced represen- tation, but was rather a selection of elements of existing understandings that had been reformulated to create a new understanding of the sacred site. That new understanding reflected and served the interests of political, commercial, and individual imperatives and the ambitions of its creators and their support- ing structures—government, publishers, and so on. We may suspect that this was the modern version of similar processes in the past. What was of particular significance was that, in the colonial context of its creation, this construction was accepted by the majority of Indians as repre- senting their beliefs. Indians, who had no reason to doubt authorities such as Sherring, henceforth engaged with the site partly or wholly in accord with the new construction. With the addition of the later findings and revelations of Indian authorities such as Swami Pranavananda, the ultimate result was the modern understanding of Kailas-Manasarovar, one that was very different, and far more cohesive, than that which had existed prior to the 1906–1912 period. chapter 16 From Theosophy’s Mahatmas to a Globalised Kailas

Introduction

Further testament to the comparatively recent origins of the modern under- standing of Kailas-Manasarovar is the lateness of its appearance in the West- ern esoteric imagination. Tibet itself first appeared in such constructions with Marco Polo’s accounts of its ‘necromancers’ and by the mid-19th century it had become a fully formed imaginal place in Western consciousness.1 But Kailas did not become a specific aspect of that construction until long after the 1903–1904 Younghusband mission had revealed the prosaic reality of the actual Lhasa. That encounter with reality forced the imaginal centre of Tibet to be relocated from there to its ‘hidden valleys’ and ‘remote caves and monasteries’, places which could be reconceptualised as sites conserving Tibet’s presumed esoteric aspects after these were found absent from the ‘Holy City’ of Lhasa. That relo- cation process involved a period (1904–1930s) in which Westerners referred to a number of sites in Tibet in terms that potentially situated them as supreme esoteric centres,2 before Kailas emerged as that centre in the popular imagina- tion. Although Sherring and Hedin established the understanding of Kailas as the supreme pilgrimage site for Hindus and Buddhists and romanticised its landscape, they did not locate it in any universalist framework. While acknowl- edging and advancing aesthetic and geographical causes for its sanctity they did not suggest Kailas was a spiritual goal, or otherwise a place of mystical or esoteric importance to anyone outside of those traditions. That it became understood in such terms was a later development; one not articulated in West- ern discourse until at least the 1930s, with its more extreme manifestations much later.

1 On which see Bishop (1989); also see Brauen (2004). 2 Shigatse, Shambala, and Amnye Machen were prominent candidates. The American anthro- pologist and adventurer Joseph Rock promoted the significance of Amnye Machen mountain; see for example, Rock (1930). Nicholas Roerich, probably unique among Westerners in sup- porting the sanctity of the Panchen Lama above that of the Dalai Lama, advanced interest in Shigatse rather than Lhasa, as well as in the mythical land of Shambala; see Roerich (1996). On the attempts of an American Theosophist to reach his “Master” in Shigatse in 1926, see McKay (1997: 95).

© koninklijke brill nv, leiden, 2015 | doi: 10.1163/9789004306189_019 a globalised kailas 411

The roots of this esoteric construction lay in the Theosophist movement and in the broader sense arose within the late 19th century and early 20th century European (and Russian) interest in such matters as Egyptology, Celtic mythol- ogy, Freemasonary, and aspects of Spiritualism, notably communication with the spirit world. These were the imaginal framework shaping the constructions of the creators of the universal Kailas. As was the case with the cosmological construction of Kailas in the Indic tra- dition, this Western-imagined Kailas allowed a dynamic process of attachment of all manner of imaginings to the central symbol. The result was that there are today a loosely-linked series of esoteric associations with Kailas-Manasarovar that collectively form a powerful image of a spiritual centre. This image is par- ticularly prominent within the myriad spiritualities of the New Age movement as it emerged in the 1960s, drawing on an eclectic selection of Asian, among many other, religious beliefs, symbols and traditions. This image of a spiritual centre is also one that features strongly in the modern promotion of the site in popular media and in travel brochures, literature and websites.

The Modern Esoteric Kailas

We can define the core of a modern esoteric understanding of Kailas as embrac- ing a conception of the region as a location for some form of Gnostic Truth, some ultimate or nameless wisdom beyond established faiths. This spiritual, or other-worldly aspect—for it is often associated with alien beings—is com- monly understood as including knowledge preserved for thousands of years by a handful of spiritual masters whose lineage may be indigenous to the region but is frequently traced back to such real or imagined realms as Mesopotamia, Pharaonic Eygpt, or Atlantis. As a casual trawl of the internet will show, at some point in this discourse, rational possibilities give way to improbabilities and eventually to total fantasy. Thus we read of the pseudo-scientific discovery that;

Mt. Kailas could be a vast, human-built pyramid, the centre of an entire complex of smaller pyramids, a hundred in total. This complex, moreover, might be the centre of a worldwide system connecting other monuments or sites where paranormal phenomena have been observed.3

3 “Russians Reveal Kailash Pyramid Mystery”; Evelina Rioukina, unece Vedic Empire http:// vedicempire.com; accessed 10 December, 2010. Emphasis added. 412 chapter 16

Then there is Kailas as a point of access to a subterranean world inhabited by an advanced race who follow the “Melchizedek teachings”, whose power is “generated from the originating or Supreme Center, we call Shamballa.” This “Agartha Network” of vast cities includes an Atlantean outpost and is repre- sented on earth by the Dalai Lama, whose messages are transmitted through “certain secret tunnels connecting the inner world of Agartha with Tibet.”4 The idea of Kailas as allowing access to subterranean cities is a powerful motif, one might even be guided by an on-line shaman on a visionary journey through such a place, a voyage;

into Mt. Kailash’s Subterranean Crystal City … … of many inhabitants, great structures of crystal, crystals that are gathered and carved out as transponders reaching out into the universes, other dimensions, the many realms unseen presently by our surface eyes … and by that be nurtured into a state of the unity of self fulfilling immortality … with the nurturing of the highest will in positive light empowerment.5

These imaginary constructions may embrace Christian aspects notably via links to the disappearance in the Kailas region of the Indian convert Sadhu Sundar Singh.6 Or, with the characteristic ecumenicalism of the New Age, Kailas may become a site where religious distinctions merge, a place where the fused souls of Adam and of Śiva live under the ice-cap.7

4 Albero Cogliani (Nityanand), “Shamballa, Agartha and the Galactic Federation of Light”, 26 August 2008; http://forums.canadiancontent.net/alternate-theories/76470-galactic -federation-light-have-you.html; accessed 11 December 2010. 5 “The Line into Mt. Kailash’s Subterranean Crystal City, Inside the Crystalline Grid Where True Life Begins”; http://www.davidicke.com/forum/showthread.php?t=121210&page=2; accessed 10 December 2010. Interestingly, there is Hindu textual sanction for the general concept; Vayu Purana 41.55 (Tagare, 1987: 272), states that, “there is an inaccessible city within the (belly of) the mountain, with only a cave as the main entrance.” Also see, for “golden underground cities wherein live Nagas”; Varaha Purana 81.2. 6 Sadhu Sundar Singh was a Sikh convert to Christianity who maintained the lifestyle of a renunciate. He claimed to have met a 300 year old Eygptian Christian, a member of a secret society going back to the time of Christ, when he visited Kailas-Manasarovar in 1913. In 1929 he set out for Kailas again via the Niti-la and was never heard of again; see, for example, Thompson (1997). 7 http://www.abovetopsecret.com/forum/thread85245/pg1, accessed 10 December 2010. The original reads “where the soul of adam and shiva came to together, which it now resides deep within mount Kailas under the thick ice on the top. [sic]” a globalised kailas 413

Yet other imaginings suggest some prior engagement with Asian history or traditions in that they reflect certain indigenous ideas and even current move- ments and debates. The Order of Nazorean Essenes,8 for example, hold as an article of faith that the true Buddhist teachings are those of an earlier Buddha named Shenrab and “the system of Dzogchen Enlightenment once taught by Him and His female counterparts on Mt. Kailash.” They claim that this form of “proto-Indian European Religion” emerged around Kailas “in the ancient Zhang Zhung kingdom [which] taught the ancient gnosis under the name of Yungdrung Bon.” As they note in an explicit link to the modern Bön ‘nativist’ movement, “many remnants of this once thriving faith still exist among the modern Bonpos.” Similarly we read reflections of local traditions unadorned by Hinduism, references to the;

Nagas, the ‘serpents’ which live in extensive underground palaces in the rocky Himalayas … [and] are able to fly in space … possess amazing magi- cal powers and intelligence … The deep caverns of the Nagas contain fabu- lous treasures, illuminated by flashing precious stones. The subterranean abodes are known to be in certain parts of both the Himalayas and Tibet, particularly around the Lake of the Great Nagas—Lake Manosarowar.9

The Theosophical Impulse

It is significant that there were no Western constructions of this kind associ- ated with Kailas until well into the 20th century, at least a generation after the Sherring-Hedin construction had passed into general knowledge of the region. In 1917 for example, a young American traveller and spiritual seeker Edwin Schary, travelled to western Tibet in search of spiritual masters (‘Mahat- mas’) that he believed could be found there. He was the first known Westerner whose journey in that region was motivated by non-Christian spirituality. Yet although he encountered Indian and Tibetan pilgrims as he passed by Kailas- Manasarovar and noted their beliefs in an account of his voyage (published only in 1937), the site had no particular resonance with his understandings of the sacred Himalayas and their concealed spiritual wisdom. Schary describes

8 http://www.essenes.net/index.php?option=com_content&task=view&id=14&Itemid=311; accessed December 2010. 9 Andrew Tomas, “On The Shores Of Endless Worlds” (Souvenir Press), 160, cited on http://www .abovetopsecret.com/forum/thread74894/pg1; accessed 10 December 2010. 414 chapter 16 how he had dreamt of a specific cave which became the object of his journey, and how he actually found the cave he had dreamt of. But neither in his dream nor in reality was that cave located at Kailas (it was more than a week’s jour- ney westwards), and in any case there was no Mahatma there and its discovery vouchsafed him no great spiritual enlightenment.10 As Schary’s search for Mahatmas indicates, he was familiar with the ideas of the Theosophists.11 These included links to ancient creeds, other-worldly dimensions, an ideal of universal brotherhood and most importantly in this context, divine revelations from Mahatmas or ‘Masters’ resident in the Hima- layas. The Theosophists’ blend of esoteric teachings from numerous (real or imagined) spiritual traditions was the most significant precursor of the New Age fusion of Asian and Western esotericism. They provided many of its main themes, not least this concept of a secret order preserving timeless spiritual wisdom which might be accessed by the seeker.12 The Theosophical Society had been founded in 1875 by Helena Petrovna Blavatsky (1831–1891), a widely travelled Russian émigré and spiritual teacher whose influences included Russian Orthodox and Coptic Christianity, Bud- dhism and the religion of the Druze, Sufism, Kabbalah, Masonic, Rosicrucian, and many other strands of belief. Blavatsky spent the years 1879–1885 in India where her society established a lasting presence. Particularly in the decades before the First World War, the ideas of the Theosophists had considerable impact on the small but influential group of Europeans and Indians at the inter- face of cross-cultural exchange in colonial India and were even allowed some credence by the more philosophically inclined colonial officials despite the pre- dominant rationalist and scientific tendencies of the period. A skilled linguist and charismatic charlatan, Blavatsky characteristically cul- tivated a wide range of influential contacts. In Europe for example, the explorer Sir Richard Burton and the Italian revolutionary Giuseppi Mazzini were among her followers or acquaintances,13 and in India she continued to associate with spiritual and political leaders. She firstly came under the influence of the founder of the Arya Samaj, Vedic scholar and social reformer Swami Dayananda Sarasvati, a key figure in the modern Hindu nationalist narrative (as noted in Chapter 10). The Theosophical Society’s connection to the emerging Indian

10 Schary (2005: esp., 196–207). 11 Schary (2005: 21 & see 37–42 on his stay in Adyar and attempts to meet Annie Besant). 12 The Mahatmas were not entirely a Western invention; they fuse ‘Wise Man from the East’ mythology with aspects of the ideal renunciate, the 84 Mahasiddhas, and the rsis. Also see note 47. 13 Johnson (1994: 38–43, 65–66). a globalised kailas 415 nationalist movement became explicit with Blavatsky’s successor as head of the Theosophists in India, Annie Besant, becoming a founding member of the Indian Congress party.14 Theosophy was thus a major spiritual and intellectual movement in India, and informed the ideas of both Western and Indian edu- cated classes there. While Blavatsky’s claim to have visited Tibet was almost certainly fabricated, the Theosophists did establish connections with the Sikkim-based British agents Sarat Chandra Das (1849–1917) and Urgyen Gyatso (1851–1915), who had travelled in Tibet in 1881–1882. From them Blavatsky apparently obtained Tibetan texts from Chandra Das’s main Tibetan contact, the Sengchen Lama, who was the Panchen Lama’s Chief Minister in Shigatse. Chandra Das was the leading British Indian expert on Tibet in the pre-Younghusband era, and these men were all key figures in events surrounding late 19th century British approaches to Tibet.15 They were well-qualified to act as cultural intermediaries and from them Blavatsky seems to have gained some knowledge of Tibetan beliefs, but her Theosophical teachings were little affected by the realities of Tibetan (or Indian) practices. The predominant Tibetan mythological construct that was used in Blavat- sky’s work was the ‘Hidden Kingdom’ myth of Shambala.16 Consistent with the conclusion that Kailas was of little renown in the 1880s, it largely escaped her attention and she mentions the region only in passing, as the home of Śiva for example, or in regard to a possible sighting there of her supposed Master “Koothumi.”17 It was, however, Blavatsky’s claim that there were Mahatmas in the Himalayas who communicated divine wisdom to her that was to be her major legacy in the esoteric construction of Kailas. Just as earlier layers of belief survive within Asian traditions, so this belief in Himalayan Masters has con- tinued into the modern era,18 inspiring others to seek out those Mahatmas.

14 Besant also made a translation of the Bhagavad Gita (an important text in the nationalist context and in regard to the construction of modern Hinduism), which version was favoured by Gandhi; Paine (1999: 79). 15 See Johnson (1994: esp., 191–206); also see, McKay (2002); Pedersen (2001); Kaschewsky (2001); Korom (2001); Lopez (2001). 16 On Shambala, see for example, Bernbaum (1980). 17 Blavatsky (1883: 98–99). I have not read Blavatsky’s complete works, but a brief perusal and the secondary sources give no indication that she considered the site being of any real significance. 18 The Chilean fascist diplomat Miguel Serrano, for example, claimed membership of an order “rumoured to have come from India, or perhaps from Tibet”, which was run by a Himalayan Master who informed Serrano that his order had its headquarters in a cave 416 chapter 16

Blavatsky seemed to identify them as Indian rather than Tibetan and as res- ident in Shigatse rather than western Tibet, but later seekers imagined the Mahatmas manifesting in a variety of forms and as residing specifically in the vicinity of Kailas.

Mystics and Mythologisers

The French explorer Alexandra David-Neel (1868–1969), who travelled through Tibet in the 1920s, did not articulate an esoteric Kailas. David-Neel was a Theosophist, attending Blavatsky’s London salon in the late 1880s and rent- ing an apartment in their Benares complex as late as 1914.19 But as a serious student of Tibetan mysticism she seems to have maintained some intellectual distance from the movement’s wider shores. Rather than seeking Blavatsky’s Mahatmas, her enquiries focussed on traditional figures among the indige- nous authorities of the Buddhist Himalayas. She had however, met Kawaguchi Ekai,20 and during her years in Sikkim she had as her informant and transla- tor Kazi Dawa-Samdup (1868–1922), Sir John Woodroffe’s Tibetan translator. David-Neel described the Kazi as “an occultist and even, in a certain way, a mystic”,21 and had Kailas then enjoyed any particular renown in Indo-Tibetan mysticism it is hard to imagine that she would not have referred to it in her writings, which included numerous accounts of esoteric practices. Kailas is similarly insignificant in the works of the Russian artist and mystic Nicholai Roerich (1874–1947), who travelled across the Tibetan plateau in the late 1920s. Like Blavatsky, Roerich’s esoteric focus was with Shambala rather than Kailas, though he apparently concluded that the former lay somewhere

at Kailas. Serrano had met Roerich, who claimed Blavatsky had been to Tibet and had attended one of the four secret ashrams—at Shigatse, Everest, Kailas, and Tholing— ashrams Roerich said were mentioned in (unspecified) Tibetan texts (none of which seem to have come to light); see Serrano, (1963: 44–47, 187–188). 19 David-Neel (1971: 49). 20 http://www.alexandra-david-neel.org/anglais/biog5.htm; accessed 31 October 2010. 21 David-Neel (1971: 20–21): Kazi Dawa Samdup was born in Kalimpong and educated at the Bhutia Boarding School in Darjeeling, where he studied Tibetan under Lama Urgyen Gyatso. He became headmaster of the (other) Bhotia Boarding School in Gangtok, and served as a translator for the Chogyal (Maharaja) of Sikkim, for the 13th Dalai Lama during his 1910–1912 exile in India, and at the Simla Convention in 1913–1914. Appointed Tibetan Lecturer at Calcutta University where he died in 1922; see Samdup (2008: 155– 158). a globalised kailas 417 to the north of that mountain.22 Roerich’s well-known painting entitled Path to Kailas dates to 1933, by which time he was living in Kinnaur (near to the Kinnaur Kailas), and that much reproduced work was his only contribution to a mystical image of a Kailas mountain. This painting dates however, to the era in which Kailas began to take on meaning in Western esoteric thought. The Nobel prize-winning Irish poet Wil- liam Butler Yeats (1865–1939), was a significant figure in this process. He had a life-long interest in both Hinduism and esoteric spiritualism, joining the Theosophical society in London in 1888 and the Rosicrucian Order of the Golden Dawn two years later.23 Around 1930, Yeats met Sri Purohit Swami (1882–c. 1936), a Maharashtran Brahman renunciate who had come to London. Yeats helped the Swami publish, and wrote introductions to, several works on Hindu culture. These included The Holy Mountain (1934), an account by Puro- hit Swami’s guru Bhagwan Sri Hamsa, of his 1908 journey to Kailas-Manasaro- var. The Holy Mountain was the first English-language book to portray Kailas with universal esoteric associations. While predominantly a prosaic account of a Hindu pilgrimage, authenticated not only by mention of various local dignitaries but by a quotation from Hedin whose visit had been only a month earlier, Hamsa’s story claimed one, if not two, mystical experiences (neither actually inexplicable in scientific terms). Firstly he encountered a learned “Hindu Mahatma” living in a cave, “who told me that he knew all the languages of the world.” Hamsa then received a mystical vision of his tutelary deity (Dattatreya24). The multi-lingual Mahatma clearly appealed to followers of the Theosophical vision, but perhaps most significantly Hamsa stressed, in the

22 The Russian writer Andrew Tomas (1974) locates Shambhala between 45 and 50 degrees north latitude, possibly somewhere in the Karakorum range northwest of Mt. Kailas. He states that “Roerich told him he had found the Holy Place northwest of Mt. Kailas, although his discovery is not unambiguously corroborated in Roerich’s own writings.”; http://www .victoria-lepage.org/shambhala_spiritual_axis.htm; accessed 11 December 2010. 23 Jeffares (1984: n. 13 6). 24 Sri Hamsa, who was Marathi, claimed to have been initiated into the Giris at Kailas by Śiva himself in the form of Dattātreya. Dattatreya was “the semidivine founder and leader” of the Nine Nāths of western India, and also the popular toponym of a group of mountain peaks there; (White 1997: 88–89). The cult arose in Maharashtra in the 13–14th centuries and Dattatreya replaced Bhairava as the tutelary deity of the Dasnami order; (Clarke 2006: 67, 234–235). There is a Dattatreya gotra in Kashmir (Sanderson 2003–2004: ns37 & 39 363–364); and there is also a Dattatreya rsi, who is referred to by Atkinson as an incarnation of Visnu and credited with giving the Himachal mahatmyas to the Kashi raja: (Atkinson 1884: 40). 418 chapter 16 context of the joy of sighting Manasarovar, that his emotional experience was not limited by religion. “I am sure [he wrote], that the sight of this magnificent Lake Mānas will arouse similar feelings in everyone, whoever he may be, Hindu or what you will.”25 This theme of universality was soon to be emphasised in regard to the spiritual, rather than the emotional encounter. Hamsa’s publishers, Faber and Faber, obviously saw commercial potential in this work and the frequency with which The Holy Mountain has been reprinted since 1934 demonstrates the correctness of their judgement.26 But in endorsing this account of Kailas, Yeats (and by implication Hamsa and Purohit Swami), was articulating what was then a novel or fringe belief. In his report on the book and Yeats’s introduction, one of Faber and Faber’s directors, the famed poet T.S. Eliot, wrote that;

Uncle William [W.B. Yeats] is as loony as ever … it’s a good long introduc- tion and I suppose it would help the sale of the book. I don’t mind it [the introduction] being mostly nonsense …27

It was in 1933–1934, while preparing to write that introduction, that Yeats wrote his renowned poem Meru,28 the mountain which, as he makes clear in The Holy Mountain, he understood as identical to Mount Kailas. Yeats was familiar with classical Indian literature. He had read translations of Kalidasa, Patanjali, and Upanisadic and Epic texts, but his framework of understanding of the site apparently derived from reading Hedin and Kawaguchi’s accounts of Kailas-Manasarovar. Despite Yeats’s beliefs in fairies and suchlike there is no hint of the esoteric in this verse, only the poetic use of Hedin’s construction. While Everest sits oddly in this context (although the reports of the early 1920s climbers there had mentioned hermits), the author’s renown made famous the lines:

Hermits upon Mount Meru or Everest, Caverned in night under the drifted snow, Or where that snow and winter’s dreadful blast Beat down upon their naked bodies, know

25 Hamsa (1998: 139, 144–145, 165–172, 178–182). 26 The Holy Mountain ran to four editions in 1934 and has been reprinted on a further three occasions since 1984. 27 Matthew Evans, in The Guardian, Saturday 6 June 2009, available at http://www.guardian .co.uk/books/2009/jun/06/t-s-eliot-faber-matthew-evans; accessed 16 December 2010. 28 Jeffares (1984: 333). a globalised kailas 419

That day brings round the night, that before dawn His glory and his monuments are gone.29

Also behind the construction of a New Age Kailas was a student of Yeats, the anthropologist Dr. W.Y. (Walter) Evans-Wentz (1878–1965). A Theosophist from his youth, Evans-Wentz studied at Stanford where he was inspired by Yeats, and later at Oxford (where he focussed on Celtic mythology), although his doctorate was from the University of Rennes in Brittany.30 By 1919, via research in Egypt on its ancient mythology, Evans-Wentz reached India where he became a yoga practitioner, if not a Hindu convert. He then obtained and had translated and published, four significant Tibetan texts, most famously The Tibetan Book of the Dead (1927, with a foreword by Sir John Woodroffe), and Tibet’s Great Yogi Milarepa. Although these were published under his name, Evans-Wentz did not read Tibetan. He hired a Tibetan translator who, in a sign of the small cast of relevant figures here, was Kazi Dawa-Samdup (who died before these works were published). Evans-Wentz had a particular interest in sacred mountains and was certainly aware of Kailas-Manasarovar. In 1931 he sought permission from the British Government of India to cross the frontier and visit Kailas, but was refused— probably his contacts with Indian nationalists were held against him.31 Evans- Wentz bought a property in Almora intending to establish a spiritual cen- tre there, but returned permanently to America on the outbreak of war in 1939. There he inherited property that included Mount Tecate/Cuchama, a site sacred to local indigenous tribes which he eventually willed to the nation. His final work entitled Cuchama and Sacred Mountains was posthumously pub- lished in 1981 and includes several pages concerning Mount Kailas/Meru, which he calls “the holiest spot on the planet Earth.”32 That globalising claim and the description of Kailas is drawn, however, from the seminal work of Anagorika Govinda, discussed below.

29 Janet Zimmerman Marsh, “The influence of Hinduism in William Butler Yeats’s ‘Meru’”, Yeats Eliot Review, Winter 2005; FindArticles.com. accessed, 02 Feb 2010. 30 Evans-Wentz (1989: xix); also see, http://en.wikipedia.org/wiki/Walter_Evans-Wentz. 31 Oriental and India Office Collection, l/p&s/12/4289–256; National Archives of India, For- eign and Political Department, 1919, General b, May 19; as early as a report of 1 May 1918 by C.R. Cleveland, Director of Criminal Intelligence, Evans-Wentz was considered “suspi- cious” and a “scallywag”, despite the support of the us Consul. Also see l/p&s/12/4287, various correspondence. 32 Evans-Wentz (1989: 55). 420 chapter 16

Evans-Wentz’s Theosophical lens problematises the works he published. Donald Lopez has drawn attention to this in the context of Evan-Wentz’s inter- pretation of The Tibetan Book of the Dead;

[T]he book is read as a code (a system of constraints) to be deciphered against another text that is somehow more authentic … for Evans-Wentz, the urtext is Madame Blavatsky’s The Secret Doctrine …33

While Evan-Wentz’s Theosophical beliefs shaped his work, they had less affect on his friend and protege, Lama Anagorika Govinda (1898–1985). Govinda was German, born Ernst Lothar Hoffmann and educated in Switzerland and Italy.34 After military service in the First World War and a subsequent bohemian existence in the 1920s artist colonies of Capri, his interest in Asian philosophies led him to move there in 1928. Govinda was firstly ordained as a Theravadin Buddhist monk in Ceylon (Sri Lanka) and in 1931 found a Tibetan Buddhist guru in Sikkim and began his studies of that tradition. For Govinda, Tibet was, “the last living link that connects us with the civilizations of a distant past.”35 Despite that view and although he contributed articles to the Theosophical Journal, Govinda does not seem to have considered himself a Theosophist. He was a gifted and highly quotable author who wrote primarily for a Western audience from the assumed perspective of a Buddhist insider. He reportedly read, though he did not speak, Pali, Tibetan, and perhaps Sanskrit,36 yet as Lopez has noted he relied primarily on Western sources and Buddhist works in translation.37 As a result, despite his assumed authority, his work represents a Western interpretation of Buddhism, at best a cultural fusion rather than an actual Himalayan Buddhist perspective. In the mid-1930s, Evans-Wentz met Govinda, whom he encouraged and promoted. On his departure for America in 1939 he installed the German in his Almora property. There Govinda, a prolific writer and lecturer with an

33 Lopez (1999: 85; also see, 47–67). Lopez records that although Evans-Wentz later claimed Dawa Samdup to have been his “guru”, their meetings were brief and formal. 34 Though frequently described as German-Bolivian, Govinda’s connections to Bolivia were through his maternal grandfather, a soldier of fortune who served as a General under Simon Bolivar; Govinda himself does not appear to have been there. For a biography of Govinda see, Winkler (1990). 35 Winkler (1990: 126). 36 Winkler (1990: 125); Sangarakshita (1991: 453); Govinda’s (anonymous) editor states that he was, “well versed in Pali, Sanskrit and Tibetan”; Govinda (1991: xiii). 37 Lopez (1999: 61–62, 125). a globalised kailas 421 appealingly sage-like appearance and personality, became a central figure in the Western expatriate spiritualist community down to the 1970s, providing a link between generations of Western spiritual seekers. He knew David-Neel, whose works he had read, and Roerich, who arranged for a wing in the Alla- habad museum where Govinda’s art went on permanent display.38 But he also hosted ‘Beat’ poets Gary Snyder and Alan Ginsburg,39 along with growing num- bers of Western spiritual seekers who took the overland trail to India in the 1960s. To a large extent Govinda’s reputation derived from his journey to the Kailas region in 1948. He recounted this in several chapters of his best-selling The Way of the White Clouds (an extract from which was published as a foreword to Evans-Wentz’s Cuchama and Sacred Mountains). Govinda’s work, part spiritual autobiography, part Tibetan Buddhist primer, has been reprinted more than 50 times and he also described Kailas in a more mystical wider context in later writings.40 Adding to the impact of his work was that of his wife, Li Gotama, a Bombay Parsi photographer and artist who had studied under Rabindranath Tagore. She accompanied Govinda on his journey to Kailas and published a series of articles on the journey in the Illustrated Weekly of India in 1951. Several of her paintings went on permanent display in the Prince of Wales Museum in Bombay,41 drawing further attention to the Kailas complex. Tibet was closed to Europeans after the Chinese invasion in 1950 and Prana- vananda’s guide-book was rare and almost unknown in the West. Way of the White Clouds thus became the major source on Kailas after its timely publica- tion in 1966, which coincided with the emergence of a New Age synthesis from the spiritual experiments of the ’60s counter-culture. That the site was inac- cessible to outsiders for a generation before and after the work appeared could only add to the mountain’s appeal,42 just as the closing of Tibet’s frontiers had added to the attraction of that land to 19th century Western explorers. Kailas partook of the lure of the inaccessible and the forbidden.

38 Winkler (1990: 50, 129); Sangarakshita (1991: 274); Govinda (1984: 81, 85). 39 Snyder (2007: 97). 40 See for example, Govinda (1991: 3–12). 41 Winkler (1990: esp., 111). 42 See for example, Thurman and Wise (2000: 6) where Thurman describes how the “inac- cessibility of Mt. Kailash gradually came to represent to me the suppression of the power of human goodness that we need to make our dreams of peace come true … Trapped as it was behind communist lines, it seemed to me that the mountain itself needed to be liberated, restored to its spiritual purpose.” 422 chapter 16

As in Hamsa’s work, there is actually much in Way of the White Clouds that is empirical43 and that would not be out of place in the works of Sherring or Hedin. Govinda acknowledges the region’s aesthetic appeal and geograph- ical significance, but he also incorporates an esoteric world removed from the scientific materialism and colonial context of those writers. In that the work includes extensive accounts of such phenomena as oracles, strange lights, apparent miracles, and wise incarnations in the Himalayas, it merely comple- ments the existing imaginal construction of Tibet. But in locating his Kailas pilgrimage as the culmination of the narrative, he focussed that imaginal con- struction on the Kailas-Manasarovar complex.44 Furthermore, developing a possibility touched on by Hamsa and Prana- vananda, Govinda specifically identified Kailas-Manasarovar as a universal sacred site. Although Govinda considered himself a Buddhist, he came to reject sectarian and cultural boundaries on the nature of ultimate truth and in this work he implicitly relegated the indigenously worshipped Kailas deities to purely local manifestations of a wider, higher, global spiritual force (just as Hinduism and Buddhism had overwritten local beliefs). Turning away from indigenous sources, Tibetan or Indic, he subsumed distinct bodies of indige- nous knowledge beneath the umbrella of the wider spirituality he imagined, implicitly reinforcing the Western Orientalist appropriation of authority over local Asian knowledge. While the belief in the Mahatmas of the Himalayas was too appealing a myth to vanquish from popular culture, Govinda transformed this concept, articulating an understanding of the mountain itself as the Mahatma;

There are many religious orders in this world … with rules and regulations, with dogmas and rituals, with vows and initiations. But the brotherhood of those who have performed the pilgrimage to Kailas, who have gone through the trials of dangers and hardships, and were rewarded with the glorious vision of the sacred land, has received an initiation of the most profound nature …45

No longer was it necessary to study Asian languages or religions to partake of Kailas’s spiritual power; the act of visiting the site was now sufficient to

43 This author used Govinda’s work as a guidebook on the Kailas parikrama in 1986. 44 The Illustrated Weekly of India funded Govinda’s journey as an exploration of the temples at Tsaparang. 45 Govinda (1984: 208). a globalised kailas 423 initiate the pilgrim into a Gnostic world above and beyond local Asian tradi- tions. That understanding ‘opened’ the site to all who would—political realities permitting—venture there. Kailas was now globalised as a sacred centre. In his later writings, Govinda went further, specifically claiming that Kailas manifested a religion that was “truly universal” and inescapable in its power. Kailas provided;

The experience of a higher reality which is conveyed through a strange combination of natural and spiritual phenomena, which even those who are unaffected by religious beliefs cannot escape … [Kailas was] the seat and center of cosmic powers, the axis which connects the earth with the universe, the super-antenna for the influx and outflow of the spiritual energies of our planet.46

Govinda thus constructed the idea of Kailas-Manasarovar as a universal spiri- tual centre. No longer was its sanctity restricted to followers of Asian religions, it was appropriated by an emerging culture, revealed as a pilgrimage place for all spiritual seekers. Like Sherring and Hedin, Govinda had taken elements from earlier traditions, reinterpreted them, and constructed a new understanding. He provided the framework for a globalised image; a new Kailas for a New Age. Once the framework for a New Age understanding of Kailas was established by Govinda, visionaries and all manner of fringe believers and fantasists could use that framework to paint in their own imaginings. Kailas was not removed from Buddhist or Hindu meaning, but it was added to the list of terms symbolic of esoteric mysteries, terms such as Atlantis/Lourdes/Pyramids/Dalai Lama, and so on, that are almost interchangeable in the more extreme esoteric construc- tions of the New Age.

Agents and Agencies in the Universalist Kailas Construction

The modern esoteric image of Kailas-Manasarovar first emerged in the 1930s. In earlier accounts of esoteric traditions in Tibet by such seminal figures as Blavatsky, David-Neel, and Roerich, Kailas was hardly mentioned. But these

46 Govinda (1984: 6–7); emphasis added. Also see Thurman and Wise (2000: 5), where Thurman writes that; “Lama Govinda told us that he had discovered an ultimate place of power on the north face of the mountain where one could plant one’s deepest wish for the world, and all the Buddhas and gods and dakini-angels would see to its accomplishment. Mt. Kailas could be that powerful.” 424 chapter 16 pioneers of the exploration of the esoteric traditions of Tibet were all to some extent or other involved with Theosophy and the earliest imaginings of the site are associated with the Theosophist’s idea of a secret society of Himalayan Mahatmas.47 By the late 1920s, there was a small group of seekers such as David-Neel, Roerich, and Evans-Wentz, who had developed a wide knowledge of Tibetan and Indian beliefs however filtered through Western and Theosophical lens that knowledge may have been. In addition to their shared intellectual back- ground they were contemporaries who knew each other personally or read each other’s works with approval. Collectively they formed a mutually rein- forcing body of knowledge and we may assume that they discussed sacred sites between themselves. Yet Kailas-Manasarovar apparently had no partic- ular imaginal significance for them until Evans-Wentz and Sri Hamsa’s works began to transform the construction of Sherring and Hedin by illuminating the spiritual possibilities of that region. Once that construction took shape, the research sensibilities of this group were more informed by Theosophy (and later in Govinda’s case by his identifi- cation as a Buddhist), than by critical source analysis and the established tropes of anthropological fieldwork.48 Ultimately, a lack of critical thought charac- terised both their belief in Theosophy and their approach to Asian esoteric traditions and understandings. Privileging European ideas and philosophies over local sources, they appropriated and reconceptualised those sources in line with their own intellectual traditions. None therefore, ever interrogated the myth of Kailas. This was never entirely a Western construct. There was significant input from Sri Hamsa. His work was published with the support of W.B. Yeats, the distinguished literary figure whose introduction occupied nearly a quarter of the book. Yeat’s patronage ensured that Hamsa’s book joined the list of those whose works were well known, but whose conclusions were uncontested.49

47 While the concept of Himalayan Mahatmas was unknown in Indo-Tibetan tradition, and initially bemused its people, the concept was not too far removed from that of a respected renouncer from a high social background. It was also a positive image and one modern Himalayan renouncers and spiritual teachers might not be adverse to drawing on. Certainly the esoteric construction of Kailas has fostered the interests of both the peoples of that region and the Bön Zhang-zhung ‘nativists’. 48 Govinda for example, was unaware that Professor Tucci had visited Tsaparang (and Man- asarovar) a decade before him; Sangharakshita (1991: 457–458). 49 Tucci’s full account of the site was only available in Italian until the 1980s, and Tucci accepted the Sherring-Hedin construction uncritically. a globalised kailas 425

Hamsa’s Holy Mountain is, as noted, actually prosaic in tone, concise and largely empirical, with an accessible European mountain landscape aesthetic. Its Biblical language surely reflects the hand of a Western editor—perhaps Yeats himself. On the two occasions the author strays into the realms of mys- tical experience the first simply repeats an unlikely claim by a renouncer, the second, a religious experience, is not inexplicable in scientific terms (visions induced by altitude, exhaustion, and faith). The account was thus at least plausible to the critical reader, while fulfilling Western fantasies of wise old Mahatmas and revealing where they could be found. Hamsa’s work was thus transitional, a modernist text which acted as a bridge between the scientific constructions of Sherring, Hedin, and Pranavananda, and the emerging eso- teric construction of Kailas. Until the 1930s, and for some time after, Kailas was largely obscured beneath imaginings around a competing myth, that of the Hidden Land of Shambala. That culminated in James Hilton’s fictional Shangri-La, as portrayed in his best-selling Lost Horizon (1933). But however attractive that mythology, the fictional place could not long compete with a “cosmic pillar that upholds the vault of heaven”,50 when it became clear that that pillar could actually be visited. Here it was the work of Anagorika Govinda that was most significant. He had been to the sacred site and his authority transformed the mountain itself into a Mahatma; simply being in the presence of this mountain could not, he claimed, fail to enlighten even the non-believer. Uniquely placed as a bridge between generations and the alternative spiritual traditions of Theosophy and the New Age, Govinda, like Hamsa before him, presented a Kailas that was not incompatible with the ‘scientific’ construction of Sherring and Hedin, but which directed its appeal to all spiritual seekers rather than empiricists and followers of Indo-Tibetan religions. Despite his claim (no doubt genuinely felt), to being a Buddhist, it was primarily Govinda who gave a structure to an eclectic series of imaginings around the sacred mountain, removing the site from its Indo-Tibetan cultural context and transforming it into a globalised mountain for a globalising world. While there are internet fantasies of Kailas that are entirely a Western con- struct, there is a complex relationship between Eastern and Western agency in the mythological construct of Kailas. Some of the processes that were involved in the construction replicated those that occurred in Indo-Tibetan history. Hin- duism and Buddhism had long since appropriated local traditions and under-

50 Thurman and Wise (2000: 3). 426 chapter 16 standings of the site. Govinda’s universalising construct was simply a continu- ation of that process, subsuming them in turn under a universal religion. Sim- ilarly, this Kailas became a symbol used in many different claims to authority, appearing in different forms in distinct bodies of knowledge, just as had been the case in its Indo-Tibetan employment, albeit that modern technology made them more widely available. It is notable that until at least the 1970s there was little or no input into the Kailas construction from Tibetan representations. Although the esoteric inter- ests of the Sikkimese, Kazi Dawa-Samdup, apparently coincided with those of his patrons (David-Neel, Evans-Wentz, and Sir John Woodroffe), it is also diffi- cult to detect his having any input into the image. Dawa-Samdup’s translations of texts associated with the Cakrasamvara Tantra51 and the life of Milarepa might suggest his understanding of Kailas was from within the mainstream Tibetan monastic Buddhist tradition, but he abdicated the role of translating that into Western terms to his patrons. Tibetan voices began to emerge into the Western mystic Kailas through the filter of Anagorika Govinda’s vision, which was ultimately universal and Bud- dhist only in the Western sense. His work did however, coalesce with the greatly increased Western interest and participation in forms of Buddhism in the 1950s and ’60s. But by the time the site opened to Western travellers in the 1980s, there were other Tibetan discourses circulating in the wider world. In particular, the Bönpo appealed to the Western imagination, thus as Tibetan concepts and imaginings came to be linked to a wider global vision of Kailas-Manasarovar it was to a large extent a Bön Kailas that became paramount. In that discourse it was the imagined—and claimed—timeless esotericism of Bön that often bet- ter chimed with New Age sensibilities than the monastic discipline of Tibetan Buddhism.

51 Taylor (2001: 131); the published text is Śrīchakrasambhāra Tantra, Ranchi, Tantrik Text series under the general editorship of Arthur Avalon (Sir John Woodroffe), 1918. Conclusions

Myths and Realities

Rather in the manner of Stanley Tambiah’s “Galactic polities”, sacred sites rise and fall in prominence over time. Here we have seen the process by which one sacred complex was elevated to a status transcending local, regional and national sacralities; how Tibet’s Mount Kailas—rather than a Mount Baekdu or a Doi Ang Salung—became a universal sacred centre attracting pilgrims from every land. We began by outlining the ubiquitous nature of sacred mountains in the realms from Persia to the China sea and the characteristic association of such mountains with esoteric spiritual practitioners (‘renunciates’) who were con- sidered to be equipped with the magical powers necessary to venture to these divine realms. We saw that both the cosmological concept of a central World- mountain and the specific sacrality of certain earthly mountains is indicated in Asian sources dating back well over 2,000 years. In what we termed the West- ern Himalayan Cultural Complex however, it was lakes that were a focus of sacred geography. Their sacrality was primarily associated with the residence there of Naga deities and the lake now known as Manasarovar was sacred in that context before the sacrality of Kailas was superimposed onto it and the new complex re-imagined as one of many dyadic formulations on the Tibetan plateau. Despite the universal assumption (Asian and Western) that the sacrality of Kailas is ancient, there is no real evidence that the earthly mountain was con- sidered sacred until the 12/13th centuries. Bön sources describing Kailas as the sacred centre of Zhang-zhung are many centuries after the fact and must be located within the context of later contestation for authority over the definition of regional sacred geography. That contest saw the subjugation and integra- tion of local culture(s) by the World-religions (or more precisely the compet- ing sects within those frameworks). These shared a central World-mountain- based cosmology and refocused landscape sacralities towards mountains at the expense of lakes. (The status of Manasarovar suggests that Zhang-zhung was more closely associated with the Western Himalayan Cultural Complex than with that we now call ‘Tibetan’.) The early Indic Kailas was an imagined heavenly world understood to be accessible only to powerful renunciates and Hero-kings. It was a literary trope, a poetical device, a metaphor, a symbol, and an ideal. References to it were not intended to be taken literally, even where they might reflect aspects of reality

© koninklijke brill nv, leiden, 2015 | doi: 10.1163/9789004306189_020 428 conclusions or actual landscapes and geographies. It was to the Indian reader what Camelot was to the English. Around the 4–5th centuries ce., representations of the heavenly Kailas mul- tiplied in Indic art, architecture, and literature. The poetry of Kalidasa, the com- position of the Abhidharmakosa, the consolidation of the Buddhist canon, the later sections of the Epics, and the early Puranas all date—very approximate- ly—to that period. Prominent among the ideological vectors favouring this process was a wider tendency towards the conceptual assimilation of heaven and earth as two forms of one reality, allied to the belief that an individual could become a deity (and thus transformed reach, or dwell, at Kailas). But that early Kailas is not the later Hindu Kailas for it was not the abode of Śiva, the attribution that became the defining characteristic of Kailas mountains in later medieval Indic understanding. So at the beginning of the second millennium ce., ‘Kailas’ did not relate in any meaningful way to the modern complex and there was no distinct associa- tion of the site with any World-religion identity. There were sacred toponyms, literary devices, cosmologies, and renunciates practicing near a sacred lake, but there was no formulated sacred geography, no mountain associated with either Śiva or Demchok or worshipped by circumambulation or pilgrimage. The acquisition of sacrality is a process, but the disjunctions here are so signifi- cant as to indicate a complete absence of continuities in regard to the sacrality of the mountain. Around the beginning of the 11th century Rinchen Zangpo, with his subju- gation of local deities and journeys via Kailas-Manasarovar and Kinnaur Kailas may have played a role in the sacralisation process at one or both sites, but a Buddhist sacred centre at Kailas-Manasarovar can only be identified from around the beginning of the 13th century. Even then the construction was not immediately established in its entirety. Buffetrille has pointed out for exam- ple, that (pilgrimage ‘opener’) Götsangpa’s immediate disciple is not recorded as ever making the circumambulation.1 There are clearly contradictions and fractures in these histories that even today are not entirely reconciled. None- the-less, it is from around the early 13th century that Kailas emerges at the centre of various claims indicative of efforts by the Kargyu and their sub-sects to situate their protective deity Demchok and their Abhidharma cosmology onto the Ngari landscape as a religio-political strategy. The politics of the Kargyu are clearly important here. They moved to the periphery, then claimed their new realm as the ontological centre of the world.

1 Buffetrille (1998: 20). conclusions 429

Their claims must be considered in the context of those forces—religious, polit- ical, and logistical—that drove them from the political centre of the emerg- ing Tibetan state to seek the patronage of the Guge-Purang kings. Other sects peripheral to Lhasa power structures, particularly the Nyingma and the Bön, also added their claims to association with the site, indicating its ‘open’ sta- tus. The Bön claim may have predated or inspired Kargyu response for the Ölmolungring construction emerges around the 11th century. But relevant Bön discourse seems to be concerned with internal identity construction or refor- mulation utilising history in sectarian competition for spiritual supremacy in the region. Their disconnection from any distinct Tise yullha and the multi- plicity of deities associated with Tise, as well as their unsystemised and poorly- formed character in Bön texts, is also notable in that regard. Existing sources do not verify the historical reality of their claimed Kailas-centred Zhang-zhung.

The Indic Kailas and the Tibetan Tise evolved as sacred centres in distinct historical trajectories. Although the Pasupatas may have traversed the Kailas- Manasarovar region as early as the first millennium ce., there is no evidence that they held it sacred nor is there any evidence for a Kailas-centred Śaivite sacred geography predating that of the Kargyu’s Buddhist construction in the 12/13th centuries. From around that period the Pasupatas faded out of Himala- yan history and were superceded by the Nath order of Śaivite renunciates, who began to penetrate into the upper Himalayas from at least the 13th century. An abundance of Himalayan toponyms testifying to their deployment of the nath suffix as a territorial and sacralising claim indicate their dynamic engage- ment in sacred landscape construction. They were in communication with the Kargyu, the two groups communicated at least in matters alchemical, with the Nath’s search for alchemical substances suggested as the impetus for the development of a religo-economic circuit north across the Himalayas via Man- asarovar and Tirthapuri to Chamba and on to Jvalamukhi. Informed by that engagement—knowledge of, for example, the myth of Śiva’s subjugation by Demchok—the Naths could well have concurred with the Kargyu’s identifica- tion of Tise as the Indic heavenly mountain (if not with the subjugation order). That would mean that we can date the understanding of the mountain in Tibet as the Indic Kailas as arising around the 13/14th centuries at the earliest, but clear evidence is lacking. By the 16/17th centuries the Naths were joined in these regions by the Giri order. Given that they are now the dominant renunciate presence at the var- ious Kailas mountains, it is possible that the full identification of the earthly Kailas as the abode of Śiva was made by the Giris and thus dates only to 430 conclusions this late medieval period. The absence of Kailas-Manasarovar from so many sacred geographical schemas that predate this period, such as the pithas (dis- cussed in Chapter 4), or the jyotirlingas (places where Śiva was considered to have manifested),2 would seem to support a later date than the 13/14th cen- turies. Kailas-Manasarovar comes more clearly into Indic focus with Garhwal and Kumaon consolidation in the upper Gangetic regions in the 16/17th cen- turies, roughly contemporary with the Kinnaur and Manimahesh sites becom- ing clearly visible in the sources. The Nath presence at Himalayan courts such as Chamba during this period when the Kailas-abode-of-Śiva model began to be applied in the Indian Himalayas suggests however, that the model could have been adopted by both sects. What is clear is that while a post-13th cen- tury date for the development of the Kailas-abode-of-Śiva model is possible, it only emerges prominently from around the 16/17th centuries. If those sites held earlier sacred meanings, this was at least a time of their re-imagining and systemisation. There were significant developments in the 17th century. The Panchen La- ma’s visit to Kailas preceeded the collapse of Guge, with Ladakh and half a century later Lhasa gaining control over Kailas-Manasarovar. To the south, the kingdoms of Garwhal (1623–1624) and Kumaon (1670) both despatched forces towards Ngari and in particular king Baz Chand was concerned with the north- ern reaches of his kingdom. Renunciate expansion into the upper Himalayas— and to Kailas-Manasarovar—was thus in the interests of the Indic kingdoms to the south as part of a process of territorial expansion involving the Sanskritisa- tion of local culture (discussed especially in Chapters 7 and 8). None-the-less, the Tibetans continued to retain political authority over areas to the south of the Himalayan watershed until Gurkha expansion paved the way for British- Indian sovereignty, so the impact of those campaigns was apparently limited. That Tibetan sovereignty extended over the sources of the Jadh and Bhai- girathi feeders of the Ganges river, including Gangotri, explains the medieval Indic focus on Badrinath as the Ganges’ source. Perhaps the alternate or ideal belief that Manasarovar was the source was found wanting by renunciates who travelled via Badrinath to Kailas, or they may have recognised Satopanth Tal as that heavenly lake, but such understandings were shaped by various fac- tors, including sectarian orientation. Disparate representations of these sacred geographies cannot necessarily be reconciled, or easily attributed to specific groups.

2 On which see Eck (2012: 201–214). conclusions 431

Our conclusions indicating a period of re-invention and systemisation not only of the renunciate movement but of Hindu sacred geography during the late medieval period seem broadly consistent with related findings. William Pinch has noted that four major and many minor famines brought large numbers of recruits into the renunciate movements during the 1554–1704 period,3 which probably stimulated renunciate expansion into the Himalayas. Mathew Clark identifies this as a formative era in the rise to prominence of the Dasnami orders, including those whose renunciates crossed the Himalayas to central Tibet. Already numerous there by the 1730s, their numbers increased during the time of the cosmopolitan 3rd Panchen Lama, when Chait Singh, Raja of Benares from 1770 to 1781, opened lines of communication with the Panchen using Giri renunciates as intermediaries.4 The Giri’s knowledge of Tibet, including its sacred geography, must have been greatly enhanced during this period, although certain aspects of their knowledge seem contradictory, as we have seen with the Ganges’ source issue. Hermann Goetz, the early historian of Chamba, concluded that the impact of the Mughals in the Himalayas brought about a “[H]indu revival [which] started in the second quarter of the 15th century, reached its apogee in the late 16th and early 17th centuries, and died in the course of the 18th century.”5 While reports of its death may be exaggerated, this is consistent with the understanding of a Sanskritic revitalisation around that period, and that does emerge in mahatmyas promoting new understandings of Himalayan sacred geography. It was not only renunciate activity that was then stimulated, popular pil- grimage to Pan-Indic shrines must always have been economically and logis- tically problematic for ordinary people, whose sacred journeys were largely restricted to local or regional shrines. But the developing economy and infra- structure—along with the security delivered by British rule—greatly stimu- lated such movement. Peter van der Veer informs us that;

Between 1780 and 1820 the number of pilgrims who went annually to Allahabad, Gaya and Benares may have trebled. Pilgrimage … was a status ritual which spread from old ruling dynasties via new dynasties and large landowners to emerging service and business classes at the end of the eighteenth and beginning of the nineteenth centuries.6

3 Pinch (2006: 80). 4 Huber (2008: 194–196); also see Petech (1988: 52–53, 61). 5 Goetz (1969: 101, 106, 112); also see Pollock (2001), (2011). 6 van der Veer (1988: 214). 432 conclusions

Thus while the rise to prominence of Kailas as an Indic pilgrimage was largely due to its promotion by Sherring and Hedin, imperatives to this and other pilgrimage sites such as Gangotri existed both in terms of a growing ‘middle class’ financially able to undertake major pilgrimages, and with an ideological movement emphasising Sanskritic sources and sacred geographies. Colonial government contributed through enhanced provision of law and order and with improved transport—notably the railways. There was, therefore, a nexus of factors that came together both to elevate the sanctity of Kailas and to enable it to function as a popular Indic pilgrimage site in the late colonial and modern period.7

On Origins

There are distinct layers of sacrality apparent at the various Kailas sites, with indications that at least three of them—Manimahesh, Kinnaur, and Kailas- Manasarovar8—had some early association with funeral rites (as with Sato- panth Tal). We can envisage a model of cremation grounds located between the periphery of village settlement and the mountains to which the souls of the departed were considered bound. The cremation ground then serves as a liminal place, with its spiritual power later acknowledged by the renunciates who dwelt there and sought to confront its demonic forces as part of their con- struction of charismatic power. Standing stones also represent a layer of sacrality, but their meaning is now lost to all but conjecture, with the dominant local beliefs in the earli- est recorded histories being worship of the powers of Nagas and territorial deities. Their continuing centrality in Western Himalayan Culture has been obscured by the dominant discourse of Hinduism and Buddhism, wherein local deities are reimagined and articulated as manifestations or lesser companions of World-religion deities—with implied social equivalencies in post-conquest society. Although the categories of Naga and territorial deity may overlap, as in Kamru, Nagas were commonly considered to reside outside of the settled vil- lage world, being particularly associated with bodies of water. In that sense they represent the wild and untamed in contrast to the settled and subjugated vil-

7 One fortuitous co-incidence noted by Mittal (1985: 76–77) is that trans-border trade declined as Indian prices rose in the 1920s, with porters then left available to carry for pilgrims. 8 Cutler (1996: 141–143) states that there were four charnel grounds at Kailas-Manasarovar. conclusions 433 lage environment. But other than sharing the wild spaces with the mountains, there is no specific link between Nagas and mountains as both were under- stood in the Western Himalayas.9 Indeed it is difficult to locate mountains in the local sacred schema. They seem of peripheral importance, prominent only as home to the souls of the departed ancestors or as the backdrop to lakes where dwelt the Nagas. The processes involved in each of these conceptions seem separate, although consistent with a model in which the acquisition of place sacredness progresses through numerous agencies and disparate traditions, each of which adds to the total sanctity of the site. Any systemisation of the different, even contradictory, elements occurred under certain political impulses, such as the incorporation of new territory and populations, or with the rise of significant new hegemonies and world-views such as that of modernity. In the case of sacred sites on the frontiers of both major empires and local kingdoms, their location only added to this fluidity and multiple agency, for frontiers are places of dynamic and organic transformation, cultural interaction, liminal status, and deliberate imprecise definition. In this liminal zone the relationship between the local and the Sanskritic has been one of on-going, though not necessarily continuous, interpenetration. The Vedic depiction of the Himalayas testifies to the antiquity of this relation- ship and the local does not necessarily, or always, predate the Sanskritic, as we saw in Chapter 7 in the case of the competing traditions at Harsar. Nor can the local be seen as a unitary whole. While we have used the concept of the Western Himalayan Cultural Complex as a holistic heuristic device, it must be understood as including disparate elements, and groups such as the Jads and the Gaddis have maintained distinct cultural markers that are not always com- patible with the proposed Cultural Complex. Indeed the extent to which any Himalayan community has enabled the creation of a Sanskritic cosmos has varied with time and location. There is a process by which Brahmans, both temple and renunciate, are agents of Sanskritisation and introduce Pan-Indic ritual and mythological forms whose frameworks may embrace local themes. But the extent to which the Brahmanical becomes hegemonic varies consid- erably, as we have seen in the case of Kinnaur. Furthermore, the Brahmanical agents both carried local traditions from their own region of origin (as we sus- pect in the case of the toponyms from the Mt Abu region) and were influenced

9 The local conception of the mountains as places of the dead could however, complement a Tibetan concept of them as frontier protectors. In both understandings the mountains existed as liminal zones on the edge of the known, ‘tamed’ world, beyond which lay the wild, the Other, and thus the threatening forces of the unknown. 434 conclusions by those of the new realm,10 while certain conceptions may have always been shared. The river goddess Gangama for example, seems to have been conceived of by both local and Sanskritic traditions and the two may have met as recently as the early 19th century, under Gurkha auspices. Modern political borders do not deliminate cultural borders, and elements identified here as characteristic of the Western Himalayan Cultural Complex clearly overlap with local traditions in the Zhang-zhung/western Tibet regions discussed here. While a separate study would be required to clarify these issues, we must note here that while there is plentiful evidence for the sacrality of lakes in local western Tibetan traditions prior to the hegemony of World-religions, evidence for the sacrality of mountains is lacking. If there were groups whose world-view identified and sacralised specific mountains prior to the introduc- tion of the Meru cosmological concept they have left little trace in the Indo- Tibetan historical record, at most we might allow the possibility of groups that drew on Turkic or Chinese concepts of sacred mountain geography having been present in the region.

On Toponyms

Among the disparate elements of the various bodies of knowledge concerning Kailas-Manasarovar are its assorted toponyms. These articulate specific claims to the association of features of the sacred landscape with the traditions, world- views, and cultural Heroes of each body of knowledge. Toponyms are thus an area of contestation, even an apparently neutral agreed name such as modern Hardwar/Haridwar actually serves as a sectarian identity marker depending upon how it is pronounced.11 This contestation is clear in historical analysis particularly in regard to the imposition of Sanskritic or Tibetan toponyms in place of earlier names (probably) reflecting the Western Himalayan Cultural Complex. But it is also seen within those broader cultures as in, for example, the addition to a place name of the suffix nath, signifying the impact of the renunciate sect of that name. This is characteristic of Himalayan geography. Its toponyms are a process that not only reflect historical changes but which demonstrates extant layers of belief and culture.12

10 A point noted by Shulman (1980: 8–9). 11 As ‘Hardwar’ it is Śaivite, as ‘Haridwar’ it is Vaisnavite (Hari=Vishnu). 12 See Chayet (1997) for a discussion of some of these issues, not least transcriptions; and see Witzel (1993) for an enquiry into historical toponyms in the context of Nepali river names. conclusions 435

We have seen that in early texts such as those of Pali Buddhism there are toponyms not otherwise known, evidence of failed claims to the authority of representation of particular sacred sites (although the difficulty of identifying the sites referred to problematises this conclusion). What is notable is that we cannot always identify if a site is intended to be an earthly location, a heavenly realm, a metaphor, a reference to myth or some combination of these. Further problematising Himalayan toponyms is that certain mountains (and their place-centred deities), may “move”; Buffetrille has discussed four Tibetan mountains, including Kailas, where there are legends of their foreign origin, or of them flying away and/or being pinned down to prevent this.13 Territorial deities such as Badri or Kelangnaga also move, seemingly with (at least some of) their migrating or exiled followers. Thus we cannot assume that a toponym in one text, or from one period, necessarily equates to the same place in another, as demonstrated in the case of Gaumukh. Conceptually, both the toponym and the place it is attached to can move. Certain toponyms such as Suvarnabhumi, Uddiyana or Kailas are applied to different places which may be widely separated,14 be assigned different locations by different traditions, or be entirely mythical. In addition, we have seen that Kailas (and Meru) originally referred to a heavenly location, only later coming to be given an earthly setting, a process that does not seem unique. Thus, with very few exceptions, textual references to toponyms can only be understood within specific contexts.15 Even a toponym fixed since the earliest written sources does not necessarily represent a precise geographical feature over time. The gaṅgā (Ganges) river referred to in the 10th book of the Rgveda is certainly the river known by that name today. But its full course and origins were not known in Vedic times, numerous sections of it bear additional or perhaps earlier local toponyms, and in one prominent later understanding, all rivers are ultimately the Ganges (something that enables a ritualist to readily coin a new but appropriate ritual for any river [deity] encountered, even if unknown). Toponyms such as Badri(nath) and Jvala(mukhi) demonstrate the regional tendency towards a pre-World-religion naming of a place after the deity that is conceptualised as ruling over it. There may also be a previously unnoticed aspect to sacred toponyms that emerges from a wider regional approach to the

13 Buffetrille (1996: 77–89); and see Rgveda 10.44.8, which refers to moving mountains. 14 Suvarnabhumi sometimes refers to the upper Indus region, sometimes to a location in southeast Asia; see Tucci (1956: 93–105). Uddiyana usually refers to the vicinity of the Swat valley, but sometimes seems to be in Orissa or Bengal. 15 Porter (1993: 76), discussing the mythical journey to Kunlun by King Mu, states that, “these place names must be read as cosmological referents rather than geographical ones.” 436 conclusions analysis of Kailas; that another layer of meaning, or wider purpose, lies behind the etymology of some of the names associated with the site. It appears that deliberately or not, these toponyms served to broaden the appeal of the site beyond a single linguistic group by embracing homonyms and homophones in various local languages. In this context, social rather than linguistic, it is interesting to look more closely at the accepted etymologies of the modern Indic and Tibetan toponyms, Kailas, Manasarovar, Tise, and Mapham, which have been relatively stable for a millennium or more. As we have seen the toponym Kailāsa is considered to have non-Sanskritic roots, although it may be linked to the Sanskrit ‘crystal’ (kelāsa). While in the early colonial period it was widely reported as being commonly used to mean ‘snowy mountains’ generally, it was also applied to a heavenly mountain and to specific earthly mountains in various Himalayan locations, as well as being used as a metaphor. The toponym does not, however, seem to have cross-cultural significance in recorded history. As we have seen, Tise is capable of multiple interpretations. Like Kailas, the origins of the name may not be recoverable and the name appears in different spellings including Te-se and more commonly Ti-tse or Ti-tshe, thus allowing a considerable range of meanings within Tibetan or local languages, including simply ‘peak’.16 But in the accumulation of the sacred a toponym that embraces multiple understandings of its meaning can only augment the power and appeal of the site, enhancing its understanding in multiple concep- tions.17

16 Here I am indebted to Dan Martin, who points out that a passage in chapter five of the Mdzod-phug has te-tse (which could be identical to ti-tse) as corresponding to Tibetan rtse (‘point, peak’) and that if ‘ti’ is understood as the number ‘1’ or as an indefinite article, which does seem possible, then it would simply mean ‘a peak’; [cf; Kailas = ‘snowy mountain’!]. 17 If we allow the possibility of Chinese influence here, as Tucci (1949: 714) allows in the context of an equation of the se-mo female spirits with the Chinese wu via the Xi Xia word szǔ = Ch: wu, the toponym can manifest still more meanings. The Chinese character Ti has associations—including power, virtue, and charisma of the type associated with the Pacific concept of mana—as well as with the sacrifice. Ti was also a major Shang (16th–11th centuries bce) dynasty deity with power over nature—wind, rain, and so on—and Robert Eno defines Ti as both an honorific and as meaning a “father”, or “originator of a lineage” (Eno 1990; also see Smith 1968: 24). Indeed, claims Eno, “Ti stood at the apex of the spiritual hierarchy”, if not as a supreme deity in the monotheistic sense, then as the deified first ancestor spirit or lineage founder. Deriving from the root ‘father’, and associated with power and majesty, Ti was also used an honorific (Keightley 1978: 215; Smith 1968: 4–6; Eno 1990). Thus via Chinese, with Se as a people, (their?) place, or (their?) deity, Tise might conclusions 437

The Tibetans use several toponyms for Manasarovar. Maphampa (‘Unbeaten One’) is so-called, according to the 1896 guide-book to the complex, due to the lake-waters’ ‘unbeatable’ qualities. But there is a more colourful Buddhist myth associated with another toponym, Madrospa (‘Not/Never Warm One’). This tells of how for 12 years a King Myupam (sMyug sbam) compassionately fed everyone in need. The water discarded after cooking rice for them was put in a hole dug for that purpose, and so much was provided that this water became a lake, and henceforth remained cool.18 But this characteristic was not unique to Manasarovar. The Mandakini Lake was similarly described in Pali sources as always cool,19 and as we have seen, Madros was also the name of the Klu deity dwelling in the lake. Other accounts state that it “is given the name Ever-cool Lake because it is the palace of the Naga king Anavatapta, ‘Who Never Warms Up.’”20 We are thus confronted with an archaic strata in which the lake toponym in both Indic and Tibetan traditions was that of the resident Naga deity ruling over the lake, and a later Buddhacised mythology that subjugates the local deity with an alternative meaning. Manasarovar provides the clearest example of how multiple meanings may be reflected in a toponym. The common, if surprisingly colourless Indic myth explaining the name tells how rsis meditating in the vicinity of Kailas required

be understood as ‘Lord [mountain] of the Se’, ‘Lineage father of the Se’, or even ‘Sacrifice [place] of the Se’; titles that seem more appropriate than those Tucci suggested. In this context we may note that one of the systemised set of Five Sacred Mountains in China was Mount Tie/Tai, in Shandong. This is the site of an inscription ordered by the first Qin emperor, and of feng and shan sacrifices. Originally known as the Southern Mount Gang, there is another mountain to the north above Shandong known as Northern Mount Gang (see Harrist 2008: 156–162, 174–175, 219–222). Whether this complex might be linked to the Tibetan ‘Gangs’ Tise/Kailas (Gangs being a common prefix I have largely ignored here), requires a separate study, but for a discussion of the evolution of the concept of sets of sacred mountains with many parallels to Ti-se/Kailas, see Robson (2009); also see, Choi (2008, esp. 30–35). Both Buddhist and alchemical imperatives promoted on-going Sino-Indic exchanges from at least the 2nd century bce., on which see Sen (2003). 18 See Huber & Rigzin (1999: 140–141); also see, Filibeck (1988: 85). This legend (found in, for example, the Mūlasarvāstivāda-Vinayavastu) also includes the Tibetan origin myth of the Ganges, which is that when the lake overflowed it formed a ‘River of Rice Water’ = ’Bras-khu’i chu = the Ganges. 19 Malalasekera (1937) citing the sna.[Sāratthappakāsinī, Samyutta Commentary] ii.407. Always being cold seems an obvious feature of a Himalayan lake. But the toponym of Issyk Kul (Kyrygzstan) means ‘Warm Lake’. 20 Ricard (1994: 342); also see Staal (1990: 275). Thus Anavatapta and Madros are one and the same—a Naga. 438 conclusions a water source for their ritual ablutions. They petitioned Brahma, who created the lake for them;21 hence manas (‘mind’) + sarovar (‘lake’) = ‘Lake [created by] the mind [of Brahma]’. But other meanings might be understood. At the local level, for example, Mana probably reflects the name of the territorial deity of Mana village (some 8 kilometres north of Badrinath) and of the pass north from there to the Tibetan plateau. The lake may thus have been considered by Mana’s followers as within that deity’s realm. In addition, as noted, the Mahab- harata (3.83.88–89) “distinguishes between accessible (gamya) and inaccessi- ble (agamya) tīrthas, which can be reached by thought (manasā) alone and are occupied by … [various enlightened beings].”22 Thus (despite the slight variant in spelling [manasā/mānasa]), Manasarovar could also be understood as, ‘the [sacred] lake that is inaccessible to ordinary beings;’ a meaning entirely appro- priate to the early accounts of the site. Manas is also known elsewhere, as a river toponym in Bhutan and in Cen- tral Asia for example, and other homonyms might be suggested; Manas the hero of the Kyrgyz epic of that name,23 the Punjabi deity Manasā Devi24 and Manasā the Bengali goddess of snakes, who shares mythology with Kailas. A goddess Manasa was known in Bronze age Greece where she was connected to the Cretan snake cult, but her Bengali namesake—perhaps originally a local deity—emerged in Puranas from the early centuries of the second millennium ce. Like the lake, she was said to have been created by the mind, in this case, the mind of the rsi Kasyapa in response to people’s fear of serpents, and after her creation she travelled to Kailas and performed austerities there for a 1000 years. Siva then endowed her with mystical knowledge and sent her off to serve Krishna.25 Whether this mythology deliberately echoes accounts of Kailas with

21 Ramayana 1.23.7–8, (Goldman vol. 1: 170). 22 Bisschop (2006: 10). 23 On the Manas epic, see Hatto (1980: 300–327, esp., 320). 24 Bakker & Entwistle (1983: 11). 25 Smith (1980: 7, 14–16). Identical accounts of the mythology are found in the Devibhagavata- purana 9.48, 10–13, and the Brahmavaivarta Purana, 46.11–13. Also see the 15–17th century Bengali legend given by Dimock & Ramanujan (1964: 303); “One day, while sitting beside a lotus pond, Siva was overcome by the beauty of the place (Version b—Siva went to bathe in a lake called Sonadaha). Thinking of his wife he discharged his seed … [which eventually] … fell upon the head of Vasuki, king of the Nagas. From the seed, Vasuki’s mother Kadru fashioned a beautiful girl … named Manasa, and to whom Vasuki gave charge of snakes and of poison … Manasa and her snakes used to come up to Siva’s lotus pond. Because of the snakes, Siva was no longer able to pick lotuses there, and so he summoned Garuda to come and eat the snakes.” The reflection of Kailas mythology seems clear. conclusions 439 the subjugation of Madros and creation by enlightened minds is uncertain, but such associations are a fundamental aspect of the accumulation of the sacred.26 We must also consider that while the philologist, the linguist, and literate local elites might easily distinguish between mana, māna, maṇi, and so on, ordinary pilgrims and speakers of other dialects and languages may not have been so linguistically precise, particularly to the term as a prefix to sarovar (lake). In particular maṇi (Skrt: jewel) had auspicious connotations and obvi- ous potential application. But the point is not to propose or to accept any one meaning, but rather to recognise that a flexibility of interpretation could only enhance the sacrality of the site, not least at a cultural distance. Toponyms such as Manasarovar offered creative potential of interpretation. They were pregnant with cultural meanings and their application was not a philological development but a socio-political seed able to be cultivated by later cultural agents. As Rita Wright observed in regard to the Indus Valley, “[a]ncient peoples used a highly symbolic vocabulary in both literature and imagery to represent th[e] association of gods with specific natural phenomena.”27 There are, therefore, a considerable range of meanings encoded in sacred toponyms. A toponym can be the personal name of a deity or a human (renun- ciate or otherwise), and can be in more than one place, including both earthly and heavenly locations. The name of a sacred place can also be a literary device, a clever play on different words or words with different meanings and through deliberate ambiguities and slight variant spellings28 it can allow different tra- ditions a place within the sacred complex. Toponyms are creative devices available to reinforce world-views, but they are also capable of being read on multiple levels and in various contexts, and they allow retention of encoded meanings by peoples in motion. They were frequently replicated and are often distinguishable only by context.29 We must presume that the informed listener/reader understood those contexts, which are not often clear today. Ultimately, sacred toponyms float free of philological definition or geographical precision, straddling, like the imagined mountain, the heavenly and earthly worlds for a certain elasticity of identity could be useful, as is so often the case on a multi-ethnic frontier.

26 In these and other links between the western Himalayas and Bengal the Ganges might be understood as the highway of communication. 27 Wright (2010: 283). 28 A point noted by Bellezza in regard to Mount Targo; Bellezza (1997: n. 3 321–322). 29 cf. Huber (2008: 119) “The intentions behind and implications of the transfer and applica- tion of each foreign toponym into a new cultural and geographical context must thus be considered on an individual basis.” 440 conclusions

On Communitas

The vision of learned religious figures exchanging esoteric insights or respect- fully debating beliefs and practices in a spiritually harmonious atmosphere is an appealing one, at least to a modern Western audience. Given that Kailas- Manasarovar was sacred to many faiths and situated at a cultural cross-roads subject to influences from many cultures, it would be easy to assume that it was a significant site for such cultural exchanges. Such exchanges did occur, as we saw with the Kargyu and Nath renouncers at Jvalamukhi, and certain scholars have identified spiritual exchanges as a central and continuous process in the wider region. Although modernity may transform the picture, the findings of Victor Turner in regard to the experience of communitas among Christian pilgrims30 have been shown to be largely inapplicable to Asian pilgrimages. This is not least because in Hindu India caste rules are maintained more strictly in sacred spaces than secular,31 and when renunciates accompany pilgrimages there is a strong tendency for them to remain aloof from other groups.32 In any case it is problematic to assume that ordinary (‘householder’) pilgrims to Kailas- Manasarovar were particularly interested in other traditions.33 Pilgrims hold- ing to different belief systems inhabited different conceptual worlds and were (and are) engaged in a ritual display of faith within a particular religious ten- dency. It seems reasonable to assume that those who undertook difficult and dangerous mountain pilgrimages were among the more devout followers of

30 Victor Turner’s theories emerged in several works; see for example, Turner & Turner (1978). Roy (2011) describes behaviour consistent with communitas on the modern Kailas pilgrimage, suggesting that Hindu modernism may develop this aspect. For a modern example of interaction in our region, see Riaboff (2004). 31 Another factor undermining communitas is suggested by an account of pilgrimage to Kedarnath; “Everybody was walking like a snail, like a machine. Scarcely anybody had time, energy or desire to look after another; everybody was trying to keep himself alive to reach the destination safely, it was impossible for them to see who was falling or ailing”; Banerjee (1937: 82). But also see, Messerschmidt & Sharma (1982: 139–158). 32 See Aziz (1982: 125). 33 Cantwell (1995: 4) notes that at Rewalsar pilgrims show little interest in activities of other religions; also see, Bharati (1978: 82) in regard to Tibetan pilgrims to the (Sikh) Golden Temple; “… maybe their mutual lack of recognition, foisted on some sort of actual sym- biosis, represents a typical theme of the Indo-Tibetan interface in the circum-Himalayan region.” Also see Sumption (1975: 192), who states in regard to the Christian crusades that, “the pilgrims had little incentive to understand their hosts and viewed them with uncom- prehending contempt.” conclusions 441 their religious traditions and that they actually seem the least likely to have sought other forms of religious knowledge. Numerous practical barriers also acted to deter communication with pil- grims of other faiths in this region, not least language.34 Communication be- tween Indians and Tibetans, let alone those from other linguistic groups, must have been hindered by a lack of means to communicate, particularly in regard to complex religious vocabularies. And while Tibetan pilgrims of different faiths shared a language if not necessarily a dialect, Indians did not necessarily even share a language. Hindu dietary restrictions similarly acted against social interaction across cultures, although those restrictions did not apply to renun- ciates. Indeed the experiences of renunciates and pilgrims differed in many ways both ritual and conceptual. While we may conclude that ordinary pilgrims generally only had superficial relations with followers of other traditions the question of renunciate interaction is more complex. Mark Dyczkowski notes that, “[T]he Tantras and commentaries on both Buddhist and Śaiva-cum-Śākta sides of the fence contain admonitions to stay clear of one another if encoun- ters happen to take place in such sites.”35 Such admonitions, which highlight the divisions between traditions, should of course be read as indicating that these encounters did take place, otherwise there would be no need to warn against them.36 But they demonstrate that it was religious authorities (con- cerned with the preservation of the identity of their traditions) who tended to discourage communication and preserve divisions between faiths. How things might operate on the ground is suggested by the experience of Götsangpa, the Drigung Kargyupa. We noted how in 1214–1215 he was turned away by other renunciates in the Kailas region.37 Both resources and patronage would have been limited in remote sacred sites and must have dictated a limit

34 “Language was … the principal barrier”; Sumption (1975: 1). 35 Dyczkowski (2001: 145). 36 Even in modern times of more precisely defined religious distinction there are examples of ecumenical encounters among advanced religious practitioners; see, for example, oioc, l/p&s/11/201–4102; Charles Bell to Government of India, 1 August 1921. This concerns Pandit Lal Singh Upadhya, a Benares sadhu. Having met some Tibetan monks at Sarnath he went to Sakya monastery for five years and became a Buddhist monk. He then went to Samye monastery, took up Dzogchen practice and spent three years living in caves. He spent 17 years in Tibet, married a Tibetan, and was a “strong believer in Buddhism” but, being suspected of independence sympathies he was deported back to India at the colonial Government’s request. 37 Vitali (1996: n. 672 409). 442 conclusions to the number of such practitioners that a site could accommodate. Ultimately the various traditions were in competition and that was often formalised by competitive display. We saw (in Chapter 11) how Milarepa triumphed over Naro Bön-chung and Götsangpa outshone the Drigung practitioners in a magical contest, and such competitions—which imply a degree of interaction between competitors—may be common.38 But the displays seem to be more about individual status in the world of renunciates, a demonstration of charismatic as much as siddhi power. The losers are not so much defeated and downcast along with the tradition they represent, as shown to simply be of lesser—albeit still advanced—achievement. In discussing medieval “Buddhist appropriation of non Buddhist designa- tions and practices”, Davidson found that “all the religious traditions … arro- gated other’s religious activities when they appeared popular in [royal] court or on the street”, and that there was, “a range of attitudes towards other religious systems, with mutual antagonism the most frequent posture.” He reasonably concluded that “perhaps a more nuanced model would be that the various lines of transmission … in some areas … interacted, while in others they maintained concerted hostility.”39 Interaction between specialists certainly occurred in some areas, including within families.40 Intertextuality between Hindu, Buddhist, and Bön scriptures and biographies is frequently attested to,41 ritual forms were easily adaptable to different traditions, cosmology had similar conceptual roots, and alchemi- cal and medical knowledge was certainly exchanged cross-culturally.42 Tantra itself crossed barriers between sects and religions.43 In general when teachings were taken from another World-religion however, they were ‘translated’ into the idiom of the receiver although, for reasons of identity, it seems easier for Hinduism to acknowledge Buddhist origins than the reverse.44

38 I am advised by an anonymous Brill reader that a recent film by James Mallinson shows yoga and tapas competitions between renunciates at the Allahabad Kumbh mela. Huber (1999: 87–93) provides an extraordinary account of a collective display of yogic powers. 39 Davidson (2002: 191–192, 218). 40 Martin (2001: 177–178) notes that the Kargyupa pioneer Jigten Gönpo’s mother was a Bönpo while his father followed Bhairava (Śiva). Milarepa was also descended from a Bönpo family. 41 See for example, Sanderson (2001). 42 On which, see White (1996); also see Schaeffer (2002: 515–533). 43 White (2001: 8) states that, “medieval and precolonial Asian religions, rather than having been discrete Tantric Hindu, Buddhist, and Jain traditions, were, to a great extent, Hindu, Buddhist, and Jain varieties of an over-arching tradition called ‘Tantra.’” 44 There are references to at least two strands of Tibetan teachings passing to India. The first conclusions 443

A distinction can be drawn here in regard to forms of knowledge. Certain religious aspects such as issues of philosophy or of theology, conceptions of the realm of particular deities, or cultural manifestations of worship were unlikely to be negotiable in encounters at sacred sites. But empirical data, such as alchemical and medical knowledge might be exchanged because that knowl- edge had practical value and could be easily located in a new religious or cogni- tive framework. Ritual technology would seem to lie somewhere between those poles, with the basic structure of a ritual replicable in a different theological context. Kailas may thus have been a fulcrum of certain aspects of knowledge exchange, particularly those not dependent on a specific cultural context (such as ritual bathing). But a more refined position may be that cultural transmission was the function of only a very limited number of individuals. It is notable that certain figures, historical or ahistorical, are associated with the transmission of major bodies of religious knowledge throughout wide expanses of the Himalayas; the Pandavas, Sankaracharya, Padmasambhava, and to an extent Rinchen Zangpo (even if only the latter can be accepted as entirely historical). All are credited in some sense with the transmission and conversion of landscapes and peoples. This phenomena of prominent individuals to whom is attributed the trans- ference of repositories of religious culture through the Himalayas has been discussed by Lars Thomas Rodseth. In a wider discussion of the mechanisms of cultural evolution and transmission he identifies as key agents, “priestly trav- ellers” or “Gurus”; charismatic figures in the Weberian mode “who develop wide personal networks yet remain detached from local political and kinship insti- tutions.”45 Such “a priestly traveller arrives on the scene, performs a service for his kingly host, and in return is granted a local office or title upon which to build a following.” Thus in the words of Fredrik Barth, “whole, complex traditions of knowledge seem to be carried by single persons.”46 While several of those historically prominent in trans-Himalayan cultural transmission may be entirely mythical, suggesting they represent wider pro- cesses rather than individual agency, there are also historical individuals such as Rinchen Zangpo, Jigten Gönpo, or Khyung-sprul. They were missionary

concerns the Vedic rsi Vasiṣṭha who obtains teachings from the Buddha (or Śiva in the form of the Buddha!). This occurs in “Mahācīnā”, although this is generally identified with the Kinnaur region rather than actually Tibet (on which see Tucci, 1971a); Bharati (1969: 52–56); Bharati, (1966: 66–70). The other reference is to a “Lama lineage” of the Dasnami order instituted in Tibet by one Ved Giri, whose guru is said to have been Padmasambhava; see Clarke (2006: 70, 302). 45 Rodseth (1994: chapter v); also see Lewis (1994). 46 Rodseth (1994: chapter vii, 75; citing Barth 1990: 646). 444 conclusions agents of a World-religion who transformed local beliefs (at least in the under- standing of the dominant culture). They were not engaged in the conversion of followers of other World-religions but focussed on subjugating local deities. Again we see therefore, evidence that such religious divisions were important in Himalayan history and that when conversion between World-religions is indicated it is not attributed to charismatic individual agents, but is articulated in terms of what A.W. Macdonald called “ideological dramas.”47 The charis- matic individuals that regional history celebrates are its missionaries, those who transformed Western Himalayan Cultural beliefs and practices.

The Construction of Sacred Geography

Vedic culture allowed several, often inter-related means by which particu- lar geography was imagined as sacred. In addition to the tribal centre, these included territory sanctified by conquest with fire, by vratya encirclement, the holding of sacrifices, or by association with the tribal dead. (It would be consis- tent with practice elsewhere if some of these sites were previously held sacred by earlier populations, with their sanctity reformulated by the new overlords of the landscape, but we have no evidence for this in that period.) The sacred iden- tity of such places could be amplified over time by association with and expres- sion in art, mythology, cosmology, and numerous other cultural and imaginal realms. The territorial sacralisation process continued throughout Indic history, with new sites being added and existing sites raised in status (or disappear- ing), in an on-going process that at its heart is conceptually centred on both the atman-brahman identification and on the concept of India as Āryāvarta; the sacred land of the Aryan peoples. That enabled another means by which territory might be sacralised. Places originally located in the realm of deities could be identified with earthly sites, whose other-worldly aspect might be fully revealed only to the enlightened.48 Charismatic individuals, sects, communi- ties, and even historical events within local, regional and national narratives have all created sacred sites that attract ‘insiders’ and which usually seek to appeal to ‘outsiders’. While particular tendencies have attempted to impose fixed networks of sacred sites within traditions, the organic nature of the sacral-

47 Macdonald (1990: 203). 48 Aside from those sites discussed in this work, see for another example, Bakker (1986: 2–11) who concludes that Ayodhyā was originally an imaginary place. conclusions 445 isation process has acted against such conformity. Different groups concep- tualised different sacred spaces and in the widest sense there was and is no agreement on hierarchies or itineraries. Certainly by the Epic Period if not earlier, significant pilgrimage sites were closely associated with the categories of both Brahman ritualists and renunci- ates. We read for example, of an assembly of “Brahm[a]ns and ascetics, who had gathered from … great mountains and places of pilgrimage.”49 Although Himalayan sacred sites do not necessarily have resident temple Brahmans, or even temples, Brahmans have historically tended to predominate among pil- grim groups and their presence acts to legitimate the sacred claims of the site in the Sanskritic world. Those claims are frequently those advanced by renunci- ates, either articulating local claims in new formulations or acting on visionary intuition. Considerations of access, infrastructure and patronage then dictate whether temple Brahmans establish themselves on the basis of these claims or whether the site remains restricted to itinerant renunciates. It is usually the presence of resident Brahmans that legitimates and allows the presence of ordinary Hindu pilgrims. (While modern Kailas-Manasarovar has been an exception to this model, this may change, with numerous Brahmans now seek- ing to establish centres there.) There is also an understanding that sites where renunciates gather are sanc- tified by their presence. Being in the presence of the renunciates practicing there is part of the attraction of a pilgrimage site, not only, it must be remem- bered, in terms of spirituality, but in terms of human spectacle. The colourful appearance and theatrical performance of the renunciates is an important fac- tor in their appeal.50 There is, therefore, a model in which a developed pilgrimage site has both resident Brahmans ministering to pilgrims at site temples (and/or at the local water source), and renunciates residing in the vicinity. Both the religious func- tionaries and the pilgrims are generally devotees of the site deity in some form, and pilgrims may attend the ritual performances of both types of functionary and ultimately will financially support both. There is thus a degree of interde- pendency between the two types of functionary, who mutually reinforce the sanctity of the site. Another fundamental reinforcement of sanctity is the mythology attached to the place. The Indic Kailas is particularly rich in such associations; 1/3/4/5

49 MhB 3.78.15–20 (van Buitenen: vol. 2: 364). We might note that this quotation gives no indication of any major Brahman/renunciate division or inequality of status. 50 cf; the saying ‘Where there’s a shaman, there’s a showman’; (source uncertain). 446 conclusions or 7 rivers including the Ganges descending from heaven via Śiva’s tresses or Visnu’s toe, the stopping of that river and Bhaigirathi’s penance to Śiva to obtain its release, and so on. These are however, almost interchangeable with the mythology surrounding the Ganges source, as well as other sites such as Badri- nath. Sacred places are where shifting narratives intersect. As with the ‘Cow’s mouth’ at the Ganges source, myths may be applied and reapplied to real or to shifting and imagined landscapes as often as new visionary constructions are required.51 The myths are dynamic vehicles which, while retaining their essen- tial structures and characteristics allow localised manifestations and construc- tions to emerge within specific cultural boundaries according to the internal logics of that culture, with or without external stimuli. Situating Kailas-Manasarovar within wider Asian history demonstrates the extent to which so many features of the sacred site are generic rather than unique or specific. Comparison with other sacred geographies shows that countless numbers of sites have a myth of heavenly descent there by an ances- tor, are claimed as the centre of the universe,52 identified as the home of Śiva or Demchok, and so on. Most, if not all of these sites attract renunciates and their surrounds are dotted with sacred representations and features that mark, dominate and demarcate the landscape as part of a spiritual heritage or sectar- ian tradition. Sacred geography is a process of accumulation and the layering of co-existent sacralities at Kailas is remarkable; little is ever entirely discarded or lost even when it is updated, as in the case of the replacement of the Ganges by the Karnali in the formulation of the four rivers originating at Manasarovar. The congregation of sacred features at Kailas was the product of bodies of knowledge held by numerous different traditions that (aside from the New Age) might be broadly subsumed under the titles of Bön, Buddhist, Hindu, and perhaps Western Himalayan. Followers of those traditions were certainly equipped to distinguish the poetic and the empirically verifiable, as the Sakya Pandita certainly did. In addition, the difference was explicable in terms of the

51 Huber (2008: 65) notes that “text-critical analysis often reveals that a great many pilgrim- age accounts are nothing but collages … of retold narratives.” Thus as Blezer (2011a: 139) states, it is “… crucially important … to appreciate the narrative configuration of Tibetan materials that we should like to read as sources for “history”.We first need to trace carefully the context and development of the particular narremes that are engaged in the (hi)story, before we come into a position to evaluate possible referents in time and space with suf- ficient certainty.” 52 In addition to those already noted, see for example, van Kooij (1972: 9) concerning a cave on Mt Nī near Kāmarūpa, residence of the goddess Kāmākhya and identified as the centre of the cosmos by her followers; also see Burghart (1983: 370). conclusions 447 common belief that there were layers of knowledge, with the higher levels only revealed to those spiritually equipped for that knowledge. But the accumulation of sacrality had significant consequences when Charles Sherring transformed selections from these layers of knowledge into a new, and almost hegemonic body of knowledge concerning the sacred site, for the explicitly poetic was given the same value as other expressions of knowl- edge. Sherring’s construction privileged Kailas in the Western imagination and paved the way for the additional constructions of Hedin, Anagorika Govinda, and others. Thus it was the colonial context that stimulated the rise of Kailas, rather than any other sacred mountain, to global stature. Yet it would be sim- plistic to isolate this construction as divorced from indigenous processes.53 In empowering the poetic Sherring was acting within indigenous understandings, and Sherring did not transform the profane into the sacred, rather he changed the ways in which the sacred was understood and laid the foundations for the transformation of a regional pilgrimage into a World-centre. In so doing, he was in many ways fulfilling the visions and intentions of earlier redacters such as Jigten Gönpo, and had Sherring been a Śaivite Brahman or a Kargyu renouncer rather than a British official his construction would still have had validity in the Indo-Tibetan context. It was largely revelation that identified a sacred place in both Indic tradition as we have seen in the case of Swami Tapovan, and in the Tibetan traditions.54 As the author of a Tibetan guide-book to Kailas stated;

Accordingly, as for my assertions [about landscape features] that “This is a deity, and this is its palace,” it is inappropriate to hold heretical views which consider these to be exaggerations merely for the reason that they are invisible to ordinary perception.55

It is this positive acceptance of revelation that meant that a Kalidasa could be as influential in the construction of a sacred Kailas as a Jigten Gönpo or a

53 Colonial modernity acted in a multi-faceted way on traditional society, drawing out pre-existing tendencies and empowering different elements from those that might have emerged with more lineal historical trajectories. Whether the Manasakhanda emerged from a late flowering of Sanskritic culture or was an early response to a demand from colonial interests is unclear but it is apparently the first Sanskrit text that allowed that an ordinary Indian pilgrim might venture to Kailas, hitherto textually reserved for renunci- ates. If this was an indigenous creation, then it was fortunate conjunction of interests with those of the colonial. 54 Ramble (1999: 15). 55 Huber (1999: 138); also see Buffetrille (2000: 93). 448 conclusions

Götsangpa. But the acceptance of revelation went beyond acceptance of the mythical, the miraculous, and the poetic vision. It extended to the positive reception given to scientific geography by the Hindu intellectual classes, the Brahman priests and renunciates. That dynamic flexibility allowed constant re-conception and reinterpretation of earlier knowledge, and once the scien- tific geography of the Kailas region was established by the Survey of India the Himalayan renunciates quickly adapted to this new worldview. Yet while these indigenous intellectuals acknowledged the validity of scientific geogra- phy, they believed that it must be compatible with the perfect knowledge of the ancient sages. Thus in articulating a new sacred geography they—like the modern Bönpo—saw themselves as heirs to a much-older wisdom, revealing a lost knowledge that they were privileged to recover.

Renunciation and Multiple Histories

Throughout this work we have seen the centrality of renunciation to Kailas histories and in the wider context it appears that they are primary agents in the construction of Asian mountain pilgrimages. Renunciates, or at least some form of esoteric and socially liminal ritualists such as the Magi or wu, were a powerful force throughout Asia and were associated with practice in remote locations such as mountains. In the Indic world, whether in early exploration or rsi ideal, as Tantric cremation ground ritualist, alchemical trader, or royal priest, renunciates have been at the centre of the cultural tendencies that shaped the understanding of Kailas. As a category equipped to deal with the impure and the demonic and to transact with the powers of the landscape, renunciates were ‘spiritual astronauts’ going beyond the known and tamed world, conceptually embracing an array of means of dealing with the world beyond, including the sword. They were the missionary arm of expansionist cultures. Indo-Tibetan renunciation has survived because of the vitality and organic nature of its forms of propagation. While beliefs change and develop and institutions wax and wane, renunciation as a category flows like a wave to fit whatever niche is appropriate to the social circumstances—now trader, now mercenary, now ritualist, now thinly cloaked with monasticism. Renunciates’ social and geographical freedom and defining characteristic of motion, allied to their self-authority within broad traditions, has enabled them to adapt to changing circumstances and thus to survive and prosper as a category. There are of course differences between Buddhist and Hindu renunciation and their strategies of landscape conversion also vary, as we have seen. In conclusions 449 the Buddhist case, Kailas was Buddhacised through the installation of the mandala of Demchok. Hinduism did not fully embrace that mandala device but in a related understanding, utilised the concept of Śiva dwelling on a mountain,56 with the surrounding territory envisaged as Śiva bhumi: ‘Śiva’s land’. Like the Chinese model of a central and four surrounding mountains, these formulations were transferable and infinitely replicable, but they gained their widest renown when applied to our mountain in Tibet. None-the-less, without the appropriate elite patronage it is doubtful either claim would be more than a historical curiosity. It was because these claims served political interests that they were patronised and thus prospered. In each case World-religions projected landscape claims back into history, appealing to the authority of the past. Bönpo, Buddhists and Śaivite Hindus all claimed ancient sanction for their visions and imaginings, and here they were well-served by the general cultural absence of critical analysis of such claims, the Sakya Pandita being a notable exception. This mythology of Heroic founding proved a successful legitimising device, powerful enough for aspects of its vision to be accepted as history even by modern academic commentators. Some almost universal human need for such visions seems to have all-too-often arrested critical faculties and supported invented traditions, but as inspira- tional as the visions may be these competing imaginaires are essentially claims to power over the culture of others. There was no one, timeless Kailas, no universal power place beyond that of the human imagination. The meaning of Kailas mountains varied over time, there was no single model and ultimately there remain many histories. These might be divided into the academic and the religious, or more precisely, the sci- entific and the visionary, with each choosing according to their own worldview, but no ultimate reconciliation is possible. Nor can the history of Manimahesh or Sri Kailas, of Kedarnath or Gangotri or any other such site be subsumed under that of another. Each of these sites had their own histories, for there never was a single centre, only a potentially infinite number of sites claiming central- ity.

56 Śiva’s residence on a mountain does not necessarily lead to its being a Kailas mountain, however. Sir Aurel Stein climbed Kashmir’s Haramukh (Hara= Śiva: i.e. ‘Śiva’s mouth’) peak in Kashmir and visited its attendant lakes Gagabal and Budabror (the latter at least a Naga lake and place of post-cremation ash-scattering). He noted that this was locally considered the centre of Kashmir with Śiva considered to reside on the peak; (and he is also considered resident on other peaks in south India). http://www.siraurelstein.org.uk/ pilgrim3.html, accessed 29 June 2010. We might also note the photographer Arthur Neve (1900: 83–84), refers to Haramouk as the source of the Ganges and the home of Śiva. 450 conclusions

A Modern Postscript

This is a work of history concerned with how Kailas-Manasarovar came to be understood today. A separate study would be needed to describe not only the development of the site since the 1980s, the agencies behind that development, and the political and nationalist aspects of this process, but to consider wider issues such as the relationship between pilgrimage and tourism, environmental problems in the development process there, or the role of modern media in the promotion of Kailas-Manasarovar. None-the-less, we might briefly note that following the communist Chinese invasion of Tibet in 1950, Kailas was closed to Europeans and in 1959 the Sino- Indian border was closed and Indian pilgrims could no longer visit the site. Tibetan religious activities in the region were then suppressed by the Chi- nese, and monasteries and other religious structures there were systematically destroyed. For 22 years China’s atheist regime contested the sacred traditions of the site with the same brutal insensitivity that they applied to religious expres- sion elsewhere in China, particularly that by the non-Han races and minority groups. Within India, the deterioration of relations with their northern neigh- bour that culminated in the Sino-Indian war of 1962 had the effect of greatly stimulating road-building in the Himalayas. That development underpinned a subsequent massive expansion in the number of pilgrims to the sacred spaces of the Himalayas. In the late 1970s, by which time it was clear that anti-religious policies had failed, China began a gradual process of opening up to foreign tourism. Kailas was at the forefront of that process, credit for which is claimed by an Iyer Brah- min, Dr. Subramaniam Swamy. Swamy, a Mandarin-speaking former Harvard economics professor—and something of a self-publicist—was a founder of the Janata party and served five terms in the Indian parliament. He reportedly raised the issue of reopening Kailas to Indian pilgrims when he met China’s Deputy Prime Minister in Beijing in 1978. The Deputy, who had never heard of the mountain, promised to investigate the issue and when Swamy returned to China in 1980 he was informed that the site would be reopened after restora- tion of roads and temples there. In April 1981 Swamy discussed the issue with Deng Xiao-Peng, and arrangements were finalised during the visit to India by the Chinese Foreign Minister in July of that year. Two months later, the first batch of a total of 60 Indian pilgrims set off for Kailas via Almora and the Lipu Lekh pass; Subramaniam Swamy was among them.57

57 Bedi & Swamy (1984: 7–8; biography from wikipedia.org/wiki/Subramanian_Swamy. Ac- cessed 16 October 2010). conclusions 451

Since that first batch of 60 pilgrims were allowed to cross the Indian frontier, this official state-level pilgrimage has continued on an annual basis. From among the vast numbers of Indian passport holders who apply, pilgrims are selected by a lottery run by the Indian External Affairs Department. They are then catered for by the Kumaon Vikas Mandal, the tourist wing of the Uttarakhand state government, with logistical support from the Indo-Tibetan Border Police. Their numbers have gradually increased; in 2009 there were 16 official groups, each made up of 60 pilgrims. The amount charged to join these groups has also risen steadily, which effectively restricts the official pilgrimage to middle and upper-class Indian passport holders. But in addition to the official parties, thousands of Indians now visit Kailas with Chinese or Nepali tour groups. In 2009, along with tens of thousands of Chinese and other foreign tourists, there were an estimated 12,000 Indian pilgrims present at the site, most of whom had joined commercial tours to be there.58 The reopening of the Kailas region to Indian pilgrims coincided with a grad- ual easing of restrictions on Tibetan religious practices in the region; monaster- ies began to be rebuilt, hermitages were reoccupied, and businesses catering to pilgrims were established. These changes were soon witnessed by foreign vis- itors. Tibet was briefly, and perhaps accidently, opened to foreign tourists in 1981–1982 and several hundred took advantage of the opportunity to go there, although none apparently reached Kailas. But in 1984, with the full opening of Tibet to foreigners, it became possible for them to reach Kailas from within China-Tibet or from Nepal (although not directly from India).59 Despite occa- sional interruptions of access to the site due to political protests in Tibet, the number of tourists has since grown exponentially and the site is now served by a regional airport. While individuals hitch-hike or even cycle there, the major- ity travel on organised tours. These are advertised not just in Kathmandu or California, but in Russia and in New Zealand, in Japan, Brazil and most of the developed world, while in another development worthy of a separate study, Han Chinese have become a major part of the regional tourist market. These visitors come to Kailas with an understanding of the site that has been constructed by Sherring, Hedin, Pranavananda and Anagorika Govinda. Those steeped in the modern body of systemised knowledge naturally express

58 http://www.tibetinfonet.net/content/update/163; accessed 15 February 2011. 59 On 1980s Western travellers to Kailas, see Snelling (1990: 313–378). His conclusion that Victor Chan and Bradley Rowe were the first Westerners there since the 1940s is incorrect, however. The journey of a New Zealander, Steven from Dunedin, who visited the site in 1984 and whose sirname I did not record, predated them by some months. This is a reminder of the extent to which our histories rely on those who record their activities. 452 conclusions their experiences of the place within that framework. But the understanding of Kailas as a spiritual place beyond the profane world has been challenged by the influx of visitors whose expectations are not always met. Western Bud- dhist and Kailas chronicler John Snelling quoted a 1988 interview with British traveller Bradley Rowe, who reported “… a few people who spent a couple of months in the area, hanging out mostly in caves. They were members of the ganja-smoking [i.e. hashish or marijuana using] community …” Snelling con- cluded that such developments “had altered the tone and atmosphere of the place.”60 But Snelling’s was an imagined “tone and atmosphere”, an imagined utopia. This was actually the Zone of Consecrated Warriors, Tantric renunciates, al- chemical pioneers, and charismatic showmen, not to mention bandits and passing armies. It was a place of contestation, where local beliefs in the sacral- ity of lakes where the serpent-deities dwelt had been overlaid by the histories, mythologies, and understandings of three colonising World-religions. Kailas- Manasarovar is a manifestation of diverse bodies of knowledge synthesised as a result of colonial imperatives, influences, and constructions. It is not a place of timeless sanctity but a conjunction of dynamic historical processes, social, political, religious, and economic, and those processes were as complex and multi-dimensional as those at any other site of human expression and histori- cal meaning.

60 Snelling (1990: 371–372); also see Bellezza (1995); and see Giri (2000: esp., 174–175) on difficulties experienced there by pilgrims. On imagined purity, cf. Harrist (2008: 271) who refers to Zhang Dai, a young man of privilege who visited Mount Tai, the sacred Chinese mountain, around 1628. The young man kept himself apart from the hordes of common pilgrims and attendant beggars, later writing that the, “beggars exploited Mt. Tai for money, while the visitors exploited Mt. Tai for fame. The land of Mt. Tai, once pure, was now everywhere desecrated by these two groups.” Bibliography

Abbreviations atbs Arbeitskreis für Tibetische und Buddhistische Studien (Universität Wien) cup Cambridge University Press ebhr European Bulletin of Himalayan Research icrhc The Institute for Comparative Research in Human Culture ieshr Indian Economic and Social History Review iias International Institute for Asian Studies (Leiden) iitbs International Institute for Tibetan and Buddhist Studies IsMEO Istituto Italiano per il Medio ed Estremo Oriente jaar Journal of the American Academy of Religion jaos Journal of the American Oriental Society jras Journal of the Royal Asiatic Society lwta Library of Tibetan Works and Archives nit Namgyal Institute of Tibetology oup Oxford University Press SOASBull. Bulletin of the School of Oriental and African Studies suny State University of New York vöaw Verlag der Österreichischen Akademie der Wissenschaften wzks Weiner Zeitschrift fur die Kunde Sudasiens

Abhedananda, Swami. 1987. Swami Abhedananda’s Journey into Kashmir and Tibet. Calcutta: Ramakrishna Vedanta Math. Achard, Jean Luc. 2004. Bon Po Hidden Hidden Treasures: A Catalogue of Gter Ston Bde Chen Gling Pa’s Collected Revelations. Leiden: Brill. , 2008. Enlightened Rainbows: The Life and Works of Shardza Tashi Gyeltsen. Leiden: Brill. Acharya, Jayaraj. (ed./trans.) 2000. The Nepāla-Māhātmya of Skandapurāna: Legends on the Sacred Places and Deities of Nepal. Jaipur/New Delhi: Nirala publications. Adriansen, R., H.T. Bakker, & H. Isaacson. 1988. The Skandapurāṇa, Volume 1; adhyāyas 1–25, Critically Edited with Prolegomena and English Synoposis. Groningen: Egbert Forsten. , 1994. “Towards a Critical Edition of the Skandapurāṇa.” Indo-Iranian Journal 37.4; 325–331. Agarwala, A.P. (ed.) 2003. Uttaranchal. The Abode of the Gods. New Delhi: Nest and Wings. Agrawala, Vasudeva. 1963. Matsya Purāna a study: an exposition of the ancient Purāṇa- Vidyā. Varanasi: All-India Kashiraj Trust. 454 bibliography

Ahmad, Zairuddhin. 1968. “New Light on the Tibet-Ladakh-Mughal War of 1679–1684.” East and West 18.3–4; 340–361. Ahuja, Ravi. 2003. ‘“The Bridge Builders’: some notes on railways, pilgrimage and the British ‘Civilizing Mission’ in colonial India.” In Colonialism as Civilizing Mission: The Case of British India. H. Fischer-Tine & M. Mann (eds.) London: Anthem Press; 195–216. Akasoy, Anna, Charles Burnett, and Ronit Yoeli-Tlalim (eds.) 2011. Islam and Tibet. Interactions along the Musk Routes. Farnham: Ashgate. Alay, Josep Lluís. 2011. “The Early Years of Khyung sprul rin po che: Hor (1897–1919).” Revue d’Etudes Tibétaines 20; 205–230. Aldenderfer, Mark. 2011. “The Material Correlates of Religious Practice in far Western Tibet: 500bce–500ce.” In Emerging Bon: The Formation of Bon Traditions in Tibet at the Turn of the First Millennium ad. Henk Blezer (ed.) Halle: iitbs; 13–33. Aldenderfer, Mark, and Zhang Yinong. 2004. “The Prehistory of the Tibetan Plateau to the Seventh Centurya.d.: Perspectives and Research from China and the West Since 1950.” Journal of World Prehistory 18.1; 1–55. Aldrich, Michael R. 1977. “Tantric Cannabis Use In India.” Journal of Psychedelic Drugs (Beloit/Wisc.) 9.3; 227–233. Ali, S.M. 1983. The Geography of the Puranas. Delhi: People’s Publishing House, (first published, 1966). Allen, Charles. 1982. A Mountain in Tibet: The Search for Mount Kailas and the Sources of the Great Rivers of India. London: Andre Deutsch. Aris, Michael. 1979. Bhutan: The Early History of a Himalayan Kingdom. Warminster: Aris & Phillips. , 1988. Hidden Treasures and Secret Lives: A Study of Pemalingpa (1450–1521) and the Sixth Dalai Lama (1683–1706). Simla: Indian Institute of Advanced Studies. Aris, Michael, and Aung San Suu Kyi (eds.) 1980. Tibetan Studies in Honour of Hugh Richardson: Proceedings of the International Seminar of Tibetan Studies, Oxford 1979. Warminster: Aris & Phillips. Atkinson, Edwin T. 1884. Notes on the History of the Himalayan Districts of North-west Provinces of India. Allahabad, (no publisher indicated; apparently a reprint of his Himalayan Gazetteer entries, the cited copy was consulted in soas library; er1527 /58437). Avalon, Arthur. (pseud. Woodroffe, John) (ed.) 1913.The Tantra of the Great Liberation: (Mahānirvāna Tantra) a translation from the Sanskrit with introduction and commen- tary. London: Luzac. Aziz, Barbara. 1982. “A Pilgrimage to Amarnath: The Hindu’s Search for Immortality.” Kailash ix. 2–3; 121–138. Bajpai, S.C. 1991. Kinnaur: A Restricted Land in the Himalaya. New Delhi: Indus. Bakker, Hans. 1986. Ayodhyā: The History of Ayodhya from the 7th Century bc to the bibliography 455

Middle of the 18th Century: Its Development into a Sacred Centre. Groningen: Egbert Forsten. , (ed.) 1990. The History of Sacred Places in India as Reflected in Traditional Literature: Papers on Pilgrimage in South Asia. [Panels of the viith World Sanskrit Conference, Leiden, 1987.] Leiden: Brill. , (ed.) 1992. The Sacred Centre as the Focus of Political Interests. Groningen: Egbert Forsten. , (ed.) 2004. Origin and Growth of the Purāṇic Text Corpus With Special Reference to the Skandapurāṇa. Delhi: Motilal Banarsidass. Bakker, Hans, & Alan Entwistle (eds.) 1983. Devi: The Worship of the Goddess and its Contribution to Indian Pilgrimage. A report on a seminar and excursion. Groningen: Institute of Indian Studies (University of Groningen). Balikci, Anna. 2008. Lamas, Shamans and Ancestors: Village Religion in Sikkim. Leiden: Brill. Banerjee, Nityanarayan. 1937. The Himalayas: In and Across. Calcutta: SunitiKumar Mondai. Bangha, Imre. 2011. “Writing Devotion. The Dynamics of Textual Transmission in the Kavitāvalī of Tulsīdās.” In Forms of Knowledge in Early Modern Asia. Sheldon Pollock (ed.) Durham/London: Duke University Press; 140–170. Bargiacchi, Enzo Gualtiero. 2008. A Bridge Across Two Cultures: Ippolito Desideri S.J. (1684–1733). A Brief Biography. Firenze: Istituto Geograpfico Militare. Barkhuis, Roelf. 1983. “The Four Pīṭhas, A report of research spring 1983.” In Devi: The Worship of the Goddess and its Contribution to Indian Pilgrimage. Hans Bakker and Alan Entwistle (eds.) Groningen: Institute of Indian Studies (University of Groningen); 58–69. Barron, P.(‘Pilgrim’). 1990. NoteofWanderingintheHimmala:ContainingDescriptionsof some of the Grandest Scenery of the Snowy Range; Among others of Naineetal. Nainital: Gyanodaya Prakashan, (first published, 1844). Barth, Frederick. 1990. “The Guru and the Conjurer: Transactions in Knowledge and the Shaping of Culture in Southeast Asia and Melanesia.” Man 25.4; 640–653. Basham, A.L. 1951. History and Doctrines of the Ajivakas. London University (soas) Ph.D. thesis. , 1991. The Wonder that was India. Calcutta: Rupa & co (first published, London, 1954). Bäumer, Betttina Sharada. 2010. “Text and Temple in Orissa: The Rajarani Temple of Bhubaneswar.” In Heritage Conservation and Research in India. 60 Years of Indo- Austrian Collaboration. Gabriela Krist, Tatjana Bayerova (eds.) Wien: Böhlau Verlag; 41–46. Bausani, Alessandro. 2000. Religion in Iran: from Zoroaster to Baha’ullah. n.y.: Biblio- theca Persica Press. 456 bibliography

Beal, Samuel. (trans.). 1969. Si-Yu-ki. Buddhist Records of the Western World translated from the Chinese of Hiuen Tsiang (a.d. 629). Delhi: Oriental Books (first published, 1884). Bechert, Heinz. 1984. “Buddhist Revival in East and West.” In The World of Buddhism. Heinz Bechert and Richard Gombrich (eds.) London: Thames and Hudson; 273–285. Beckwith, Christopher I. 1987. The Tibetan Empire in Central Asia: A History of the Struggle for Great Power among Tibetans, Turks, Arabs, and Chinese during the Early Middle Ages. Princeton: University Press. , 2009. “The Central Eurasian Culture Complex in the Tibetan Empire: The Imperial Cult and Early Buddhism.” In EintausendJahreasiatisch-europäischeBegeg- nung. Gedenkband für Dr.Peter Lindegger. Ruth Erken (ed.) Frankfurt am Main: Peter Lang; 221–238. , 2009a. Empires of the Silk Road. A History of Central Eurasia from the Bronze Age to the Present. Princeton/Oxford: Princeton University Press. , 2011. “On Zhangzhung and Bon.” In Emerging Bon: The Formation of Bon Traditions in Tibet at the Turn of the First Millennium ad. Henk Blezer (ed.) Halle: iitbs; 164–184. Bedi, Rahul K., & Subramanian Swamy. 1984. Kailas & Manasarovar. After 22 years. In Shiva’s Domain. New Delhi: Allied Publishers. Bell, C. 1989. “Religion and Chinese Culture: Toward an Assessment of ‘Popular Reli- gion’.”History of Religions 29.1; 35–57. Bellezza, John Vincent. 1993. “Quest for the Four Fountains of Tibet.” Himal 6.1; 41– 44. , 1995. “Kang Rimpoche Trashed and Commercialised.”Himal 8.1, 23–25. , 1997. Divine Dyads: Ancient Civilization in Tibet. Dharamsala: lwta. , 2001. Antiquities of Northern Tibet: Pre-Buddhist Archaeological Discoveries on the High Plateau. Delhi: Adroit Publishers. , 2002. Antiquities of Upper Tibet: Pre-Buddhist Archaeological Sites of the High Plateau (Findings of the Upper Tibet Circumnavigation Expedition, 2000). Delhi: Adroit Publishers. , 2005. Spirit-mediums, Sacred Mountains and Related Bon Textual Traditions in Upper Tibet: Calling Down the Gods. Leiden: Brill. , 2010. “‘gShen-rab Myi-bo’. His life and times according to Tibet’s earliest literary sources.”Revue d’Etudes Tibétaines 19; 31–118. , 2011. “Territorial Characteristics of the Pre-buddhist Zhang Zhung Paleo-cul- tural Entity—A Comparative Analysis of Archaeological Evidence and Popular Bon Literary Sources.” In Emerging Bon: The Formation of Bon Traditions in Tibet at the Turn of the First Millennium ad. Henk Blezer (ed.) Halle: iitbs; 53–116. Belton, Pauline. 1984. “The Concept of the Sacred Place in Hinduism.” London Univer- sity (soas) Religious Studies ma thesis. bibliography 457

Berar, A.S. 1999. Kailash Mansrovar: Thirty Days of Adventure and Ecstasy. Delhi: Na- tional Book Shop. Beresford, B. 1987. “The Lost Kingdom of Gu-gé.” The Middle Way 62.1; 37–42. Berg, E. 1994. “Journeys to the Holy Centre: the Study of Pilgrimage in Recent Himalayan Research.”ebhr 6; 3–19. Berglie, Per-Arne. 1980. “Mount Targo and Lake Dangra: A Contribution to the Religious Geography of Tibet.” In Tibetan Studies in Honour of Hugh Richardson: Proceedings of the International Seminar of Tibetan Studies, Oxford 1979. Michael Aris & Aung San Suu Kyi (eds.) Warminster: Aris & Phillips; 39–43. Bernbaum, Edwin. 1980. The Way to Shambala: A Search for the Mythical Kingdom Beyond the Himalaya. Cambridge ma.; Anchor Press. Berry, Scott. 1990. A Stranger in Tibet: The Adventures of a Zen Monk. London: Harper Collins. , 1995. Monks, Spies and a Soldier of Fortune: The Japanese in Tibet. London: Athlone. Berti, Daniela. 2009. “Divine Jurisdictions and Forms of Government in Himachal Pradesh”. In Territory, Soil and Society in South Asia. Berti, D., & Tarabout, G. (eds.) New Delhi: Manohar; 1–19. , 2009a. “Kings, Gods, and Political Leaders in Kullu (Himachal Pradesh).” In Bards and Mediums: History, Culture and Politics in the Central Himalayan Kingdoms. Lecomte-Tilouine, Marie (ed.) Almora: Almora Book Depot; 107–136. Bharati, Agehananda. 1963. “Pilgrimage in the Hindu Tradition.”History of Religions 3.1; 135–167. , 1965. The Tantric Tradition. London: Rider & co. , 1969. “References to Tibet in Medieval Indian Literary Documents.” Tibet Society Bulletin iii (Bloomington); 46–71. , 1978. “Actual and Ideal Himalayas: Hindu Views of the Mountains.” In Hima- layan Anthropology: The Indo-Tibetan Interface. J. Fisher (ed.) The Hague: Mouton; 77–82. Bhardwag, Surinder M. 1973. Hindu Places of Pilgrimage in India: a study in cultural Geography. Berkeley etc: University of California Press. Bharti, K.R. 2003. Travels to Highlands of Himachal. Delhi: Indus. Bhatt, M.S. 1987. Vedic Tantrism: A Study of Rgvidhāna of Śaunaka with Text and Trans- lation. Delhi etc: Motilal Banarsidass. Bhattacharji, R. 1986. “Jadh Ganga Valley, 1985.”Himalayan Journal 42; 49–55. Bhattacharya, Swapna. 1996–1997. “Some Observations on Mahayana-Tantric Beliefs in Ancient Myanmar.” Journal of Ancient Indian History xx; 52–67. Bhawa, Seema. 1998. Religion and Art of the Chamba Valley (a.d. 700–1300). Delhi: Agam Kala Prakashan. Birnbaum, Raoul. 1984. “Thoughts on T’ang Buddhist Mountain Traditions and Their Context.”Tang Studies 2: 5–23. 458 bibliography

Birtalan, A. 1995. “Typology of Stone Cairn Obos (Preliminary Report, Based on Mongo- lian Fieldwork Material Collected in 1991–1995).” In Tibetan Mountain Deities: Their Cults and Representations: Proceedings of the 7th International Seminar for Tibetan Studies, Graz 1995. Anne-Marie Blondeau (ed.) Wein: vöaw; 199–210. Bishop, Peter. 1989. The Myth of Shangri-la: Tibet, Travel Writing and the Western Cre- ation of Sacred Landscape. London: Athlone Press. Bisschop, Peter C. 2006. Early Śaivism and the Skandapurāṇa: Sects and Centres. Gronin- gen: Egbert Forsten. Bisschop, Peter C. and Arlo Griffiths. 2003. “The Pāśupata Observance (Atharvaveda- pariśiṣṭa 40).”Indo-Iranian Journal 46.4; 315–348. Bjerken, Zeff. 1998. “Cracking the Mirror: A Critical Geneology of Scholarship on Tibet- an Bon and the ‘Canonical’ Status of The Crystal Mirror of Doctrinal Systems.” The Tibet Journal 23.4; 92–107. , 2004. “Exorcising the Illusion of Bon ‘Shamans’: A Critical Geneology of Sha- manism in Tibetan Religions.”Revue d’Etudes Tibétaines 6: 4–59. Blavatsky, H.P. 1883. “Existence Of The Himalayan Mahatmas.” The Theosophist 3.51 (December); 98–99. , 1888. The Secret Doctrine. Pasadena: Theosophical University Press. Blezer, Henk. 2007. “Heaven My Blanket, Earth My Pillow: Wherever Rin po che lays his head down to rest is the original place of Bon.” Acta Orientalia (Oslo) 68; 75–112. , 2008. “sTon pa gShen rab, six marriages and many more funerals.” Revue d’Etudes Tibétaines 15; 421–479. , 2009. “The Silver Castle revisited—a few notes.” Acta Orientalia (Oslo) 70; 217–223. , 2010. “Greatly Perfected in Space and Time: Historicities of the Bon Aural Transmission from Zhang zhung.” In The Earth Ox Papers: Proceedings of the Inter- national Seminar on Tibetan and Himalayan Studies, Held at the Library of Tibetan Works and Archives, September 2009 on the Occasion of the “Thank you India” Year. Roberto Vitali (ed.) Dharamsala: lwta. (Tibet Journal special edition, vol. 34.3–35.2, 2009–2010); 71–160. , 2010a. “The Two Conquests of Zhang zhung and the many Lig-Kings of Bon: A structural analysis of the Bon ma nub pa’i gtan tshigs.” Edition, editions: L’écrit au Tibet, evolution et devenir. (Collectanea Himalayica 3) Anne Chayet, Cristina Scherrer-Schaub, Françoise Robin, Jean-Luc Achard (eds.) München: Indus Verlag; 19–63. , 2010b. “William of Ockham, Jan van Gorp and Tibetan Studies: Some Notes on Dating the mDo ’dus.” In Études tibétaines en l’honneur d’Anne Chayet, J.-L. Achard (ed.) Hautes Etudes Orientales—Extrême Orient vol. 49; 1–50. , 2011. “It all happened in Myi Yul Skyi Mthing: A crucial nexus of narratives pointing at the proto-heartland of Bon?” In Buddhist Himalaya: Studies in Religion, bibliography 459

History and Culture. Vol. 1 Tibet and the Himalaya. Alex McKay & Anna Balikci- Denjongpa (eds.) Gangtok: nit; 157–178. , 2011a. “In Search of the Heartland of Bon—Khyung Lung Dngul Mkhar the Silver Castle in the Garuḍa Valley”. In Emerging Bon: The Formation of Bon Traditions in Tibet at the Turn of the First Millennium ad. Henk Blezer (ed.), Halle: iitbs; 117–163. , (ed.) 2011b. Emerging Bon: The Formation of Bon Traditions in Tibet at the Turn of the First Millennium ad. Henk Blezer (ed.) Halle: iitbs. , 2013. “The Paradox of Bon Identity Discourse: Some Thoughts on the rMa Clan and on The Manner of bsGrags pa bon, ‘Eternal’ Bon, New Treasures, and New Bon.” In Buddhism and Nativism: Framing Identity Discourse in Buddhist Environments. Henk Blezer (ed.), Leiden: Brill; 123–157. , (ed.) 2013. Buddhism and Nativism: Framing Identity Discourse in Buddhist Environments, Leiden: Brill. , (Forthcoming) “Breaking the Paradigm: Tibetan Bon po-s and their Origin Nar- ratives.” In Emerging Religions: Breaking the (Religious) Historical Paradigm. Henk Blezer (ed.), Leiden: Brill. , (ed.) (Forthcoming). Emerging Religions: Breaking the (Religious) Historical Paradigm. Leiden: Brill. , (Forthcoming). The Three Pillars of Bon: Doctrine, ‘Location’ & Founder, Vol- ume i: Doctrine, Part i: Antecedents of Bon Religion in Tibet and Part ii: Tibetan Texts. Leiden: Brill. , (Forthcoming). The Three Pillars of Bon: Doctrine, ‘Location’ & Founder, Vol- ume ii, Part i: Location of Origin and Part ii: Apparatus. Leiden: Brill. Blondeau, Anne-Marie. (ed.) 1998. Tibetan Mountain Deities: Their Cults and Represen- tations: Proceedings of the 7th International Seminar for Tibetan Studies, Graz 1995. Wein: vöaw. Blondeau, Anne-Marie, and Ernst Steinkellner (eds.) 1996. Reflections of the Mountain: Essays on the History and Social Meaning of the Mountain Cult in Tibet and the Himalaya. Wien: vöaw. Bloomfield, Maurice. 1924. “On False Ascetics and Nuns in Hindu Fiction.” Journal of the American Oriental Society 44; 202–242. , 1978. The Atharvaveda and the Gopatha Brāhmaṇa. Delhi: Asian Publication Services, (first published, Strassburg, 1899). Boileau, Gilles. 2002. “Wu and shaman.”bsoas: 65.2; 350–378. Bollee, Willem. 2007. “A Note on the Pāsa Tradition in the Universal History of the Digambaras and Śvetâmbaras: Guṇabhadra, Mahāpurāṇa, Utt. 73 (Pārśva).”Interna- tional Journal of Jaina Studies 3.2: 1–60. Bosch, P. van den. 1944. Atharvaveda-Parisista: chapters 21–29, Introduction, Translation and Notes. PhD thesis. University of Groningen. 460 bibliography

Bouillier, Véronique. 1991. “Growth and decay of a Kanphata Yogi monastery in south- west Nepal.”ieshr 28.1: 151–170. , 1997. “Conflits autour d’un monastere Sannyāsī: documents du Kwāṭhanḍo maṭh à Bhatgaon (Népal).” In Les Habitants du Toit Du Monde: Études Recueillies en Hommage Àlexander W. Macdonald. Samten G. Karmay & Phillipe Sagant (eds.) Nanterre: Société d’ethnologie; 723–735. Boyce, Mary. 1975. A History of Zoroastrianism: Volume One: The Early Period. Leiden: Brill. Brashier, K.E. 2001–2002. “The Spirit Lord of Baishi Mountain: Feeding the Deities or Heeding the Yinyang.”Early China 26–27; 159–231. Brauen, Martin. 1997. Mandala. Sacred Circle in Tibetan Buddhism. London: Serindia. , (ed.) 2002. Peter Aufschnaiter’s Eight Years in Tibet. Bangkok: Orchid Press. , 2004. Dreamworld Tibet: Western Illusions. Bangkok: Orchid Press, (first published as Traaumwelt Tibet: Westliche Trugbilder. Berne: Verlag Paul Haupt; 2000). Bray, John. 2003. “Ladakhi and Bhutanese Enclaves in Tibet.” In Alex McKay (ed.) The History of Tibet: The Modern Period: 1895–1959. The Encounter with Modernity, vol. 3. London: RoutledgeCurzon; 191–201 (first published, 1997. In Recent Research on Ladakh 7. Theirry Dodin and Heinz Räther (eds.) Ulm: Ulmer Kulturanthropolo- gische Schriften; 89–104). Briggs, George Weston. 2001. Gorakhnāth and the Kānphaṭa Yogīs. Delhi: Motilal Banar- sidass (first published, Calcutta, 1938). Brinkhaus, H. 1991. “The Descent of the Nepalese Malla Dynasty as Reflected by Local Chroniclers.” jaos iii.i; 118–122. Brockington, John. 1988. The Sanskrit Epics. Leiden: Brill. , 2003. “The Sanskrit Epics”. In The Blackwell Companion to Hinduism. Gavin Flood (ed.) Oxford: Blackwell; 116–129. Broido, M. 1987. “Sa-skya Paṇḍita, the White Panacea and the Hva-Shang Doctrine.” Journal of the International Association of Buddhist Studies 10.2; 27–68. Bronkhorst, Johannes. 1998. The Two Sources of Indian Asceticism. Delhi: Motilal Banar- sidass (first published, Bern, 1993). , 2007. Greater Magadha: Studies in the Culture of Early India. Leiden: Brill. Brooks, Douglas Renfrew. 1996. Auspicious Wisdom: The Texts and Traditions of Śrīvidyā Śākta Tantrism in South India. Delhi: Manohar. Brough, John. 1953. The Early Brahmanical System of Gotra and Pravara: A Translation of the Gotra-Pravara-Mañjarī of Puruṣottama-Paṇdita, with an Introduction. Cam- bridge: cup. Brown, C.W. 1984. “‘The Goat is Mine, The Load is Yours’: Morphogenesis of ‘Bhotiya- Shauka’, u.p., India.” Lund: Lund Studies in Social Anthropology. Bruhn, Klaus. 1998. “Bibliography of Studies Connected with the Āvaśyaka-Commen- bibliography 461

taries.” Appendix in Catalogue of the Papers of Ernst Leumann in the Institute for the Culture and History of India and Tibet. Birte Plutat (comp.) Hamburg: Franz Steiner Verlag. Bue, Erberto Lo. 1988. “The Dharmamaṇḍala by Buddhaguhya.” In Orientalia Iosephi Tucci Memoriae Dicata. C. Cuverant, G. Gnoli and L. Lanciotti (eds.) (Serie Orientale Roma: 56.3) Rome: IsMEO; 787–801. Buffetrille, Katia. 1994. The Halase-Maratika Caves (Eastern Nepal). A sacred place claimed by both Hindus and Buddhists. Pondichery: Institut Française de Pondich- ery. , 1996. “One day the mountain will go away: Preliminary remarks on the flying mountains of Tibet.” In Reflections of the Mountain: Essays on the history and social meaning of the mountain cult in Tibet and the Himalaya. Anne-Marie Blondeau & Ernst Steinkellner (eds.) Wien: vöaw; 77–89. , 1997. “The Great Pilgrimage of A-myes rma-chen.” In Mandala and Landscape. A.W. Macdonald (ed.) New Delhi: d.k. Printworld; 75–132. , 1998. “Reflections on Pilgrimages to Sacred Mountains, Lakes and Caves.” In Pilgrimage in Tibet. Alex McKay (ed.) Richmond: Curzon Press; 18–34. , 2000. Pèlerins, Lamas et Visionnaires: Sources Orales et Écrites sur les Pèleri- nages Tibétains. Wien Universität: Arbeitskreis für Tibetische und Buddhistische Studien. , 2009. “Khyung mo Monastery (A’mdo) and Its ‘Map’ of ’Ol mo lung ring.” In Bon The Everlasting ReligionOf Tibet: Tibetan Studies in Honour of Professor David L. Snell- grove, New Horizons in Bon Studies 2. Samten G. Karmay and Donatella Rossi (eds.) (Special edition of)East and West 59.1–4; 313–326. , 2012. “Low Tricks and High Stakes Surrounding a Holy Place in Eastern Nepal: the Halesi-Māratika Caves.” In Revisiting Rituals in a Changing Tibetan World. Katia Buffetrille (ed.) Leiden: Brill; 163–208. , 2014. “The pilgrimage to Mount Kha ba dkar po: a metaphor for bar do?” In Searching for the Dharma, Finding Salvation—Buddhist Pilgrimage in Time and Space, Christoph Cueppers and Max Deeg (eds.) Lumbini: Lumbini Research Insti- tute; 197–220. Buffetrille, Katia, & Hildegard Diemberger (eds.) 2002.Territory and Identity in Tibet and the Himalayas, piats 2000: Tibetan Studies: Proceedings of the Ninth Seminar of the International Association for Tibetan Studies, Leiden 2000. Leiden: Brill. Buffetrille, Katia, & Robert Kostka. 2002. Kailash. Map of the Holiest Mountain in the World. Vienna. Weiner Studien zur Tibetologie und Buddhismuskunde. Buffetrille, Katia, & Donald S. Lopez (trans.) 2010. Burnouf, Eugène. Introduction to the History of Indian Buddhism. Chicago/London: University of Chicago Press (first published in French, 1884). Bujard, Marianne. 2009. “State and local cults in Han religion.” In Early Chinese Reli- 462 bibliography

gion. Part One: Shang through Han (1250bc–220ad) vol. 2. John Lagerwey and Marc Kalinowski (eds.) Leiden: Brill; 777–813. Burghart, Richard. 1978. “The Disappearance and Rediscovery of Janakpur.” Kailash: A Journal of Himalayan Studies 6.4; 257–287. , 1983. “Wandering Ascetics of the Ramanandi Sect.” History of Religion 22.4; 361–380. Callewaert, Winand. 1998. “On the Way to Kailash.” In Pilgrimage in Tibet. Alex McKay (ed.) Richmond: Curzon Press; 108–116. Campany, Robert Ford. 2009. Making Transcendents. Ascetics and Social Memory in Early Medieval China. Honolulu: University of Hawai‘i. Campbell, June. 1996. Traveller in Space: In Search of Female Identity in Tibetan Bud- dhism. London: Athlone. Cantwell, Cathy. 1983. “To Meditate Upon Consciousness as Vajra: Ritual ‘Killing and Liberation’ in the Rnying-Ma-Pa Tradition.” In ContributionsonTibetanandBuddhist Religion and Philosophy. Proceedings of the Csoma de Körös Symposium held at Velm- Vienna, Austria, 13–19 September, 1981, vol. 1. Ernst Steinkellner & Helmut Tauscher (eds.) Wien: atbs; 107–118. , 1995. “Rewalsar: Tibetan Refugees in a Buddhist Sacred Place.” The Tibet Jour- nal 20.1; 3–9. Caplan, P. 1973. “Ascetics in Western Nepal.” The Eastern Anthropologist 26.2; 173–182. Carpenter, David. 1992. “Language, Ritual and Society: Reflections on the Authority of the Veda in India.” jaar lx.i: 57–77. Castro, Andrea de, and David Templeman. (eds.) 2015. Asian Horizons: Tucci, Himala- yan, Indian, Buddhist and Central Asian Studies. Rome/Melbourne: Serie Orientale. Chakrabarti, Chinmoy. 2003. “My Silent Ascetic.”Himalayan Journal 59; 167–172. Chandler, David. 2000. A History of Cambodia. Bangkok: Silkworm Books (first pub- lished, Boulder: Westview Press). Chandola, Khenanand. 1987. Across the Himalayas through the ages: a study of relations between central Himalayas and western Tibet. New Delhi: Patriot. Chandra, Lokesh. (ed.) 1968. Tibetan Chronicles of Padma-dkar-po. New Delhi: Interna- tional Academy of Indian Culture. Chang, Garma C.C. (trans.) 1977. The Hundred Thousand Songs of Milarepa, 2 vols. Boulder: Shambala. Channa, Subhadra Mitra. 2013. The Inner and Outer Selves: Cosmology, Gender, and Ecology in the Himalayas. Delhi: oup. Chatterji, B.R. 1940. “Tholing Monastery in Western Tibet (A cultural link between Greater India, Pāla Bengal, and Tibet).” Journal of the United Provinces Historical Society xiii.ii; 30–34. Chaturvedi, B.K. n.d. [c. 1999?]Shankaracharya. Delhi: Books for all. Chayet, Anne. 1997. “Pays réel et pays sacré, réflexions sur les toponymes tibétains.” In bibliography 463

Les Habitants du Toit Du Monde: Études Recueillies en Hommage Àlexander W. Mac- donald. Samten G. Karmay & Phillipe Sagant (eds.) Nanterre: Société d’ethnologie; 35–51. Chhodak, Tenzing. 1978. “The 1901 Proclamation of H.H. Dalai Lama xiii.” The Tibet Journal 3.1; 30–38. Choephel, Gedun. 1978. The White Annals. Dharamsala: lwta. Choi, Mihwa. 2008. “Contesting Imaginaires in Death Rituals during the Northern Song Dynasty.” University of Chicago Divinity School PhD thesis. Clark, Mathew. 2006. The Daśnāmī-Saṃnyāsīs: The Integration of Ascetic Lineages into an Order. Leiden: Brill. Clarke, John. 1998. “Hindu Trading Pilgrims.” In Pilgrimage in Tibet. Alex McKay (ed.) Richmond: Curzon Press; 52–70. Cleary, Thomas. (trans.) 1986. The Flower Ornament Scripture: A Translation of the Avatamsaka Sūtra, 3 vols. Boston & London: Shambhala. Coe, Michael D. 2003. Angkor and the Khmer Civilization. London: Thames and Hud- son. Cohn, Bernard. 1964. “The Role of the Gosains in the Economy of Eighteenth and Nineteenth Century Upper India.”ieshr 1.4; 175–182. Colebrook, H.T. 1812. “On the Source of the Ganges in the Himádri or Emodus.” Asiatik Researches 11; 429–445. Cone, Margaret and Richard Gombrich. (ed/trans.) 1977. The Perfect Generosity of Prince Vessantara: A Buddhist Epic. Oxford: Clarendon Press. Cort, John. 2010. Framing the Jina: Narratives of Icons and Idols in Jain History. New York: Oxford University Press. Cousens, Diana. 2008. Triloknath. Monash University PhD. Thesis. , 2010. “Multiculturalism on the Tibetan border: the temple of Triloknāth in Lahul.” In New Views of Tibetan Culture. David Templeman (ed.) Caulfield: Monash University Press; 53–69. Cousins, Lance. 1996. “The dating of the Historical Buddha: A Review Article.” jras 3.6.1: 57–63. Cutler, Nathan S. 1996. “Mount Kailasa: Source for the Sacred in Early Indian and Tibetan Tradition.” California Institute of Integral Studies Ph.D. thesis. Dagkar, Namgyal Nyima. 1997. ‘Stag-gzig and Zhang-zhung in Bon Sources.’ In Tibetan Studies i and ii: Proceedings of the 7th Seminar of the International Association for Tibetan Studies, Graz 1995, vol. 2. Helmut Krasser et al., (eds.) Wien: vöaw; 687– 700. Dalmia, V. 1997. The Nationalization of Hindu Traditions: Bhāratendu Hariśchandra and Nineteenth century Banaras. Delhi: oup. Dalton, Jacob and Sam van Schaik. 2006. Tibetan Tantric Manuscripts from Dunhuang. A Descriptive Catalogue of the Stein Collection at the British Library. Leiden: Brill. 464 bibliography

Dandamayev, M., and V. Livshits. 1988. “Zattumēšu, a Magus in Babylonia.” Acta Iranica 28, Hommages et Opera Minora vol. xii, A Green Leaf, Papers in Honour of Professor Jes P. Asmussen. Leiden: Brill; 457–459. Daniels, Christine. 1994. “Defilement and Purification: Tibetan Buddhist Pilgrims at Bodhnath, Nepal”. University of Oxford D.Phil. thesis. Dargyay, Eva K. 1981. “Sangha and State in Imperial Tibet.” In Tibetan History and Language: Studies dedicated to Uray Géza on his seventieth birthday, vol. i. Ernst Steinkellner (ed.) Wien: atbs; 111–127. , 1994. “Srong-btsan Sgam-po of Tibet: Boddhisattva and King.” In Monks and Magicians. Religious Biographies in Asia. P. Granoff and K. Shinohara (eds.) Delhi: Motilal Banarsidass; 99–117 (first published, Ontario, 1988). Das, Sarat Chandra. 1983. A Tibetan Dictionary. Delhi: Motilal Banarsidass (first pub- lished, 1902). Das, Sur. 2012. “An Unpublished Account of Kinnauri Folklore.” (Introduced by Arik Moran). ebhr: 39; 9–40. David-Neel, Alexandra. 1971. With Mystics and Magicians in Tibet. n.y.: Dover, (first published, 1931.) Davidson, Ronald M. 1991. “Reflections on the Maheśvara Subjugation Myth: Indic Materials, Sa-skya-pa Apologetics, and the Birth of Heruka.” The Journal of the Inter- national Associations of Buddhist Studies 14.2; 197–234. , 1995. “The Boddhisattva Vajrapāṇi’s Subjugation of Śiva.” In Religions of India in Practice. Donald S. Lopez (ed.) Princeton: University Press; 547–555. , 2002. Indian Esoteric Buddhism: A Social History of the Tantric Movement. n.y.: Columbia University Press. , 2005. Tibetan Renaissance: Tantric Buddhism in the Rebirth of Tibetan Culture. n.y.: Columbia University Press. , 2011. “Himalayan Buddhist Valleys as Tantric Ecologies.” In Buddhist Himalaya: Studies in Religion, History and Culture, vol. i. Tibet and the Himalaya. Alex McKay and Anna Balikci (eds.) Gangtok: nit; 135–150. Davis, Wade. 2011. Into the Silence. The Great War, Mallory, and the Conquest of Everest. usa: Alfred A. Knopf. Dazey, Wade Hampton. 1990. “Tradition and Modernization in the Organization of the Daśanāmī Saṃnyāsīns.” In Monastic Life in the Christian and Hindu Traditions: A Comparative Study. Austin B. Creel and Vasuadha Narayanam (eds.) Lewiston: The Edwin Mellen Press; 281–321. Deasy, H.H.P. 1901. In Tibet and Chinese Turkestan: being a record of three years’ explo- ration. London. T. Fisher Unwin. Decleer, Hubert. 2000. “Si tu Paṇ chen’s translation of the Svayaṃbhū Purāṇa and his role in the development of the Kathmandhu Valley pilgrimage guide (gnas yig) literature.”Lungta 13; 33–64. bibliography 465

Deeg, Max. 1993. “Shamanism in the Veda: the Keśin–Hymn (10.136), the Journey to Heaven of Vasiṣṭha (r.v.7.88) and the Mahāvrata ritual.” Nagoya Studies in Indian Culture and Buddhism 14; 95–144. Denwood, Philip. 2005. “Early Connections between Ladakh/Baltistan and Amdo/ Kham.” In Ladakhi Histories: Local and Regional Perspectives. John Bray (ed.) Leiden: Brill; 31–40. Desai, Nileshvari. 1968. Ancient Indian Society, Religion and Mythology as Depicted in the Mārkaṇḍeya-Purāṇa (A Critical Study). Baroda: m.s. University of Baroda. Dexter, Miriam Robbins, & Victor H. Mair. 2010. Sacred Display; Divine and Magical Female Figures of Eurasia. n.y.: Cambria Press. Dhami, K.S. 2001. “Expedition to Sri Kailash.” The Himalayan Journal 57: 171–173. Diemberger, Hildegard. 1998. “The Horseman in Red: On Sacred Mountains of La stod lho (Southern Tibet).” In Tibetan Mountain Deities: Their Cults and Representations: Proceedings of the 7th International Seminar for Tibetan Studies, Graz 1995. Anne- Marie Blondeau (ed.) Wien: vöaw; 43–56. Dietz, Siglinda. 1988. “Remarks on four Cosmological Texts from Tun-huang.” In Tibetan Studies: Proceedings of the 4th Seminar of the International Association for Tibetan Studies, Schloss Hohenkammer—Munich 1985, Vol. i. Helga Uebach & Jampa Pan- glung (eds.) Munich: Bayerische Akademie der Wissenschaften; 111–117. Dikshit, K.N. 1984. “Late Harappa in Northern India.” In Frontiers of the Indus Civiliza- tion: Sir Mortimer Wheeler commemoration volume. B.B. Lal & S.P. Gupta (eds.) New Delhi: Books and Books; 253–269. Dimock, E.C., & Ramanujan, A.K. 1964. “The Goddess of Snakes in Medieval Bengali Literature Part 11.”History of Religions 3.2; 300–323. Dirks, Nicholas. 1993. “Colonial Histories and Native Informants: Biography of an Ar- chive.” In Orientalism and the Postcolonial Predicament. Carol A. Breckenridge and Peter van der Veer (eds.) Philadelphia: University of Pennsylvania Press; 279–313. Dodin, Thierry, & Heinz Räther. (eds.) 2001. Imagining Tibet: Perceptions, Projections & Fantasies. Boston: Wisdom Press (first published in German. 1997. Mythos Tibet: Wahrnehmungen, Projektionen, Phantasien. Köln: Dumont). Dodson, Michael. 2007. Orientalism, Empire and National Culture: India, 1770–1880. Lon- don: Palgrave Macmillan. Dollfus, Pascale. 1996. “No sacred mountains in Central Ladakh?” In Reflections of the Mountain: Essays on the History and Social Meaning of the Mountain Cult in Tibet and the Himalaya. Anne-Marie Blondeau & Ernst Steinkellner (eds.) Wien: vöaw; 3–22. Doniger, Wendy. (ed.) 1993. Purāṇa Perennis: Reciprocity and Transformation in Hindu and Jaina Texts. n.y.: suny. Dowman, Keith. 1985. Masters of Mahamudra: Songs and Histories of the Eighty-four Buddhist siddhas. n.y.: suny. , 1988. The Power places of Central Tibet: The Pilgrim’s Guide. London: Routledge and Kegan Paul. 466 bibliography

Dowson, J. 1928. A Classical Dictionary of Hindu Mythology and Religion, Geography, History and Literature. London: Trubner (first published, 1879). Draper, John. 1986. “Nāgas.” Contributions to Nepalese Studies 13.2; 149–166. Driem, George van. 2001. “Zhang-zhung and its next of kin in the Himalayas.” In New Research on Zhangzhung and Related Himalayan Languages. (Bon Studies 3). Yasu- hiko Nagano & Randy J. LaPolla (eds.) Osaka: National Museum of Ethnology; 31– 44. Duchesne-Guillemin, J. 1973. Religion of Ancient Iran. Bombay: Tata Press (first pub- lished, Paris, 1962). Dumont, Louis. 1970. “World Renunciation in Indian Religions.” In Religion, Politics and History in India: Collected Papers in Indian Sociology. Louis Dumont. Paris/Leiden: Mouton (first published in Contributions in Indian Sociology 4. 1960; 33–62). Duncan, Jonathan. 1801. “An Account of Two Fakeers, with their Portraits.” Oriental Memoirs. London; 37–52. Dunlop, D.M. 1973. “Arab Relations with Tibet in the 8th and early 9th centuries a.d.” Islam Tetkikleri Enstitusu Dergisi (Istambul University); 301–318. Dunnell, Ruth W.1996. TheGreatStateofWhiteandHigh:BuddhismandStateFormation in Eleventh-century Xia. Honolulu: University of Hawai’i. Dyczkowski, Mark S.G. 1988. The Canon of the Śaivāgama and the Kubjikā Tantras of the Western Kaula Tradition. n.y.: suny. , 1995–1996. “Kubjika the Erotic Goddess. Sexual Potency, Transformation and Reversal in the Heterodox Theophanies of the Kubjika Tantras.”Indologica Taurinen- sia xxi–xxii; 123–140. , 2001. “The Inner Pilgrimage of the Tantras: The Sacred Geography of the Kubjikā Tantras with Reference to the Bhairava and Kaula Tantras.” Journal of the Nepal Research Centre 12; 43–100. , 2004. A Journey in the World of the Tantras. Varanasi: Indica Books. Eck, Dianna. 1981. “India’s Tīrthas: ‘Crossings’ in Sacred Geography.”History of Religions 20.4; 323–344. , 1985. Darsan: Seeing the Divine Image in India. Chambersburg pa: Anima publications (first published, 1981). , 2012. India: A Sacred Geography. n.y.: Three Rivers Press. Eggermont, P.H.L. 1965–1966. “New Notes on Aśoka and His Successors.” Persica 2; 227–271. Ehrhard, Franz-Karl. 1997. “A ‘Hidden Land’ in the Tibetan-Nepalese Borderlands.” In Maṇḍala and Landscape. A.W. Macdonald (ed.) New Delhi: d.k. Printworld; 335–364. , 1998. “Sa-‘dul dgon-pa: A Temple at the Crossroads of Jumla, Dolpo and Mus- tang.” Ancient Nepal: Journal of the Department of Archaeology 140; 3–20. , 2001. “Review” (of Katia Buffetrille. 2000). Indo Iranian Journal 44.3: 279–285. bibliography 467

, 2003. “Pilgrims in Search of Sacred Lands.” In Sacred Landscape of the Himala- yas. Niels Gutschow, et al. (eds.) Wien. vöaw; 95–110. Eimer, H., & Germano, D. (eds.) 2002. The Many Canons of Tibetan Buddhism: piats 2000: Tibetan Studies: Proceedings of the Ninth Seminar of the International Associa- tion for Tibetan Studies, Leiden 2000. Leiden: Brill. Ekvall, Robert B., and James F. Downs. 1987. Tibetan Pilgrimage. Tokyo: Institute for the Study of Languages and Cultures of Asia and Africa. Elizarenkova, Tatyana. 1995. Language and Style of the Vedic Rsis. n.y.: suny. Elliot, Mark C., 2000. “The Limits of Tartary: Manchuria in Imperial and National Geographies.” The Journal of Asian Studies 59.3; 603–646. Engelhardt, Isrun. 2002. “The Closing of the Gates: Tibetan-European Relations at the End of the Eighteenth Century.” In Tibet, Past and Present. Tibetan Studies 1. piats 2000: Tibetan Studies: Proceedings of the Ninth Seminar of the International Association for Tibetan Studies, Leiden 2000. Henk Blezer (ed.) Leiden: Brill; 229–246. Eno, Robert. 1990. “Was there a High God Ti in Shang Religion?”Early China 15; 1–26. Ensink, Jacob. 1979. “Hindu Pilgrimage and Vedic Sacrifice.” In Ludwik Sternbach Felic- itation Volume (part one). J.P. Sinha (ed.) Lucknow: Akhila Bharatiya Sanskrit Pari- shad; 105–117. Erdosy, G. (ed.) 1995. The Indo-Aryans of Ancient South Asia: Language, Material Culture and Ethnicity. Berlin/n.y.: Walter de Gruyter. Evans-Wentz, W.Y. 1989. Cuchama and Sacred Mountains. (Frank Waters and Charles L. Adams, eds.) Athens: Swallow Press/Ohio University Press (first pub- lished, 1981). , (ed.) 1951. Tibet’s Great Yogi Milarepa: A Biography from the Tibetan being the Jetsün-Kahbum or Biographical History of Jetsün-Milarepa, according to the Late Lāma Kazi Dawa-Samdup’s English rendering edited with introduction and annota- tions by W.Y.Evans-Wentz. London/n.y.: oup (first published in English 1928; original Tibetan text by Gtsaṅ-smyon He-ru-ka [1452–1507]). Falk, Harry. 1989. “Soma i and Soma ii.”bsoas lii.i; 77–90. , 1997. “The Purpose of Rgvedic Ritual.” In Inside the Texts, Beyond the Texts: New Approaches to the Study of the Vedas. Michael Witzel (ed.) Cambridge: Harvard Oriental Series (Opera Minora vol. 2); 69–88. Falk, Nancy. 1973. “Wilderness and Kingship in Ancient South Asia.”History of Religions 13.1; 1–15. Falkenhausen, Lothar von. 1995. “Reflections on the Political Role of Spirit Mediums in Early China: The Wu officials in the Zhou Li.”Early China 20; 279–300. Farquhar, J.N. 1925. “The Fighting Ascetics of India.”Bulletin of the John Rylands Library 9.2; 431–452. Faure, Bernard. 2001. “Japanese Tantra, the Tachikawa-ryu and Ryobu Shinto.” In Tantra in Practice. David Gordon White (ed.) Delhi: Motilal Banarsidass; 543–556. 468 bibliography

Feldhaus, A. 1990. “The Image of the Forest in the Mahatmyas of the Rivers of the Deccan.” In The History of the Sacred Places of India as Reflected in the Traditional Literature [Panels of the viith World Sanskrit Conference, Leiden, 1987, vol. iii]. Hans Bakker (ed.) Leiden: Brill; 93–99. Filibeck, Elena De Rossi. 1988. Two Tibetan Guide Books to Ti se and La phyi. Bonn: vgh Wissenschaftsverlag. Finn, L.M. (trans.) 1986. The Kulacūḍāmaṇi Tantra and the Vāmakeśvara Tantra with the Jayaratha Commentary. Wiesbaden: Otto Harrassowitz. Flattery, David S., & Martin Schwartz. 1989. Haoma and Harmaline: The Botanical Identity of the Indo-Iranian Sacred Hallucinogen “Soma” and its Legacy in Religion, Language,andMiddleEasternFolklore. Berkeley: University of California Press (Near Eastern Studies, vol. 21). Fleming, Peter. 1984. Bayonets to Lhasa. Oxford: oup (first published, 1961). Flugel, Peter. 2010. “The Jaina Cult of Relic Stūpas.” Numen 57.3–4; 389–504. Francke, A.H. 1907. A History of Western Tibet: one of the unknown empires. London: Partridge & co. , 1977. Antiquities of Indian Tibet. 2 vols. Delhi: Archaeological Survey of India (first published, 1914). Fraser, James. 1820. Journal of a Tour through part of the snowy range of the Himālā Mountains and to the sources of the rivers Jumna and Ganges. London: Rodwell and Martin. French, Patrick. 1994. Younghusband: The Last Great Imperial Adventurer. London: HarperCollins. Gamble, Ruth. 2010. “Laughing Vajra: the outcast clown, satirical guru and smiling Buddha in Milarepa’s songs.” In New Views of Tibetan Culture. David Templeman (ed.) Caulfield: Monash University Press; 137–166. Ganhar, J.N. 1975. Jammu Shrines and Pilgrimages, New Delhi: Ganhar. Gerard, Alexander. 1993. Account of Koonawur in the Himalaya. (George Lloyd; ed.) New Delhi: Indus (first published, 1841). , 1996. Tours in the Himalaya. Account of an attempt to penetrate by Bekhur to Garoo, and the Lake Manasarowara. New Delhi: Indus (first published, 1840). Germano, David. 2002. “The Seven Descents and the Early History of the rNying ma Transmissions.” In The Many Canons of Tibetan Buddhism: piats 2000: Tibetan Stud- ies: Proceedings of the Ninth Seminar of the International Association for Tibetan Stud- ies, Leiden 2000. Helmut Eimer & David Germano (eds.) Leiden: Brill, 225–263. Gernet, Jaques. 1995. Buddhism in Chinese Society: An Economic History from the Fifth to the Tenth Centuries. n.y.: Columbia University Press, (first published in French, 1956). Getty, A. 1914. TheGodsofNorthernBuddhism,theirhistory,iconography,andprogressive evolution through the northern Buddhist countries. Oxford: Clarendon Press. bibliography 469

Ghosh, J.M. 1930. Sannyāsi and Fakir Raiders in Bengal. Calcutta: Bengal Secretariat Book Depot. Ghurye, G.S. 1964. Indian Saddhus. Bombay: Popular Pradakshan (first published, 1953). , 1972. Two Brahmanical Institutions Gotra and Charana. Bombay: Popular Pra- kashan. Gingrich, Andre. 1996. “Hierarchical merging and horizontal distinction—A compara- tive perspective on Tibetan mountain cults.” In Reflections of the Mountain: Essays on the History and Social Meaning of the Mountain Cult in Tibet and the Himalaya. Anne-Marie Blondeau and Ernst Steinkellner (eds.) Wien: vöaw; 237–239. Giri, Swami Bikash. 2000. Sumeru Parvat: 12 Years of Kailas Manasavoar Pilgrimage and Transformation. Pithoragarh: Kailash Ashram (first Hindi publication, 1999). Giri, Swami Sadānanda. 1976. Society and Sannyāsin (A History of the Daśnāmī San- nyāsins). Varanasi: Sannyāsi Sanskrit Mahāvidyālaya. Glover, M.B. 1930. “Around the Kanawar Kailas.”Himalayan Journal ii; 81–86. Goetz, Herman. 1969. Studies in the History and Art of Kashmir and the Indian Himalaya. Wiesbaden: Otto Harrassowitz. , 1955. Early Wooden Temples of Chamba. Leiden: Brill. Gombrich, Richard F. 1975. “Ancient Indian Cosmologies.” In Ancient Cosmologies. Carmen Blacker and Michael Loewe (eds.) London: Allen and Unwin; 110–142. , 1985. “The Vessantara Jātaka, the Rāmāyaṇa, and the Dasaratha Jātaka.” jaos 105.3; 427–437. , 1992. “Dating the Buddha: A Red Herring Revealed.” In The Dating of the Historical Buddha, 2 vols. Heinz Bechert (ed.) Göttingen: Vandenhoeck & Ruprecht; vol. ii, 237–259. , 2006. Theravada Buddhism. London: Routledge (first published, 1988). Gonda, Jan. 1975. Vedic Literature, vol. 1: Saṃhitās and Brāhmaṇas,[A History of Indian Literature]. Wiesbaden: Otto Harrassowitz. , 1977. Medieval Religious Literature in Sanskrit. Weisbaden: Otto Harrasowitz. Gorvine, William M. 2006. “The Life of a Bonpo Luminary: Sainthood, Partisanship and Literary Representation in a 20th Century Tibetan Biography.” University of Virginia Ph.D. thesis. Goss, Robert. 1993. “The hermeneutics of madness: A literary and hermeneutical anal- ysis of the ‘Mi-la’i-rnam-thar’ by gTsang-smyon Heruka.” Harvard University Ph.D. thesis. Goudriaan, Teun & Sanjukta Gupta. 1981. Hindu Tantic and Śākta Literature. Wies- baden: Otto Harrassowitz. Goudriaan, Teun and J.A. Schoterman. 1988. The Kubjikāmatatantra; Kulālikāmnāya version. Critical Edition. Leiden: Brill. , (eds./trans.) 1994. The Kubjuka Upanisad. Groningen: Egert Forsten. Govinda, Lama Anagorika. 1984. The Way of the White Clouds. A Buddhist Pilgrim in Tibet. London: Rider & co (first published, 1966). 470 bibliography

, 1991. Insights of a Himalayan Pilgrim. Oakland: Dharma publishing. Granoff, Phyllis. 1997. “Heaven on Earth: Temples and Temple Cities of Medieval India.” In India and Beyond: Aspects of Literature, Meaning, Ritual and Thought. Essays in Honour of Frits Staal. Dick van der Meij (ed.) London: Kegan Paul International; 170–193. , 2006. “Mountains of Eternity: Raidhū and the Colossal Jinas of Gwalior.”Rivista di Studi Sudasiatici, 1.1: 31–50. Gray, David B. 2001. “On Supreme Bliss. A Study of the History and Interpretation of the Cakrasaṃvara Tantra.” Columbia University Ph.D. thesis. , 2005. “Eating the Heart of the Brahmin: Representations of Alterity and the Formation of Identity in Tantric Buddhist Discourse.” History of Religions 45.1: 45– 69. , 2006. “Mandala of the Self: Embodiment, Practice, and Identity Construction in the Cakrasamvara Tradition.” Journal of Religious History 30.3; 294–310. , 2007. The Cakrasamvara Tantra (The Discourse of Śrī Heruka) Śrīherukābhid- hāna: A Study and Annotated Translation. n.y.: The American Institute of Buddhist Studies at Columbia University. . 2007A. “The Cakrasamvara Tantra: Its History, Interpretation, and Practice in India and Tibet.”Religion Compass 1.6: 695–710. Griffiths, Arlo. 2003. “The Textual Divisions of the Paippalāda Saṃhitā.” wzks xlvii: 5–37. Griffiths Arlo, and Annette Schmiedchen (eds.) 2007. The Atharvaveda and its Pippalā- daśākhā: Historical and Philological Papers on a Vedic Tradition. Aachen: Shaker Ver- lag. Gross, Robert Lewis. 1992. The Sadhus of India: A Study of Hindu Asceticism. Jaipur: Rawat. Grünendahl, R. 1993. “Zu den beiden Gandhamādana-Episoden des Āranyakaparvan.” Studien zur Indologie und Iranistik 18; 103–138. Gupta, Linda. 2002. “Tantric Incantation in the Devī Purāna: Padamālā Mantra Vidyā.” In The Roots of Tantra. Katherine A. Harper & Robert L. Brown (eds.) n.y.: suny; 231–249. Gupta, P.L. 1988. Numismatic History of Himachal Pradesh. Delhi: b.r. Publishing. Gupta, Sanjukta, and Richard Gombrich. 1986. “Kings, Power and the Goddess.”South Asia Research 6.2; 123–138. Gurung, Kalsang Norbu. 2011. “History and Antiquity of the Mdo ‘dus in Relation to the Mdo Chen po Bzhi.” In Emerging Bon: The Formation of Bon Traditions in Tibet at the Turn of the First Millennium ad. Henk Blezer (ed.) Halle: iitbs. Gutschow, Niels, Axel Michaels, Charles Ramble, and Ernst Steinkellner. (eds.) 2003. Sacred Landscape of the Himalayas: proceedings of an international conference at Heidelberg, 25–27 May 1998. Wien: vöaw. bibliography 471

Haarh, Erik. 1969. The Yar-luṅ Dynasty: a study with particular regard to the contribution of myths and legends to the history of Ancient Tibet and the origin and nature of its kings. Copenhagen: g.e.c. Gad. , 2003. “Extract from The Yar-luṅ Dynasty.” In The History of Tibet: The Early Period; to c. ad 850 The Yarlung Dynasty. Alex McKay (ed.) London: RoutledgeCur- zon; 142–155. Habibullah, A.B.M. 1976. “An Early Arab Report on Indian Religious Sects.” In History and Society: Essays in Honour of Professor Niharranjan Ray. Debiprasad Chattopad- hyaya (ed.) Calcutta: K.P. Bagchi; 433–444. Halbfass, Wilhelm. (ed.) 1995. Philology and Confrontation: Paul Hacker on Traditional and Modern Vedanta. n.y.: suny. Halkias, Georgios T. 2009. “Until the Feathers of the Winged Black Raven Turn White: Sources for the Tibet-Bashahr Treaty of 1679.” In Mountains, Monasteries and Mosques: Recent Research on Ladakh and the Western Himalayas. Proceedings of the 13th Colloquium of the International Association for Ladakh Studies. John Bray & E. Elana de Rossi Filibeck (eds.) Pisa/Rome: Fabrizio Serra Editore; 61–86. Hamsa, Bhagwān Shri. 1997. The Holy Mountain: being the story of a pilgrimage to Lake Mānas and of initiation on Mount Kailās in Tibet. Delhi: Pilgrim Books (first published, London, 1934). Handa, Devendra. 2002. “History and Coinage of the Kunindas.” In Where Mortals and Mountain Gods Meet: Society and Culture in Himachal Pradesh. Laxman Thakur (ed.) Shimla: iias; 193–208. Handa, O.C. 2002. History of Uttaranchal. Delhi: Indus. , 2004. Naga Cults and Traditions in the Western Himalaya. Delhi: Indus. Hansen, Valerie. 1990. Changing Gods in Medieval China, 1127–1276. Princeton: Univer- sity Press. Hara, Minoru. 1994. “Pāśupata Studies ii.” wkzs. xxxviii; 323–335. Hardwicke, Thomas. 1801. “Narrative of a Journey to Sirinagur.” Asiatik Researches 6; 309–347. Harley, B., and D. Woodward. (eds.) 1987. The History of Cartography, Volume 2, Book 1: Cartography in the Traditional Islamic and South Asian Societies. Chicago/London: University of Chicago Press. Harris, Ian. 2006. Cambodian Buddhism: History and Practice. Chiang Mai: Silkworm books (first published, Hawai’i, 2005). Harrist, Robert E. 2008. The Landscape of Words: Stone Inscriptions from Early and Medieval China. Seattle /London: University of Washington Press. Hartsuiker, Dolf. 1993. Sādhus: Holy Men of India. London: Thames and Hudson. Hartzell, James F. 1997. “Tantric Yoga: A Study of the Vedic Precursors, Historical Evo- lution, Cultures, Doctrines and Practices of the 11th Century Kaśmīri Śaivite and Buddhist Unexcelled Tantric Yogas.” Columbia University Ph.D. thesis. 472 bibliography

Hatto, A.T. 1980. “Kirghiz. Mid-nineteenth century.” In Traditions of Heroic and Epic Poetry.VolumeOne:TheTraditions. A.T. Hatto (ed.) London: The Modern Humanities Research Association; 300–327. , (ed./trans.) 1990. The Manas of Wilhelm Radloff. Wiesbaden: Otto Harras- sowitz. Hausner, Sondra L. 2007. Wandering with Sadhus: Ascetics in the Hindu Himalayas. Bloomington: Indiana University Press. Havell, E.B. 1924. The Himalayas in Indian Art. London: John Murray. Hazra, R.C. 1975. Studies in the Purāṇic Records on Hindu Rites and Customs. Delhi: Motilal Banarsidass (first published, Dacca, 1940). Hedin, Sven. 1909–1913. Transhimalaya; discoveries and adventures in Tibet. London: Macmillan, 3 vols. , 1917–1922. Southern Tibet; discoveries in former times compared with my own researches in 1906–1908, 9 vols. Stockholm: Lithographic Institute of the General Staff of the Swedish Amy. Heesterman, J.C. 1962. “Vratya and Sacrifice.”Indo-Iranian Journal vi.i: 1–37. , 1984. ‘“Orthodox’ and ‘Heterodox’ Law: Some Remarks on Customary Law and the State.” In Orthodoxy, Heterodoxy and Dissent in India. S.N. Eisenstadt, R. Kahane, & D. Shulman (eds.) Amsterdam/Berlin/n.y.: Mouton; 149–167. , 1985. The Inner Conflict of Tradition: Essays in Indian Ritual, Kingship and Society. Chicago: University of Chicago Press. , 1988. “Householder and wanderer.” In Way of Life: King, Householder, Re- nouncer: Studies in honour of Louis Dumont. T.N. Madan (ed.) Delhi: Motilal Banar- sidass (first published, 1982); 251–271. , 1991. “Hinduism and Vedic Ritual.”History of Religions 30.3; 296–305. , 1993. The Broken World of Sacrifice: An Essay in Ancient Indian Ritual. Chicago/ London: University of Chicago Press. , 1995. “Warrior, Peasant and Brahmin.” Modern Asian Studies 29.3; 637–654. , 1997. “Ritual and Ritualism: The Case of Ancient Indian Ancestor Worship”. In India and Beyond: Aspects of Literature, Meaning, Ritual and Thought. Essays in Honour of Frits Staal. Dick van der Meij (ed.) London: Kegan Paul International: 249–270. Heim, A. & A. Gansser. 1939. “Central Himalaya; geological observations of the Swiss expedition 1936.” Memoir Society Helvetica Science Nature, 73. , 2004. The Throne of the Gods: An Account of the First Swiss Expedition to the Himalayas. Bangkok: White Lotus (first published, n.y., 1939). Hein, Veronika. 2007. “A Preliminary Analysis of Some Songs in Tibetan Language Recorded in Spiti and Upper Kinnaur.” In Text, Image and Song in Transdisciplinary Dialogue. piats 2003: Tibetan Studies: Proceedings of the Tenth Seminar of the Inter- national Association for Tibetan Studies, Oxford, 2003. Deborah Klimburg-Salter, Kurt Tropper and Christian Jahoda (eds.) Leiden: Brill; 235–247. bibliography 473

Heller, Amy. 2005. “A Thang ka Portrait of ‘Bri gung rin chen dpal, ‘Jig rten gsum mgon (1143–1217).” Journal of the International Association of Tibetan Studies 1: 1–10. Henss, M. 1997. “Wall-Paintings in Western Tibet: The Art of the Ancient Kingdom of Guge, 1000–1500.” Marg xlviii.i; 33–61. Herbert, J.D. 1825. “An Account of a Tour made to lay down the Course and Levels of the River Setlej or Satúdrá, as far as traceable within the limits of British authority, performed in 1819.”Asiatic Researches xv: 339–428. Hill, Nathan. 2006. “The Old Tibetan Chronicle Chapter 1.”Revue d’Etudes Tibétaines 10; 89–101. Hiltebeitel, Alf. 2004. “More Rethinking the Mahābhārata: Toward a Politics of Bhakti.” Indo-Iranian Journal 47.3–4; 203–227. Hilton, James. 1933. Lost Horizon. London: Macmillan. Hitchcock, John T. 1978. “An Additional Perspective on the Caste System.” In Himalayan Anthropology: The Indo-Tibetan Interface. James F. Fisher (ed.) The Hague/Paris: Mouton; 111–120. Hodgson, J.A. 1822. “Journal of a Survey to Heads of the Rivers, Ganges and Jumna.” Asiatik Researches 14; 60–152. Hodgson, J.A., and J.D. Herbert. 1822. “An account of Trigonometrical and Astronomical Operations for determining the Heights and Positions of the Principal Peaks of the Himálaya Mountains.” Asiatik Researches 14; 187–372. Hoek, A.W. van den, D.H.A. Kolff, & M.S. Oort. (eds.) 1992. Ritual, State and History in South Asia: Essays in Honour of J.C. Heesterman. Leiden: Brill. Hoey, W. 1907. “The Five Rivers of the Buddhists.” jras 38; 41–46. Hofinger, Marcel. 1982. Le Congrès du Lac Anavatapta (Vies de Saints Bouddhiques). Louvain: Institut Orientaliste Louvain-la-Neuve (first published, 1954). Holmberg, David. 2011. “From Naturalization to Rationalization in Western Tamang Buddhist Practice.” In Buddhist Himalaya: Studies in Religion, History and Culture, vol.1,TibetandtheHimalaya. Alex McKay & Anna Balikci-Denjongpa (eds.) Gangtok: nit; 186–187. Holt, John Clifford. 2009. Spirits of the Place: Buddhism and Lao Religious Culture. Honolulu: University of Hawai’i Press. Huber, Toni. 1994. “What You See Is Not What You Get: Remarks on the Traditional Tibetan Presentation of Sacred Geography.” In Tantra and Popular Religion in Tibet. Geoffrey Samuel, H. Gregor & E. Stutchbury (eds.) New Delhi: Aditya Prakashan; 39–52. , 1997. “Colonial Archaeology, International Missionary Buddhism and the First Example of Modern Tibetan Literature.” In Bauddhavidyāsudhākaraḥ: Studies in Honour of Heinz Bechert On the Occasion of His 65th Birthday. (Indica et Tibetica 30). Petra Kieffer-Pülz and Jens-Uwe Hartmann (eds.) Swisttal-Odendorf; 297–318. , 1997a. “Ritual and Politics in the Eastern Himalaya: The Staging of Processions 474 bibliography

at Tsa-ri.” In Les Habitants du Toit Du Monde: Études Recueillies en Hommage Àlexan- der W. Macdonald. Samten G. Karmay & Phillipe Sagant, (eds.) Nanterre: Société d’ethnologie; 221–260. , 1999. The Cult of Pure Crystal Mountain: Popular Pilgrimage and Visionary Landscape in Southeast Tibet. n.y./Oxford: oup. , 1999a. “Putting the Gnas Back into Gnas-skor: Rethinking Tibetan Pilgrimage Practice.” In Sacred Spaces and Powerful Places in Tibetan Culture: A Collection of Essays. Toni Huber (ed.) Dharamsala: ltwa; 77–104 (first published, 1994. The Tibet Journal 19.2; 23–60). , (ed.) 1999b. Sacred Spaces and Powerful Places in Tibetan Culture: A Collection of Essays. Dharamsala: ltwa. , (ed/trans.) 2000. The Guide to India: A Tibetan Account by Amdo Gendun Chöphel. Dharamsala: lwta. , 2001. “Review” (of dBa’ bzhed. The Royalnarrative concerning the bringing of the Buddha’s Doctrine to Tibet. Pasang Wangdu and Hildegard Diemberger [eds.]) Acta Orientalia 62; 280–286. , 2003. “Where Exactly are Cāritra, Devīkoṭa and Himavat? A Sacred Geog- raphy Controversy and the Development of Tantric Buddhist Pilgrimage Sites in Tibet.” In The History of Tibet vol. ii. The Medieval Period. The Development of Bud- dhist Paramountcy. Alex McKay (ed.) London: RoutledgeCurzon; 392–424 (revised version; first published, 1990. Kailash 16.3–4; 121–164). , 2008. The Holy Land Reborn: Pilgrimage and the Tibetan Reinvention of Bud- dhist India. Chicago: University Press. , 2011. “Pushing South. Tibetan Economic and Political Activities in the Far Eastern Himalaya, ca., 1900–1950.” In Buddhist Himalaya: Studies in Religion, History and Culture. Vol. 1: Tibet and the Himalaya. Alex McKay & Anna Balikci-Denjongpa (eds.) Gangtok: nit; 259–276. Huber, Toni, and Tsepak Rigzin. 1999. “A Tibetan Guide to Pilgrimage to Ti-se (Mount Kailas) and mTsho Ma-pham (Lake Manasarovar).” In Sacred Spaces and Powerful Places in Tibetan Culture: A Collection of Essays. Toni Huber (ed.) Dharamsala: ltwa; 125–153 (first published, 1995. The Tibet Journal 20.1; 10–47). Humphrey, Caroline. 1995. “Chiefly and Shamanist Landscapes in Mongolia.” In The Anthropology of Landscape: Perceptions on Place and Space. Eric Hirsch and Michael O’Hanlon (eds.) Oxford: Clarendon Press; 135–162. Hutchinson, J., & J.Ph. Vogel. 2000. History of the Panjab Hill States. 2 vols. Shimla: h.p. Department of Language and Culture (first published, 1933). Imaeda, Yoshiro, Mathew T. Kapstein and Tsuguhito Takeuchi. (eds.) 2011. New Studies of the Old Tibetan Documents: Philology, History and Religion. Tokyo: Tokyo Univer- sity of Foreign Studies. Inden, Ronald. 1985. “The Temple and the Hindu Chain of Being.” In L’Espace du bibliography 475

temple: Espaces, itineraries, meditations. Jean-Claude Galey (ed.) Paris: Collection Purusartha 8; 53–73. , 1992. “Changes in the Vedic Priesthood.” In Ritual, State and History in South Asia: Essays in Honour of J.C. Heesterman. A.W. van den Hoek, D.H.A. Kolff, & M.S. Oort (eds.) Leiden: Brill; 556–577. , 2000. “Imperial Puranas. Kashmir as Vaisnava Center of the World.” In Query- ing the Medieval: Text and the History of Practices in South Asia. Ronald Inden, Jonathan Walters and Daud Ali. n.y.: oup; 29–98. Ingalls, Daniel H.H. 1962. “Cynics and Pāśupatas: The Seeking of Dishonour.” Harvard Theological Review lv.iv; 281–298. Insler, Stanley. 1998. “On the Recensions of the Atharva Veda and the Atharvan Hymn Composition.”wzks 42; 5–21. Irwin, John. 1983. “The Ancient Pillar-Cult at Prayāga (Allahabad): Its Pre-Aśokan Ori- gins.” jras (new series) 2; 253–280. Jackson, David P. 1976. “The Early History of Lo (Mustang) and Ngari.” Contributions to Nepalese Studies 4.1; 39–56. , 1978. “Notes on the History of Se-Rib, and Nearby Places in the Upper Kali Gandaki Valley.”Kailash 6.3; 195–227. , 1984. The Mollas of Mustang. New Delhi: ltwa. , 1994. Enlightenment by a Single Means: Tibetan Controversies on the “Self- Sufficient White Remedy” (dkar po chig thub). Wien: vöaw. Jain, J.C. 1991. The Jain Way of Life. Delhi: Manohar. Jay, J.W. 1996. “Imagining Matriarchy: ‘Kingdoms of Women’ in Tang China.” jaos 116.2; 220–229. Jeffares, A. Norman. 1984. A New Commentary on the Poems of W.B. Yeats. Stanford: University Press. Jerath, Ashok. 2001. Hindu Shrines of the Western Himalayas. Jammu: Association of Litterateurs, Folklorists and Artists. Jha, D.N. 2000. “State Formation in a Peripheral Region: The Case of Early Medieval Chamba.” In The Feudal Order: State, Society and Ideology in Early Medieval India. D.N. Jha (ed.) Delhi: Manohar; 197–209, (first published in Society and Ideology in India: Essays in Honour of Professor R.S. Sharma. D.N. Jha (ed.) Delhi, 1996). Jha, Makhan. (ed.) 1985. Dimensions of Pilgrimage an anthropological appraisal: based on the transactions of a World Symposium on Pilgrimage. New Delhi: Inter-India. Johnson, K. Paul. 1994. The Masters Revealed: Madame Blavatsky and the Myth of the Great White Lodge. n.y.: suny. Johnson, Russell, and Kerry Moran. 1989. Kailas: On Pilgrimage to the Sacred Mountain of Tibet. London: Thames & Hudson. Jones, Kenneth W. 1976. Arya Dharm: Hindu Consciousness in 19th-century Punjab. Berkerley: University of California Press. 476 bibliography

, 1989. Socio-Religious Reform Movements in British India (The New Cambridge History of India). Cambridge; cup. Jong, Albert de. 1997. Traditions of the Magi. Zoroastrianism in Greek and Latin Litera- ture. Leiden: Brill. Jong, J.W. de. 1989. The Story of Rama in Tibet. Stuttgart: University of Hamburg. , 1993. “The Beginnings of Buddhism.”The Eastern Buddhist 26.2; 1–18. Joshi, J.P., M. Bala & J. Ram. 1984. “The Indus Civilization: A Reconsideration of the Basis of Distribution Maps.” In Frontiers of the Indus Civilization: Sir Mortimer Wheeler commemoration volume. B.B. Lal & S.P. Gupta (eds.) New Delhi: Books and Books; 511–530. Joshi, Maheshwar P. 1986. “The Religious History of Uttarakhand: Sources and Materi- als.” In Ecology, Economy and Religion of Himalayas. L. Vidyarthi & M. Jha (eds.) New Delhi: Orient; 193–216. , 1988. “Kumaun: Archaeological and Historical Perspective.” In Kumaun; Land and People. A Volume in Honour of Shri Narayan Datt Tiwari. K.S. Valdiya (ed.) Nainital: Gyanodaya Prakashan; 73–86. , 1989. Morphogenesis of Kunindas (cir. 200b.c.–cir. a.d. 300). A Numismatic Overview. Almora: Shree Almora Books. , 2009. “Advent of Polities in Uttarakhand (Kumaon and Garwhal).” In Bardsand Mediums: History, Culture and Politics in the Central Himalayan Kingdoms. Lecomte- Tilouine, Marie (ed.) Almora: Almora Book Depot; 327–371. Joshi, Maheshwar P. & C.W. Brown. 1987. “Some Dynamics of Indo-Tibetan Trade through Uttarkhandda (Kumaon-Garwhal), India.” Journal of the Economic and Social History of the Orient 30; 303–317. Kaelber, Walter O. 1989. Tapta Marga: asceticism and initiation in vedic India. n.y.: suny. Kane, P.V.1953. A History of Dharmaśāstra: Ancient and Medieval Religious and Civil Law in India, vol. iv, Poona: Bhandarkar Oriental Research Institute. Kapadia, Harish. 1988–1989. “A Note on Kinnaur.” The Alpine Journal 93; 29–34. , 1989–1990. “Two is Company: Explorations and Climbs in the Mana Gad.” Himalayan Journal 47; 42–55. Kapstein, Mathew T. 1980. “The Shangs-pa bka’-brgyud: An Unknown Traditon of Tibetan Buddhism.” In Tibetan Studies in Honour of Hugh Richardson: Proceedings of the International Seminar of Tibetan Studies, Oxford 1979. Michael Aris and Aung San Suu Kyi (eds.) Warminster: Aris & Phillips; 138–144. , 2000. The Tibetan Assimilation of Buddhism: Conversion, Contestation, and Memory. n.y./Oxford: oup. , 2003. “Remarks on the Maṇi bKa’-‘bum and the Cult of Avalokiteśvara in Tibet.” In The History of Tibet. Vol. ii, The Medieval Period: c. 850–1895. The Develop- ment of Buddhist Paramountcy. Alex McKay (ed.) London: RoutledgeCurzon; 523– bibliography 477

537 (first published in Tibetan Buddhism, Reason and Revelation. S.M. Goodman & R.M. Davidson [eds.] Albany: suny; 79–93). , 2006. “New Light on an Old Friend: pt 849 Reconsidered.” In Tibetan Buddhist Literature and Praxis: Studies in its Formative Period 900–1400, piats 2003: Tibetan Studies: Proceedings of the Tenth Seminar of the International Association for Tibetan Studies, Oxford, 2003. Ronald M. Davidson and Christian Wedemeyer (eds.) Leiden: Brill; 9–30. , 2011. “Just Where on Jambudvīpa Are We? New Geographical Knowledge and Old Cosmological Schemes in Eighteenth-century Tibet.” In Forms of Knowledge in Early Modern Asia: Explorations in the Intellectual History of India and Tibet, 1500– 1800. Sheldon Pollock (ed.) Durham and London: Duke University Press; 336–364. Karambelkar, W.W. 1959. The Atharvavedic civilization: its place in the Indo-Aryan cul- ture. Nagpur: University Press. Karmay, Samten G. 1972. The treasury of good sayings: A Tibetan history of Bon. London: oup. , 1975. A General Introduction to the History and Doctrines of Bon. Memoirs of the Research Department of the Toyo Bunko, No. 33, Tokyo. (A revised version of this was published in The Arrow and the Spindle: Studies in History, Myths, Rituals and Beliefs in Tibet. Samten Karmay, 1998. Kathmandhu: Mandala Book Point; 104– 156). , 1983. “Early Evidence for the Existence of Bon as a Religion in the Royal Period.” In Contributions on Tibetan and Buddhist religion and philosophy, vol. 2. Ernst Steinkellner & Helmut Tauscher (eds.) Wien: atbs; 89–106 (reprinted in Sam- ten Karmay. 1998. Op. cit.; 157–168). , 1992. “Mount Bon ri and its associations with early myths.” In Tibetan Studies: Proceedings of the 5th Seminar of the International Association for Tibetan Studies, Narita 1989. S. Ihara & Z. Yamaguchi (eds.), Narita: Naritasan Shinshoji; vol. 2; 527– 539 (reprinted in Samten Karmay. 1998. Op cit.; 211–227). , 1994. “Mountain Cults and National Identity in Tibet.” In Resistance and Reform in Tibet. Robert Barnett and Shirin Akiner (eds.) London: Hurst Books; 112–120 (reprinted in Samten Karmay. 1998. Op cit.; 423–431). , 1996. “The Cult of Mount Murdo in Gyalrong.” Kailash xviii.i–ii; 1–16 (re- printed in Samten Karmay. 1998. Op cit.; 451–464). , 1996a. “The Tibetan cult of mountain deities and its political significance.” In Reflections of the Mountain: Essays on the History and Social Meaning of the Mountain Cult in Tibet and the Himalaya. Anne-Marie Blondeau & Ernst Steinkellner (eds.) Wien: vöaw; 59–76 (reprinted in Samten Karmay. 1998. Op cit.; 432–450). , 1998.The Arrow and the Spindle: Studies in History, Myths, Rituals and Beliefs in Tibet. Kathmandhu: Mandala Book Point. , 2000. “A comparative study of the yul lha cult in two areas and its cosmological 478 bibliography

aspects.” In New Horizons in Bon Studies (Bon Studies 2). Samten G. Karmay & Yasuhiko Nagano (eds.) Osaka: National Museum of Ethnology; 383–416. , 2003. “The Ordinance of Lha Bla-ma Ye-Shes-Od.” In The History of Tibet vol. ii. The Medieval Period. The Development of Buddhist Paramountcy. Alex McKay (ed.) London: RoutledgeCurzon; 138–139 (first published in 1980. Tibetan Studies in Honour of Hugh Richardson. Michael Aris & Aung San Suu Kyi (eds.) Warminster: Aris & Philips; 150–162; first reprinted in Karmay 1998. Op. cit.; 3–16). , 2007. The Great Perfection (rDzogs chen) A Philosophical and Meditative Teach- ing of Tibetan Buddhism. Leiden: Brill. Karmay, Samten, and Philippe Sagant (eds.) 1997. Les Habitants du Toit du Monde: Études Recueillies en Hommage à Alexander W. Macdonald. Nanterre: Société d’Eth- nologie. Karmay, Samten G., & Yasuhiko Nagano (eds.) 2000. New Horizons in Bon Studies (Bon Studies 2). Osaka: National Museum of Ethnology. , (eds.) 2002. The Call of the Blue Cuckoo: An Anthology of Nine Bonpo Texts on Myths and Rituals. Osaka: National Museum of Ethnology. , (eds.) 2003. A Survey of Bonpo Monasteries and Temples in Tibet and the Hima- laya. Bon Studies 7. Osaka: National Museum of Ethnology. Kaschewsky, Rudolf. 2001. “The Image of Tibet in the West before the Nineteenth Century.” In Imagining Tibet: Perceptions, Projections & Fantasies. Thierry Dodin & Heinz Räther (eds.) Boston: Wisdom Press; 3–20. Katz, R.C. 1989. Arjuna in the Mahabharata: Where Krishna Is, There Is Victory. Colum- bia: University of South Carolina. Kawaguchi, Ekai. 1979. Three Years in Tibet. Kathmandhu: Ratna Pustak, (first pub- lished, Madras 1909). Keightley, David N. 1978. “The Religious Commitment: Shang Theology and the Genesis of Chinese Political Culture.”History of Religions 17.3; 211–225. Kerr, Ian J. 2001. Railways in modern India. New Delhi/n.y.: Oxford University Press. Khazanov, Anatoly Michailovich & André Wink (eds.) 2007. Nomads in the Sedentary World. Abingdon: Routledge-iias (first published, 2001). Kind, Marietta. 2007. “Jag ‘dul—A Bon Mountain Pilgrimage.” In Discoveries in Western Tibet and the Western Himalayas: Essays on History, Literature, Archaeology and Art. piats 2003: Tibetan Studies: Proceedings of the Tenth Seminar of the International Association for Tibetan Studies, Oxford, 2003. Amy Heller and Giacomella Orofino (eds.) Leiden: Brill; 199–214. , 2012. The Bon Landscape of Dolpo: Pilgrimages, Monasteries, Biographies and the Emergence of Bon. (Worlds of South and Inner Asia: vol. 4). Bern, etc: Peter Lang. King, Richard. 1999a. Orientalism and Religion: Postcolonial Theory, India and “The Mystic East.” London/ n.y.: Routledge. , 1999b. “Orientalism and the Modern Myth of ‘Hinduism.”’ Numen 46.2; 146– 183. bibliography 479

Kirkland, J. Russell. 1982. “The Spirit of the Mountain: Myth and State in Pre-Buddhist Tibet.”History of Religions 21.3; 257–271. Kish, George. 1984. To the Heart of Asia: The Life of Sven Hedin. Ann Arbor: University of Michigan Press. Kleeman, Terry F. 1994. “Mountain Deities in China: The Domestication of the Moun- tain God and the Subjugation of the Margins.” jaos 114.2; 226–238. , 1994. “Licentious Cults and Bloody Victuals: Sacrifice, Reciprocity, and Vio- lence in Traditional China.” Asia Major 3.7.1; 185–211. Klimburg-Salter, Deborah. (ed.) 1982. The Silk Road and the Diamond Path: Esoteric Buddhist Art on the Trans-Himalayan Trade Routes, l.a.: ucla Arts Council. , (ed.) 1997. Tabo a Lamp for the Kingdom: Early Indo-Tibetan Buddhist Art in the Western Himalayas. Milan: Skira. , 2002. “Ribba, The Story of an Early Buddhist Temple.” In Buddhist Art and TibetanPatronage:NinthtoFourteenthCenturies,piats2000:ProceedingsoftheNinth Seminar of the International Association for Tibetan Studies, Leiden 2000. Deborah Klimburg Salter & Eva Allinger (eds.) Leiden: Brill; 1–28. , 2007. “Kha-che lugs and The Wooden Sculptures from Charang.” In ThePandita and the Siddha: Tibetan Studies in Honour of E. Gene Smith. Ramon N. Prats (ed.) McLeod Gang: Amnye Machen Insitute; 138–152. Klokke, Marijke J. 1995. “On the Orientation of Ancient Javanese Temples: The Example of Candi Surowono.” In International Institute for Asian Studies, Yearbook 1994. Paul van der Velde (ed.) Leiden: ii as; 73–86. Kolff, Dirk. H.A. 1971. “Sannyasi Trader-Soldiers.”ieshr 8; 213–218. Kooij, K.R. van. 1972. Worship of the Goddess according to the Kālikāpurāṇa: part i. A Translation with an Introduction and notes of Chapters 54–69. Leiden: Brill. Koppedrayer, K.I. 2003. “The Interweave of Place, Space, and Biographical Discourse at a South Indian Religious Centre.” In Pilgrims, Patrons, and Place: Localizing Sanc- tity in Asian Religions. Phyllis Granoff and Koichi Shinohara (eds.) Vancouver: The University of British Columbia Press; 279–296. Korom, Frank J. 1992. “Of Navels and Mountains: A Further Inquiry into the History of an Idea.” Asian Folklore Studies 51; 103–125. , 2001. “The Role of Tibet in the New Age Movement.” In Imagining Tibet: Perceptions, Projections & Fantasies. Thierry Dodin & Heinz Räther (eds.) Boston: Wisdom Press; 167–182. Kosambi, D.D. 1946. “Early stages of caste system in North India.” Journal of the Bombay Branch of the Royal Asiatic Society 22; 33–48. , 1952. “Ancient Kosala and Magadha.” Journal of the Bombay Branch of the Royal Asiatic Society 27; 180–213. Kramrish, Stella. 1946. The Hindu Temple, 2 vols. Calcutta: University Press. Krishan, Shree. 2001. “Sketch of Darma grammar.” In New Research on Zhangzhung and 480 bibliography

Related Himalayan Languages. Yasuhiko Nagano and Randy J. LaPolla (eds.) Osaka: National Museum of Ethnology; 347–400. Krishna, N. 1980. The Art and Iconography of Vishnu-Narayana. Bombay: Vakils. Kuijp, Leonard W.J. van der. 1981. “On the Life and Political Career of Ta’i-si-tu Byang- chub Rgyal-mtshan (1302–1364).” In The History of Tibet vol. ii. The Medieval Period. The Development of Buddhist Paramountcy. Alex McKay (ed.) London: Routledge- Curzon; 425–466 (first published, 1981. In Tibetan History and Language: Studies ded- icated to Uray Géza on his seventieth birthday. Ernst Steinkellner (ed.) Wien: atbs, vol. 1; 277–327). , 2004. “U rgyan pa Rin chen dpal (1230–1309) Part Two: For Emperor Qubi- lai? His Garland of Tales about Rivers.” In The Relationship between Religion and State (chos srid zung ‘brel) in Traditional Tibet: proceedings of a seminar held in Lumbini, Nepal, March 2000. Christoph Cueppers (ed.) Lumbini: Lumbini Interna- tional Research Institute; 299–339. Kulkarni, Shashank, and Chandrashekhar Tambat. 1981–1982. “Sri Kailas Expedition, 1982. A Two-man Ascent.”Himalayan Journal 39; 62–68. Kumar, Dinesh. 1990. “The Sacred Complex of Badrinath: A Study in Himalayan Pilgrim- age.” In Social Anthropology of Pilgrimage. Makhan Jha (ed.) New Delhi: Inter-India Publications; 205–216. Kumar, K.I. 1981. Expedition Kinner Kailash. New Delhi: Vision Books. Kushal, Molly. 2000. “Divining the Landscape—the Gaddi and his land.” Indian Inter- national Centre Quarterly Winter 2000–Spring 2001; 35–36. Kuznetsov, B.I. 1975. “Who was the Founder of the ‘Bon’ Religion?”The Tibet Journal 1.1; 113–114, (reprinted in 2003. The History of Tibet; vol. 1., The Early Period: to c. ad 850. The Yarlung Dynasty. Alex McKay (ed.) London: RoutledgeCurzon; 457–458). Kværne, Per. 1974. “The Canon of the Tibetan Bonpos.” Indo-Iranian Journal 16.2; 33– 35. , 1987. “Dualism in Tibetan Cosmogonic Myths and the Question of Iranian Influence.” In Silver on Lapis: Tibetan literary culture and history. Christopher I. Beck- with (ed.) Bloomington: The Tibet Society; 163–174. , (ed.) 1994. Tibetan Studies: Proceedings of the 6th Seminar of the International Association for Tibetan Studies Fagernes 1992, 2 vols. Oslo: icrhc. , 1995. The Bon Religion of Tibet: The Iconography of a Living Tradition. London: Serindia. , 1997. “Invocations to Two Bön Deities.” In Religions of Tibet in Practice. Donald S. Lopez (ed.) Princeton: University Press; 395–400. , 1998. “Khyung-sprul ’Jigs-med nam-mkha’i rdo-rje (1897–1955): An Early Twen- tieth-century Tibetan Pilgrim in India.” In Pilgrimage in Tibet. Alex McKay (ed.) Richmond: Curzon Press; 71–84. , 2000. “The Study of Bon in the West: Past, present, and future.” In New Horizons bibliography 481

in Bon Studies: Bon studies 2. Samten G. Karmay & Yasuhiko Nagano (eds.) Osaka: National Museum of Ethnology; 7–20. , 2010. “Bon and Chos—the Other Story.” In Rocznik Orientalistyczny lxiii.i (Altaica et Tibetica: Anniversary Volume dedicated to Stanislaw Godzinski on His Seventieth Birthday): 116–124. Lahiri, Manosi. 2005. Mapping India. Delhi: Niyogi books. , 2006. Here be Yaks: A Personal Account of Travels in Far West Tibet. New Delhi: Stellar Publishers. Lajpat Rai, Lala. 1967. A History of the Arya Samaj. Bombay: Orient Longmans. Lamb, Alastair. 1986. British India and Tibet 1766–1910. London: Routledge & Kegan Paul, (second, revised edition; first published 1960. Britain and Chinese Central Asia: the Road to Lhasa 1767–1905). , 1989. Tibet, China and India: A History of Imperial Diplomacy. Hertingfordbury (uk.): Roxford Books. , 2002. Bhutan and Tibet: The Travels of George Bogle and Alexander Hamilton 1774–1777; Vol. 1, Letters, Journals and Memoranda. Hertingfordbury (uk.): Roxford Books. Landor, A.H.S. 1899. In the forbidden land; an account of a journey in Tibet, capture by the Tibetan authorities, imprisonment, torture, and ultimate release …, 2 vols. London: William Heinemann. Leask, Nigel. 2000. “Francis Wilford and the Colonial Construction of Hindu Geogra- phy.” In Romantic Geographies: discourses of travel, 1775–1844. Amanda Gilroy (ed.) Manchester: University Press; 204–225. Leavitt, J. 1995. “The Demon of Ashes in Sanskrit Text and Himalayan Ritual.” In Beyond Textuality: Asceticism and Violence in Anthropological Interpretation. G. Bibeau & E. Corin (eds.) Berlin/n.y.: Mouton de Gruyter; 79–110. Lecomte-Tilouine, Marie (ed.) 2009. Bards and Mediums: History, Culture and Politics in the Central Himalayan Kingdoms. Almora: Almora Book Depot. Lewis, C.A. 1976. “The Connection between the Geographical texts of the Puranas and those of the Mahabharata.”Purana 18.1; 56–74. Lewis, Todd T. 1993. “Himalayan Frontier Trade: Newar Daispora Merchants and Bud- dhism.” In Anthropology of Tibet and the Himalayas. Martin Brauen (ed.) Zurich: Volkerkunde-museum; 165–178. , 1994. “Himalayan Religions in Comparative Perspective: Considerations Regarding Buddhism and Hinduism across their Indic Frontiers.” Himalayan Research Bulletin 14.1–2; 25–46. Leyerle, Blake. 1996. “Landscape as Cartography in Early Christian Pilgrimage Narra- tives.” jaar lxiv.i; 119–143. Lhagyal, Dondrup. 2000. “Bonpo family lineages in Central Tibet.” In New Horizons in Bon Studies (Bon Studies 2). Samten G. Karmay & Yasuhiko Nagano (eds.) Osaka: National Museum of Ethnology; 429–508. 482 bibliography

Li, Brenda W.L. 2011. “A Critical Study of the Life of the 13th-Century Tibetan Monk U rgyan pa Rin chen dpal Based on his Biographies.” Oxford University Ph.D. thesis. Lohner, Mme. [sic] 1949. “Towards the Himalaya.”Himalayan Journal 15; 20–22. Lopez, Donald S. (ed.) 1997. Religions of Tibet in Practice. Princeton: University Press. , 1999. Prisoners of Shangri-La: Tibetan Buddhism and the West. Chicago: Uni- versity Press. , 2001. “The Image of Tibet of the Great Mystifiers.” In Imagining Tibet: Percep- tions, Projections & Fantasies. Thierry Dodin & Heinz Räther (eds.) Boston: Wisdom Press; 183–200. Lorenzen, David N. 1972. The Kāpālikas and Kālāmukhas: Two Lost Śaivite Sects. (anu Oriental Series, vol. xii) New Delhi: Thomson Press, (also Berkeley; University of California Press, 1972). , 1978. “Warrior Ascetics in Indian History.” jaos 98.1; 61–75. , 1989. “New Data on the Kāpālikas.” In Criminal Gods and Demon Devotees; Essays on the Guardians of Popular Hinduism. Alfred Hiltbeitel (ed.) n.y.: suny; 231–235. , 1993. “Review” (of History and historiography of the age of Harsha. Shankar Goyal). jaos 113.3; 480–481. , 1999. “Who Invented Hinduism?” Journal of Comparative Study of Society and History 41.4; 630–659. , 2001. “A Parody of the Kapalikas in the Mattavilasa.” In Tantra in Practice. David Gordon White (ed.) Delhi: Motilal Banarsidass; 81–96 (first published, 2000. Princeton). , 2002. “Early Evidence for Tantric Religion.” In The Roots of Tantra. Katherine A. Harper and Robert L. Brown (eds.) n.y.: suny; 25–38. Loseries, Andrea. 2011. “Cakrasamvara and Syncretism in the Northwestern Himalaya.” In Buddhist Himalaya: Studies in Religion, History and Culture. Vol. 1 Tibet and the Himalaya. Alex McKay & Anna Balikci-Denjongpa (eds.) Gangtok: nit; 151–156. Loseries-Leick, Andrea. 1998. “On the Sacredness of Mount Kailasa in the Indian and Tibetan Sources.” In Pilgrimage in Tibet. Alex McKay (ed.) Richmond: Curzon Press; 143–164. Luczanits, Christian. 2008. “On the Earliest Mandalas in a Buddhist Context.” In Maha- yana Buddhism. History and Culture. Darrol Bryant & Susan Bryant (eds.) New Delhi: Sambhota Series xv, Tibet House; 111–136. , 2009. “Ritual, Instruction And Experiment: Esoteric Drawings From Dun- huang.” In The Art of Central Asia and the Indian Subcontinent in Cross-Cultural Per- spective. Anupa Pande and Mandira Sharma (eds.) New Delhi: National Museum Institute-Aryan Books International; 140–149. Macdonald, Alexander W. 1985. “Points of View on Halase, a Holy Place in East Nepal.” The Tibet Journal 10.3; 3–13. bibliography 483

, 1987. “Remarks on the Manipulation of Power and Authority in the High Himalayas.” The Tibet Journal 12.1; 3–16. , 1990. “Hindu-isation, Buddha-isation, then Lama-isation or: What Happened at La-Phyi?” In Indo-Tibetan Studies: Papers in honour and appreciation of Professor David L. Snellgrove’s contribution to Indo-Tibetan Studies. Tadeusz Skorupski (ed.) Tring: Buddhica Britannica; 199–208. , (ed.) 1997. Mandala and Landscape. New Delhi: d.k. Printworld. , 2003. “Remarks on the Manipulation of Power and Authority in the High Himalaya.” In The History of Tibet; vol. 1., The Early Period: to c. ad 850. The Yarlung Dynasty. Alex McKay (ed.) London: RoutledgeCurzon; 185–196 (first published in 1987. The Tibet Journal 12.1; 3–16). Macdonald, Ariane [Spanien]. 1962. “Note sur la diffusion de la ‘Théorie des quatre fils du ciel.’” Journal Asiatique 250; 531–548. , 1971. “Une lecture des Pelliot Tibétain 1286, 1287, 1038, 1047, et 1290, essai sur la formation et l’emploi des mythes politiques dans la religion royale de Sroṅ-bcan sgam-po.” In Études tibétaines dédiées à la mémoire de Marcelle Lalou. Paris: Librairie d’Amérique et d’Orient; 190–391. Macdonell, A.A. & A.B. Keith. 1912. Vedic index of names and subjects, 2 vols. London: John Murray. Maclean, Kama. 2008. Pilgrimage and Power: The Kumbh Mela in Allahabad, 1765–1954. n.y.: oup. McKay, Alex. 1997. Tibet and the British Raj. The Frontier Cadre, 1904–1947. Richmond: Curzon Press. , 1998. “Asceticism, Power, and Pilgrimage: Kailas-Manasarovar in Classical and Colonial Indian Sources.” In Pilgrimage in Tibet. Alex McKay (ed.) Richmond: Curzon Press; 165–183. , 1998a. Pilgrimage in Tibet. Richmond: Curzon Press. , 1999. “The British Imperial Influence on the Kailas-Manasarovar Pilgrimage.” In Sacred Spaces and Powerful Places in Tibetan Culture. Toni Huber (ed.) Dharam- sala: lwta; 305–321. , 2002. “The Drowning of Lama Sengchen Kyabying: A Preliminary Enquiry from British Sources.” In Tibet Past and Present: Tibetan Studies i. the Proceedings of the 9th International Seminar for Tibetan Studies, Leiden 2000. Henk Blezer (ed.) Leiden: Brill; 263–280. , (ed.), 2003. The History of Tibet, 3 vols. London: RoutledgeCurzon. , 2003a. ‘Tibet: The Myth of Isolation.’ In The History of Tibet. London: Rout- ledgeCurzon; vol. iii, 635–645 (first published in New Developments in Asian Studies: the i.i.a.s. Yearbook 1996. 1998. Paul van der Velde and Alex McKay (eds.) London: KeganPaul International; 302–318.) , 2009. ‘“Tracing lines upon the unknown areas of the earth’: Reflections on 484 bibliography

the Indo-Tibetan Frontier.” In Fringes of Empire: Peoples, Places and Spaces at the Margins of British Colonial India. Sameetah Agha & Elizabeth Kolsky (eds.) Delhi: oup; 69–93. , 2010. “Tibet, Sikkim, Bhutan: the myth of venereal disease.” In Studies of Medical Pluralism in Tibetan History and Society. piats 2006: Proceedings of the Eleventh Seminar of the International Association for Tibetan Studies. Königswinter 2006. Sienna Craig, Mingji Coumo, Frances Garrett and Mona Schrempf (eds.) Halle: iitbs; 41–60. , 2011. “In Search of Zhang-zhung: the ‘Grey and Empty’ Land?” In EmergingBon. The Formation of Bon Traditions in Tibet at the Turn of the First Millennium ad. piats 2006: Proceedings of the Eleventh Seminar of the International Association for Tibetan Studies. Königswinter 2006. Henk Blezer (ed.) Halle: iitbs; 185–206. , 2015. “‘A very useful lie’: Giuseppe Tucci, Tibet, and scholarship under dic- tatorship.” In Asian Horizons: Tucci, Himalayan, Indian, Buddhist and Central Asian Studies. Volume cvi; David Templeman & Andrea de Castro (eds.) Serie Orientale Roma, Rome/ Melbourne; 68–82. McKay, Alex, and Anna Balikci (eds.) 2011. Buddhist Himalaya: Studies in Sikkim History and Culture, 2 vols. Gangtok: nit. McKeown Arthur P. (ed./trans.) 2010. Rolf Stein’s Tibetica Antiqua: with additional mate- rial. Leiden: Brill. Mabbett, Ian. 1983. “The Symbolism of Mount Meru.”History of Religions 23; 64–83. Madden, Captain. 1846. “Diary of an Excursion to the Shatool and Boorun Passes over the Himalaya, in September, 1845.” Journal of the Asiatic Society 15; 79–135. , [Major]. 1848. “The Turaee and Outer Mountains of Kumaoon.” Journal of the Asiatic Society 17.1, May 1848; 349–350, & June 1848; 563–626. Maguire, Mark Patrick. 2011. “Review” (of Emplacing a Pilgrimage: The ôyama Cult and Regional Religion in Early Modern Japan. Barbara Ambros. 2008. Boston: Harvard University Press.) jaar 79.2; 541. ’”.ag, Old Persian Maguš, and English ‘MagicianץMair, Victor H. 1990. “Old Sinitic *M Early China 15; 27–47. Malalasekera, Gunapali/George P. 1937. Dictionary of Pali Names. London: John Murray. Malandra, Geri Hockfield. 1993. Unfolding a Mandala: The Buddhist Cave Temples at Ellora. n.y.: suny. Malandra, William W.1983. AnIntroductiontoAncientIranianreligion:Readingfromthe Avesta and Achaemenid Inscriptions. Minneapolis: University of Minnesota Press. Malhotra, S.S.L. 1983. Gangotri & Gaumukh: A Trek to the Holy Source. New Delhi: Allied Publishers. Malla, Kamal P. 1992. “The Nepāla-Mahāmya: A ix-Century Text or a Pious Fraud?” Contributions to Nepalese Studies 19.1; 145–158. Mallinson, W. James. 1993. ‘“Some Common Faqueers of an Inferior Order’. Saddhuism in Perspective”. University of London ma thesis. bibliography 485

, 2001. “The Khecarīvidyā of Ādinātha: A Critical Edition and Annotated Trans- lation.” Oxford University (Balliol) Ph.D. thesis. Malvania, Dalsukh. (ed.) 1970. Āgamic Index, volume 1, Prakrit Proper Names Part 1. Ahmedabhad: l.d. Institute of Indology. Mani, Vettam. 1975. Purāṇic encyclopedia: a comprehensive dictionary with special refer- ence to the epic and Puranic literature. Delhi: Motilal Banarsidass. Martin, Dan. 1982. “The Early Education of Milarepa.” Journal of the Tibet Society 2; 53–75. , 1994. Mandala Cosmology Human Body Good Thought and the Revelation of the Secret Mother Tantras of Bon. Wiesbaden: Harrassowitz Verlag. , 1994a. “Tibet at the Centre: a historical study of some Tibetan geographical conceptions based on two types of country lists found in Bon histories.” In Tibetan Studies: Proceedings of the Sixth Seminar of the International Association for Tibetan Studies, Fagernes, 1992; 2 vols. Per Kvaerne (ed.) Oslo: icrhc; vol. 1; 408–429. , 1995. “‘Ol-mo-lung-ring, the Original Holy Place.” The Tibet Journal 20.1; 48–82, (reprinted in 1999. In Sacred Spaces and Powerful Places in Tibetan Culture. Toni Huber [ed.] Dharamsala: lwta; 258–304). , 1996. “Lay religious Movements in 11th and 12th-Century Tibet: A Survery of the Sources.”Kailash xviii. iii–iv; 23–55. , 1997. “Beyond Acceptance and Rejection? The Anti-Bon Polemic included in the Thirteenth-Century Single Intention (Dgongs-gcig Yig-cha) and Its Background in Tibetan Religious History.” Journal of Indian Philosophy 25.3; 263–305. , 2001. Unearthing Bon Treasures: Life and Contested Legacy of a Tibetan Scrip- ture Revealer. With a General Biography of Bon. Leiden: Brill. Martyn, J.A.K. 1938. “Across the Gangotri-Alaknanda Watershed.”Himalayan Journal 10; 79–85. Mayer, Robert. 1990. “The Origins of the Esoteric Vajrayāna.” Paper presented at the Buddhist Forum, London University soas, October 1990. , 1997. “The Sa-skya Paṇḍita, the White Panacea, and Clerical Buddhism’s Cur- rent Credibility Crisis.” The Tibet Journal 22.3; 79–105. Mejor, Marek. 1987. “Review” (of Marcel Hofinger: 1982; op. cit.). Rocznik Orientalisty- czny 45.2; 90–93. Mehra, P.L. 1968. “Sikkim and Bhutan: An Historical Conspectus.” Journal of Indian History xlvi.1; 89–124. Meij, Dick van der. (ed.) 1997. India and Beyond: Aspects of Literature, Meaning, Ritual and Thought. Essays in Honour of Frits Staal. London/Leiden: Kegan Paul Interna- tional/iias. Mendelson, E. Michael. 1961. “The King of the Weaving Mountain.” Journal of the Royal Central Asian Society 48.3–4; 229–237. Messerschmidt, D.A., & J. Sharma. 1982. “Social Process on the Hindu Pilgrimage to Muktinath.”Kailash ix.ii–iii; 139–158. 486 bibliography

Meulenbeld, G.R. 1989. “The search for clues to the chronology of Sanskrit medical texts as illustrated through the history of bhaṅgā.” Studien zur Indologie und Iranstik, 15; 59–70. Meyer, Karl E., and Sharon Blair Brysac. 1999. Tournament of Shadows: The Great Game and the Race for Empire in Central Asia. Washington d.c.: Counterpoint. Michaels, Axel. 1990. “Pilgrimage and Priesthood at the Pasupatinatha Temple of Deo- patan (Nepal).” In The History of Sacred Places in India as reflected in traditional Literature: Papers on Pilgrimage in South Asia. [Panels of the viith World Sanskrit Conference, Leiden 1987] Hans Bakker (ed.) Leiden: Brill; 131–159. Millard, Colin. 2009. “The Life and Medical Legacy of Khyung sprul ‘Jigs med nam mkha’i rdo rje (1897–1955).” In Bon The Everlasting Religion Of Tibet: Tibetan Studies in Honour of Professor David L. Snellgrove, New Horizons in Bon Studies 2. Samten G. Karmay and Donatella Rossi (eds.) (Special edition of) East and West 59.1–4; 147–166. Miller, Jeanine. 1985. The Vision of Cosmic Order in the Vedas. London: Routledge & Kegan Paul. Mills, Martin. 1997. “The Religion of Locality: Local Area Gods and the Characterisation of Tibetan Buddhism.” In Recent Research on Ladakh 7. Thierry Dodin & Hans Räther (eds.) Bonn: Ulmer Kulturanthroplogische Schriften; 309–328. Mimaki, Katsumi. 2000. “A preliminary comparison of Bonpo and Buddhist cosmology.” In New Horizons in Bon Studies: Bon studies 2. Samten G. Karmay & Yasuhiko Nagano (eds.) Osaka: National Museum of Ethnology; 89–115. Mishra, Ibhuti. 1973. Religious Beliefs and Practices of North India During the Early Medieval Period. Leiden: Brill. Mishra, Kamala Prasad. 1975. Banaras in Transition (1738–1795): A Socio-Economic Study. Delhi: Munshiram Manoharlal. Misra, Ram Nath. 1981. Yaksha Cult and Iconography. Delhi: Motilal Banarsidass. Mithal, R.S. 1968. “The Physiographical and Structural Evolution of the Himalaya.” In Mountains and Rivers of India. B.C. Law (ed.) Calcutta: National Committee for Geography. Mittal, A.K. 1985. “Trans-Himalayan Trade with Tibet during the British Rule (With special reference to Kumaon).” Journal of Indian History lxiii: i–iii; 63–82. Modak, B.R. 1993. The Ancillary Literature of the Atharvaveda: A Study with Special Reference to the Pariśiṣṭas. New Delhi: Rashtriya Veda Vidya Pratishthan. Monier-Williams, Monier. 1976. ADictionary,EnglishandSanskrit. Delhi: Motilal Banar- sidass (first published, 1851). Moorcroft, W. 1816. “A Journey to Lake Mánasaróvara in Ún-dés, a province of little Tibet.” Asiatik Researches 12; 375–535. Moran, Arik. 2006. “Trade, Tax and Timber—The Transformation of Kingship in Bash- ahr, 1815–1914”. Oxford University ma History dissertation. bibliography 487

, 2007. “From mountain trade to jungle politics: The transformation of kingship in Bashahr, 1815–1914.”ieshr 44.2; 147–177. , 2013. “Toward a history of devotional Vaishnavism in the West Himalayas: Kullu and the Ramanandis, c. 1500–1800”. ieshr 50.1; 1–25. Moran, M. 2003. “Little Kailash Pilgrimage.” The Himalayan Journal 59; 82–92. Morinis, E. Alan. 1984. Pilgrimage in the Indian Tradition: A Case Study of West Bengal. Delhi: oup. Muller, Max, and Hermann Jacobi (trans.) 1989. Jaina Sutras (Part 1: The Akaranga Sutra): Delhi: Motilal Banarsidass (first published n.y., 1884). http://www.abebooks .co.uk/servlet/BookDetailsPL?bi=3848283332&searchurl=isbn%3D8120801237. Müller-Ebeling, Claudia, & Christian Rätsch & Surendra Bahadur Shahi. 2002. Shaman- ism and Tantra in the Himalayas. London: Thames and Hudson, (first published as Schamanismus und Tantra in Nepal. Switzerland: at Verlag). Mus, Paul. 2010. India as seen from the East. India and Indigenous Cults in Chamba. (Ian Mabbit, trans.) Caulfield: Monash University Press (first French publication, 1933). Nagano, Yasuhiko, and Randy J. LaPolla (eds.) 2001. New research on Zhangzhung and Related Himalayan Languages. Osaka: National Museum of Ethnology. Nakamura, Hajime. 1987. Indian Buddhism: a survey with bibliographical notes. Delhi: Motilal Banarsidass (first published, Japan, 1980). Nalesini, Oscar. 2008. “Assembling Loose Pages, Gathering Fragments of the Past: Giu- seppe Tucci and His Wanderings Throughout Tibet and the Himalayas, 1926–1954.” In Sanskrit Texts from Giuseppe Tucci’s Collection Part i, Francesco Sferra (ed.) Rome: Istituto Italiano per L’Africa e L’Oriente; 79–112. Nand, Nitya and Kamlesh Kumar. 1989. The Holy Himalaya: A Geographical Interpreta- tion of Garwhal. Delhi: Daya Publishing. Naquin, Susan, and Chün-fang Yü (eds.) 1992. Pilgrims and Sacred Sites in China. Berke- ley: University of California Press. Nebesky-Wojkowitz, René. 1955. Where the Gods are Mountains: Three Years among the People of the Himalayas. London: Weidenfeld & Nicolson. , 1975. Oracles and Demons of Tibet: the cult and iconography of the Tibetan protective deities. Delhi: Tiwari’s Pilgrims Book House (first published, 1956). Neumaier, E.K. 2002. “Historiography and the Narration of Cultural Identity in Zanskar.” In Tibet, Past and Present: Tibetan Studies i: piats 2000: Proceedings of the Ninth Seminar of the International Association for Tibetan Studies, Leiden 2000. Henk Blezer (ed.) Leiden: Brill; 327–341. Neve, Arthur. 1900. Picturesque Kashmir. London: Sands & co. Newman, John Ronald. 1987. “The Outer Wheel of Time: Vajrayāna Buddhist Chronol- ogy in the Kālachakra Tantra.” Ann Arbor umi, Ph.D. thesis. Ngawang Zangpo. 2001. Sacred Ground: Jamgon Kongtrul on ‘Pilgrimage and Sacred Geography’. Ithaca, n.y.: Snowlion. 488 bibliography

Noble, C. 1988. Over the High Passes: A Year in the Himalayas with the Migratory Gaddi Shepherds. London: Fontana (first published, 1987). Norbu, Namkhai. 1995. Drung, Deu and Bön. Narrations, Symbolic languages and the Bön tradition in ancient Tibet. Dharamsala: lwta. , 2009. The Light of Kailash: A History of Zhang Zhung and Tibet. Volume One. The Early Period. Merigar: Shang Shung publications. Norbu, Namkhai, & Ramon Prats. 1989. Gans Ti Se’i dKar c’ag: A Bon-po Story of the Sacred Mountain Ti-se and the Blue Lake Ma-pan. Rome: IsMEO. Norman, K.R. 1983. “Review” (of Traditions of the Seven Ṛṣis. John E. Mitchiner. 1982. Delhi: Motilal Banarsidass), jras (new series) 2; 318–319. , 1990. “Review” (‘The coming of the Aryans to Iran and India and the cultural and ethnic identity of the Dasas.’ Asko Parpola; Studia Orientalia 64: 195–302). Acta Orientalia li; 288–297. Oakley, E. Sherman. 1905. Holy Himalaya. The Religion, Traditions and Scenery of a Himalayan Province (Kumaon and Garwhal). Edinburgh & London: Oliphant Ander- son and Ferrier. Obeyesekere, G. 1987. The Cult of the Goddess Pattini. New Delhi: Motilal Banarsidass (first published, Chicago, 1984). Ohri, Vishwa Chander. 1991. Sculpture of the Western Himalayas (History and Stylistic Development). Delhi: Agam Kala Prakashan. Ohri, Vishwa Chander, & Amarnath Khanna (eds.) 1989. A Western Himalayan King- dom: History and Culture of the Chamba State. Delhi: Books and Books. Olivelle, Patrick. 1974. “The Notion of Āśrama in the Dharmaśāstras.” wzks 18; 27– 35. , 1987. “King and Ascetic.” The Adyar Library Bulletin, Festschrift for Ludo Rocher 51; 39–59. , 1993. The Āśrama System: History and Hermeneutics of a Religious Institution. n.y.: oup. Orr, W.G. 1940. “Armed Religious Ascetics in North India.” Bulletin of the John Ryland Library 24.1, 81–100. Ortner, Sherry B. 1992. High Religion: A Cultural and Political History of Sherpa Bud- dhism. Delhi: Motilal Banarsidass (first published, Princeton, 1989). O’Shaughnessy, W.B. 1839. “On the preparation of Indian Hemp, or Gunjah.” Journal of the Asiatic Society of Bengal 8.2; 732–742. Osmaston, Gordon. 1938. “Surveys and Various Expeditions: Progress of Himalayan Surveys.”Himalayan Journal 10; 177–183. , 1939. “Gangotri Triangulation.”Himalayan Journal 11; 126–131. Osmaston, Henry (ed.) 1996. Ladakh Studies: volume 8. Cumbria: privately published. Ottley, J.F.S. 1940. “The Jadh Ganga Valley and the Nela Pass.” Himalayan Journal 12; 27–29. bibliography 489

Paine, Jeffrey. 1999. Father India: Westerners Under the Spell of an Ancient Culture. n.y.: HarperPerennial. Pal, P. 1988. “The Fifty-One Sakta Pithas.” In Orientalia Iosephi Tucci Memoriae Dicata. E. Curaverunt, G. Gnoli, & L. Lanciotti (eds.) Rome: IsMEO; 1039–1060. Pande, Badri Datt. 1993. (C.M. Agrawal, trans.) History of Kumaun (Kumaun ka Itihās) Vol. i. Almora: Shyam Prakashan (first Hindi publication, 1937). Panglung, J.L. 1992. “On the Narrative of the Killing of the Evil Yak and the Discovery of Salt in the Cho-‘byung of Nyang-ral.” In Tibetan Studies: Proceedings of the 5th Seminar of the International Association for Tibetan Studies, Narita 1989. S. Ihara & Z. Yamaguchi (eds.) Narita: Naritasan Shinshoji; vol. 2; 661–667. Paramasiva, Vasudha. 2009. ‘Yah Ayodhyā Vah Ayodhyā: Earthly and Cosmic Journeys in the Ānand-laharī.’ In Patronage and Popularisation, Pilgrimage and Procession: Channels of Transcultural Translation and Transmission in Early Modern South Asia; Papers in Honour of Monika Horstmann. Heidi Rika Maria Pauwels (ed.) Wiesbaden: Harrassowitz Verlag; 101–115. Parpola, Asko. 1988. “The Coming of the Aryans to Iran and India and the Cultural and Ethnic Identity of the Dāsas.” Studia Orientalia 64; 195–302. , 1992. “The metamorphoses of Mahisa Asura and Prajapati.” In Ritual, state and history in South Asia: Essays in honour of J.C. Heesterman. A.W. van den Hoek, D.H.A. Kolff and M.S. Oort (eds.) Leiden: Brill; 275–308. , 1995. “The Problem of the Aryans and the Soma: Textual-linguistic and archae- ological evidence.” In The Indo-Aryans of Ancient South Asia: Language, Material Culture and Ethnicity. George Erdosy (ed.) Berlin/n.y.: Walter de Gruyter; 353–379. , 1997. “The Dasa and the Coming of the Aryans.” In Inside the Texts, Beyond the Texts: New Approaches to the Study of the Vedas. Michael Witzel (ed.) Harvard: Harvard Oriental Series (Opera Minora vol. 2); 193–202. , 2002. “Pre-Proto-Iranians of Afghanistan as Initiators of Śāmkta Tantrism: On the Scythians/Saka Affiliation of the Dāsas, Nuristanis and Magadhans.” Iranica Antiqua 37; 233–324. , 2004–2005. “The Nāsatyas, the Chariot and Proto-Aryan Religion.” Journal of Indological Studies 16/17; 1–62. Pedersen, Poul. 2001. “Tibet, Theosophy, and the Psychologization of Buddhism.” In Imagining Tibet: Perceptions, Projections & Fantasies. Thierry Dodin & Heinz Räther (eds.) Boston: Wisdom Press; 151–166. Petech, Luciano. 1947. “The Tibet-Ladakh-Mughal War of 1681–1683.” Indian Historical Quarterly xxiii.iii; 169–199. , 1977. The Kingdom of Ladakh c. 950–1842 a.d. Rome: IsMEO. , 1978. “The ‘Bri.guṅ.pa Sect in Western Tibet and Ladakh.” In Proceedings of the Csoma de Kórös Memorial Symposium. L. Ligeti (ed.) Budapest: Akademiai Kiadó; 313–325. 490 bibliography

, 1988. “The Missions of Bogle and Turner according to the Tibetan texts.” In Selected Papers on Asian History. Luciano Petech. Roma: IsMEO; 49–62 (first published, T’oung Pao, 1950). , 1997. “Western Tibet: Historical Introduction.” In Tabo. A Lamp for the King- dom, Early Indo-Tibetan Buddhist Art in the Western Himalaya. Deborah E. Klimburg- Salter (ed.) Milan: Skira Editori; 229–255. , 2003. “Ya-Ts‘e, Gu-Ge, Pu-Ran: A New Study.” In History of Tibet. Volume 2. The Medieval Period. The Rise of Buddhist Paramountcy. Alex McKay (ed.) London: RoutledgeCurzon; 33–52, (first published in 1980.The Central Asiatic Journal 24.1–2; 85–111; reprinted in 1998. Selected Papers on Asian History. Luciano Petech. Rome: IsMEO; 369–394.) , 2003. “The Dalai-Lamas and Regents of Tibet: A Chronological Study.” In History of Tibet. Volume 2. The Medieval Period. The Rise of Buddhist Paramountcy. Alex McKay (ed.) London: RoutledgeCurzon; 567–583 (first published in 1959. T’oung Pao xlvii; 368–394; reprinted in Selected Papers on Asian History. Luciano Petech. Roma: IsMEO; 125–147). Phuntsho, Karma. 2013. The History of Bhutan. New Delhi: Random House. Pinch, William R. 2006. Warrior Ascetics and Indian Empires. Cambridge: cup. Pinch, Vijay (William R.). 2003. “Bhakti and the British Empire.” Past and Present 179; 159–196. Pinkney, Andrea. 2013. “An Ever-Present History in the Land of the Gods: Modern Māhātmya Writing on Uttarakhand.” International Journal of Hindu Studies 17.3; 231–262. Pintchman, Tracy. 1994. The Rise of the Goddess in the Hindu Tradition. n.y.: suny. Pollack, Sheldon. 2001. “New Intellectuals in Seventeenth-Century India.” ieshr 38.1; 3–31. , (ed.) 2011. Forms of Knowledge in Early Modern Asia. Durham/London: Duke University Press. Pommaret, Françoise. 1996. “On local and mountain deities in Bhutan.” In Reflections of the Mountain: Essays on the History and Social Meaning of the Mountain Cult in Tibet and the Himalaya. Anne-Marie Blondeau and Ernst Steinkellner (eds.) Wien: vöaw; 39–58. , 2004. “Yul and yul lha: the territory and its deity in Bhutan.”Bulletin of Tibetol- ogy 40.1; 39–68. Porter, Bill. 1993. Road to Heaven: Encounters with Chinese Hermits. London: Rider. Porter, Deborah. 1993. “The Literary function of K’un-lun Mountain in the Mu T’ien-tzu Chua.”Early China 18; 73–106. Postell, M., A. Neven, & K. Mankodi. 1985. Antiquities of Himachal, vol. 1. Bombay: Project for Indian Cultural Studies. Powell-Price, J.C. 1930. “Some Notes on the early History of Kumaon.” Journal of the United Provinces Historical Society 4.2; 5–16. bibliography 491

, 1945. “Kunindas and Katyurs.” Journal of the United Provinces Historical Society 18.2; 213–223. Powers, John. 2009. A Bull of a Man: Images of Masculinity, Sex, and the Body in Indian Buddhism. Cambridge (Mass.)/London: Harvard University Press. Pranavananda, Swami. 1946. “New Light on the Sources of the Four Great Rivers of the Holy Kailas and Manasarovar.” Journal of the United Provinces Historical Society 19.1; 168–180. , 1950. Exploration in Tibet. Calcutta: privately published. , 1984. Kailās-Mānasarovar. Delhi: privately published (first published, 1949). Prescott, J.R.V. 1965. The Geography of Frontiers and Boundaries. London: Hutchinson. , 1972. Political Geography. London: Methuan. Pringle, Robert. 1885. “Ancient and Modern Methods of Treating Small-pox Epidemics in India.” Journal of the Royal Society of Arts, xxxiii; 737–739. Puranik, G.D. 1999. Beyond Himalaya: Kailas Manasarovar Yatra 1998. Mumbai: Samvad Prakashan. Quintman, Andrew. 2006. “Mi la ras pa’s many lives: Anatomy of a Tibetan biographical corpus.” University of Michigan Ph.D. thesis. , 2008. “Toward a Geographic Biography: Mi la ras pa in the Tibetan Landscape.” Numen 55.4; 363–410. Quraishi, Salim al-Din. 1997. Cry for freedom: proclamations of Muslim revolutionaries of 1857. Lahore: Sang-e Meel Publications. Ramble, Charles. 1996. “Patterns of places.” In Reflections of the Mountain: Essays on the History and Social Meaning of the Mountain Cult in Tibet and the Himalaya. Anne-Marie Blondeau & Ernst Steinkellner (eds.) Wien: vöaw; 141–156. , 1997. “Se: preliminary notes on the distribution of an ethnonym in Tibet and Nepal.” In Les Habitants du Toit du monde, Hommage à Alexander W. Macdonald Études Recueillies par les soins de Samten Karmay and Philippe Sagant. Samten Karmay and Philippe Sagant (eds.) Nanterre: Société d’ethnologie; 485–513. , 1998. “The Classification of Territorial Divinities in Pagan and Buddhist Rituals of South Mustang.” In Tibetan Mountain Deities: Their Cults and Representations: Pro- ceedings of the 7th International Seminar for Tibetan Studies, Graz 1995. Anne-Marie Blondeau (ed.) Wein: vöaw; 123–144. , 1999. “The Politics of Sacred Space in Bon and Tibetan Popular Tradition.” In Sacred Spaces and Powerful Places in Tibetan Culture: A Collection of Essays. Toni Huber (ed.) Dharamsala: lwta; 3–33. (A revised version of “Gaining Ground: Repre- sentations of Territory in Bon and Tibetan Popular Tradition.” 1995. The Tibet Journal 20.1; 83–124.) , 2008. “A nineteenth-century Bonpo pilgrim in Western Tibet and Nepal: epi- sodes from the life of dKar-ru grub-dbang bsTan-‘dzin rin-chen.” In Tibetan Studies in Honour of Samten Karmay. Revue d’études Tibétaines 15.2; 481–502. 492 bibliography

Ramble, Charles, and Martin Brauen. (eds.) 1993. Anthropology of Tibet and the Hima- laya. Zurich: Ethnological Museum of the University of Zurich. Ramble, Charles, and Marietta Kind. 2003. “Bonpo monasteries and temples of the Himalayan region.” In A Survey of Bonpo Monasteries and Temples in Tibet and the Himalaya,(Bon Studies 7). Samten G. Karmay and Yasuhiko Nagano (eds.) Osaka: National Museum of Ethnology; 669–752. Rao, Velcheru Narayana. 1993. “Purāṇa as Brahminic Ideology.” In Purāṇa Perennis: Reciprocity and Transformation in Hindu and Jaina Texts. Wendy Doniger (ed.) n.y.: suny; 85–100. Raper, F.V. 1812. “Narrative of a Survey for the purpose of discovering the Sources of the Ganges.” Asiatick Researches 11; 446–563. Rath, B.K. 1976. “The History of Mount Mahendra.” Orissa Research Journal 22.2; 80–98. Ratnagar, S. 1987. “Pastoralists in the Prehistory of Baluchistan.” Studies in History 3.2; 137–154. Rauber, Hanna. 1980. “The Humli-Khyampas of Far Western Nepal: A Study in Ethno- genesis.” Contributions to Nepalese Studies viii.i; 57–79. Rawat, A.S. 1989. History of Garhwal 1358–1947; An Erstwhile Kingdom in the Himalayas. New Delhi: Indus. Rawling, C.G. 1905. The great plateau, being an account of exploration in Central Tibet, 1903, and of the Gartok Expedition, 1904–1905. London: Edward Arnold. Rawlinson, Andrew. 1986. “Nāgas and the Magical Cosmology of Buddhism.” Religion 16.2; 135–153. Raza, Syed Aiman. n.d. (c. 2008). “Culture, Environment and Sustainable Development among the Kinnaurese of Morang Tehsil, Kinnaur district, Himachal Pradesh.” Uni- versity of Delhi (Dept. of Anthropology) Ph.D. thesis. Regmi, D.R. 1983. Inscriptions of Ancient Nepal, 3 vols. New Delhi: Abhinav Publications. Renou, Louis. 1957. Vedic India. Calcutta: Susil Gupta. Riaboff, Isabelle. 1996. “gZhon nu mdung lag, mountain god of Zanskar—A Regional scale divinity and its cult’s territorial ordering.” In Reflections of the Mountain: Essays on the History and Social Meaning of the Mountain Cult in Tibet and the Himalaya. Anne-Marie Blondeau and Ernst Steinkellner (eds.) Wien: vöaw; 23–38. , 2004. “Rituals for the local gods among the Bod of Paldar.”Études mongoles et sibériennes, centrasiatiques et tibétaines 35; 185–201. Ricard, Mathieu, et al., (eds./trans.) 1994. The Life of Shabkar: The Autobiography of a Tibetan Yogin: by Źabs-dkar Tshogs-drug-raṅ-grol, 1781–1851. n.y.: suny. Richardson, Hugh. 1980. “The First Tibetan Chos-‘byung.” The Tibet Journal 5.3; 62–73 (reprinted in Hugh Richardson, 1988. High Peaks, Pure Earth. Collected Writings on Tibetan History and Culture (Michael Aris: ed.) London: Serindia; 89–99). , 1984. Tibet and its History (revised edition). Boston/London: Shambhala (first published, 1962). bibliography 493

, 1985. A Corpus of Early Tibetan Inscriptions. London: Royal Asiatic Society. , 1988. “Further Fragments from Dunhuang.” In Hugh Richardson, High Peaks, Pure Earth. Collected Writings on Tibetan History and Culture (Michael Aris: ed.) Lon- don: Serindia; 28–36 (first published in Bulletin of Tibetology 2.3, 1965; 33–38). , 2003. “The Origin of the Tibetan Kingdom.” In The History of Tibet. vol. 1, the Early Period to c. a.d. 850. The Yarlung Dynasty. Alex McKay (ed.) London: RoutledgeCurzon; 156–164 (first published in Bulletin of Tibetology 3: 1989; 5–19. Reprinted in Richardson. 1988. Op. cit.; 124–134). , 2003. “The Political Role of the Four Sects in Tibetan History.” In The History of Tibet vol. ii. The Medieval Period. The Development of Buddhist Paramountcy. Alex McKay (ed.) London: RoutledgeCurzon; 165–173 (first published in Tibetan Review 11.9: 1976, 18–23; reprinted in Richardson. 1988. Op. cit.; 420–430). Richardson, Hugh, and David Snellgrove. 1968. A Cultural History of Tibet. London: Weidenfeld & Nicolson. Robb, Peter. 1997. “The colonial state and constructions of Indian identity: an example on the north-east frontier in the 1880s.” Modern Asian Studies 31.2; 245–283. , 2002. A History of India. Basingstoke: Palgrave. Roberts, Peter Alan. 2007.The Biographies of Rechungpa: The evolution of a Tibetan hagiography. London/ n.y.: Routledge. Robson, James. 2009. Power of Place: The Religious Landscape of the Southern Sacred Peak (Nanyue) in Medieval China. Cambridge: Harvard University Press. Rocher, Ludo. 1986. The Purāṇas. Wiesbaden: Otto Harrassowitz. , 1988. “The Concept of Boundaries in Classical India.” In The Countries of South Asia: Boundaries, Extensions and Interrelations. Peter Gaeffke and David Utz (eds.) Philadelphia: University of Pennsylvania; 1–10. Rock, Joseph. 1930. “Seeking the mountains of mystery. An expedition on the China- Tibet frontier to the unexplored Amnyi Machen Range.” The National Geographic Magazine lvii.ii; 131–185. Rockhill, W.W. 1891. “Tibet: A Geographical, Ethnographical and Historical Sketch, Derived from Chinese Sources.” Journal of the Royal Asiatic Society 23.1; 1–133 and 185–291. Rodseth, Lars Thomas. 1994. “Travel and transcendence: Lamaist expansion in the Himalayan kingdoms.” University of Michigan Ph.D. thesis. Roerich, George, N. 1996. Trails to Inmost Asia: five years of exploration with the Roerich Central Asian Expedition. Delhi: Book Faith (first published, 1931). Roerich, George N. & Gedun Choepal (ed/trans.). 1996. The Blue Annals. Delhi: Motilal Banarsidass (first published, 1949). Roesler, Ulrike. 2002. “The Great Indian Epics in the Version of Dmar Ston Kyi Rgyal Po.” In Religion and Secular Culture in Tibet. Tibetan Studies ii. piats 2000. Tibetan Studies: Proceedings of the Ninth Seminar of the International Association for Tibetan Studies, Leiden 2000. Henk Blezer (ed.) Leiden: Brill; 431–450. 494 bibliography

Rose, H.A., & J. Hutchinson. 1996. Gazetteer of the Chamba State 1904, Punjab States Gazetteer Vol. xxiia. New Delhi: Indus publications (first published, 1904). Rosenfield, J.M. 1967. TheDynasticArtsoftheKushans. Berkeley: University of California Press. Roy, Kumkum. 1993. “In which part of south Asia did the early brahmanical tradition (1st millennium b.c.) take its form?” Studies in History 9.1; 1–32. , 2011. “Peopled Landscapes—Pilgrims and (Sup)porters on the Kailas-Manasa- rovar Yatra in the 20th and 21st Centuries.” In Buddhist Himalaya: Studies in History, Religion and Culture, vol. 1, Tibet and the Himalayas. Alex McKay and Anna Balikci (eds.) Gangtok: nit; 323–338. Ruegg, David Seyfort. 2003. “Problems in the Transmission of Vajrayāna Buddhism in the Western Himalaya about the year 1000”,in TheHistoryofTibetvol.ii.TheMedieval Period. The Development of Buddhist Paramountcy. Alex McKay (ed.) London: Rout- ledgeCurzon; 123–133 (first published in 1984. Acta Indologica (Naritasan Shinshoji) vi; 369–381). Russell, Jeremy. 1996. “The Tokdens of Khampagar.” Chö Yang 7; 110–116. Said, Edward. 1978. Orientalism. n.y.: Pantheon Press. Saklani, Atul. 1987. The History of a Himalayan Princeley State: Changes, Conflicts and Awakening (An Interpretative History of Princely State of Tehri Garwal u.p.) (a.d. 1815 to 1949 a.d.). Delhi: Durga Publications. Salomon, Richard. 1999. Ancient Buddhist Scrolls from Gandhāra. The British Library Kharoṣṭhī Fragments. Seattle: University of Washington Press. , 2008. Two Gandhari Manuscripts of the Song of Lake Anavatapta (Anavatapta- gatha): British Library Kharosthi Fragment 1 and Senior Scroll 14. Seattle: University of Washington Press. Samdup, Kazi Dawa. 1919. Shrīchakrasambhara Tantra: A Buddhist Tantra. Calcutta & London: Luzac & co. Samdup, P.W. 2008. “A Brief Biography of Kazi Dawa Samdup (1868–1922).” Bulletin of Tibetology 44.1–2; 155–158. Samuel, Geoffrey. 1989. “The Body in Buddhist and Hindu Tantra: Some Notes.”Religion 19.3; 197–210. , 1993. Civilised Shamans: Buddhism in Tibetan Societies. Washington/London: Smithsonian Institution. , 1997. “Some Reflections on the Vajrayana and its Shamanic Origins.” In Les Habitants du Toit du Monde: Études Recueillies en Hommage à Alexander W. Mac- donald. Samten Karmay and Philippe Sagant (eds.) Nanterre: Société d’Ethnologie; 325–342. , 2000. “The Indus Valley civilization and early Tibet.” In New Horizons in Bon Studies: Bon studies 2. Samten G. Karmay & Yasuhiko Nagano (eds.) Osaka: National Museum of Ethnology; 651–670. bibliography 495

, 2002. “Buddhism and State in Eighth Century Tibet.” In Religion and Secular Culture in Tibet: Tibetan Studies ii: piats 2000: Tibetan Studies: Proceedings of the Ninth Seminar of the International Association for Tibetan Studies, Leiden 2000. Henk Blezer (ed.) Leiden: Brill; 1–19. , 2008.The Origins of Yoga and Tantra: Indic Religions to the Thirteenth Century. Cambridge: cup. Samuel, Geoffrey, H. Gregor and E. Stutchbury (eds.) 1994. Tantra and Popular Religion in Tibet. New Delhi: Aditya Prakashan. Sanan, Deepak, and Dhanu Swadi. 1998. Exploring Kinnaur and Spiti in the Trans- Himalaya. Delhi: Indus. Sanderson, Alexis. 1985. “Power and Purity among the Brahmans of Kashmir.” In The Category of the Person: Anthropology, Philosophy, History. Michael Carrithers, Steven Collins and Steven Lukes (eds.) Cambridge: cup; 190–216. , 1988. “Śaivism and the Tantric Traditions.” In The World’s Religions: The Reli- gions of Asia. Friedhelm Hardy (ed.) London: Routledge; 660–674. , 1994. “Vajrayāna: Origin and Function.” in Mettananda Bhikku et al., (eds.) Buddhism into the Year 2000: International Conference Proceedings. Bangkok/Los Angeles: Dhammakaya Foundation; 87–102. , 2001. “History through Textual Criticism in the Study of Śaivism, the Pañcarāta and the Buddhist Yoginītantras.” In Sources and Time. A Colloquium, Pondicherry, 11–13 January 1997. Françoise Grimal (ed.) Pondicherry: École Française d’Extrême Orient; 1–47. , 2003–2004. “The Śaiva Religion among the Khmers. Part i.” Bulletin de L’École Française d’Extrême Orient 90–91; 349–462. , 2007. “Atharvavedins in Tantric Territory: The Āṅgirasakalpa Texts of the Oriya Paippalādins and their Connection with the Trika and the Kālīkula. With critical editions of the Parājapavidhi, the Parāmantravidhi, and the *Bhadrakālīmantravid- hiprakaraṇa.” In The Atharvaveda and its Paippalādaśākhā: Historical and Phlio- logical Papers on a Vedic Tradition. Arlo Griffiths and Annette Schmiedchen (eds.) Aachen: Shaker Verlag; 195–312. Sangarakshita (D.P.E. Lingwood). 1991. Facing Kachenjunga: An English Buddhist in the Eastern Himalayas. Glasgow: Windhorse publications. Sarcar, S.C. 1931. “Some Notes on the Intercourse of Bengal with the Northern Countries in the Second Half of the Eighteenth Century.”Bengal Past and Present xli; 119–128. Sarkar, Jadunath. [and Nirod Bhusan Roy]. 1957. A history of Dashnami Naga Sanyasis. Allahabad: Sri Panchayati Akhara Mahanirvani. Satchidananda, Sri Swami. 1984. Kailas Journal. Pilgrimage into the Himalayas. Virginia: Integral Yoga publications (first published in Tamil, n.d.). Sax, W. 2002. Dancing the Self: Personhood and Performance in the Pandav Lila of Garh- wal. n.y.: oup. 496 bibliography

Saxena, Savitri. 1995. Geographical Survey of the Purāṇas. Delhi: Nag. Schaeffer, Kurtis R. 2002. “The Attainment of Immortality. From Nāthas in India to Buddhists in Tibet.” Journal of Indian Philosophy 30; 515–533. , 2011. “New Scholarship in Tibet, 1650–1700.” In Forms of Knowledge in Early Modern Asia: Explorations in the Intellectual History of India and Tibet, 1500–1800. Sheldon Pollock (ed.) Durham and London: Duke University Press; 292–310. Schaik, Sam van. 2004. “The Early Days of the Great Perfection.” Journal of the Interna- tional Association of Buddhist Studies 27.1; 165–202. , 2010. “The Limits of Transgression’: The Samaya vows of Mahāyoga.” In Eso- teric Buddhism at Dunhuang: Rites and Teachings for This Life and Beyond. Mathew T. Kapstein & Sam van Schaik (eds.) Leiden: Brill; 61–84. Schary, Edwin G. 2005. In Seach of the Mahatmas of Tibet. Varanasi: Pilgrims publishing (first published, London, 1937). Schlagentweit, Emile. 1863. Buddhism in Tibet, with an Account of the Buddhist Systems preceding it in India. Leipzig: f.a. Brockhaus. Schopen, Gregory. 1992. “Archaeology and Protestant Presuppositions in the Study of Indian Buddhism.”History of Religions 31.4; 1–23. , 1994. “Stūpa and Tīrtha: Tibetan Mortuary Practices and an Unrecognised Form of Burial Ad Sanctos at Buddhist Sites in India.” In The Buddhist Forum iii. Tadeusz Skorupski & Ulrich Pagel (eds.) London: soas; 273–293. , 1997. Bones, Stones, and Buddhist Monks. Hawai’i: University Press. Schoterman, Jan A. 1992. “The Kubjikā Upaniṣad and Its Atharvavedic Character.” In Ritual and Speculation in Early Tantrism: Studies in Honor of André Padoux. Teun Goudriaan (ed.) n.y.: suny; 313–326. Schrempf, Mona. 1999. “Taming the Earth, Controlling the Cosmos: Transformation of Space in Tibetan Buddhist and Bon-po Ritual Dance.” In Sacred Spaces and Powerful Places in Tibetan Culture: A Collection of Essays. Toni Huber (ed.) Dharamsala: ltwa; 198–224. , 2007. “Bon Lineage Doctors and the Local Transmission of Knowing Medical Practice in Nagchu.” In Soundings in Tibetan Medicine: Anthropological and Histori- cal Perspectives. piats 2003: Tibetan Studies: Proceedings of the Tenth Seminar of the International Association for Tibetan Studies, Oxford, 2003. Mona Schrempf (ed.) Lei- den: Brill; 91–126. Schubring, Walther. 2000. The doctrine of the Jainas; described after the old sources. Delhi: Motilal Banarsidass (first published, Berlin, 1935). Schwartzberg, J.E. 1994. “Cartography of Greater Tibet and Mongolia.” In The History of Cartography vol. 2. 2: Cartography in Traditional East and Southeast Asian Societies. J.B. Harley and D. Woodward (eds.) Chicago: University Press; 207–237. Schwarzgruber, Rudolph. 1939. “The German Expedition to the Gangtori Glacier, 1938.” Himalayan Journal xi; 138–148. bibliography 497

, 1939a. “The German Garwhal-Himalaya Expedition, 1938.” Alpine Journal li; 79–84. Schwieger, Peter. 1996. “sTag-tshang ras-pa’s exceptional life as a pilgrim.”Kailash xviii: i–ii; 81–108. Seeland, Klaus. 1993. “Sanskritisation and Environmental Perception Among the Ti- beto-Burman Speaking Groups.” In Anthropology of Tibet and the Himalaya. Charles Ramble and Martin Brauen (eds.) Zurich: Ethnological Museum of the University of Zurich; 354–364. Selvanayagam, Israel. 1992. “Aśoka and Arjuna as Counterfigures standing on the field of dharma: A historical hermeneutical perspective.” History of Religions 32.1; 59– 75. Sen, Tansen. 2003. Buddhism, diplomacy and trade: the realignment of Sino-Indian rela- tions, 600–1400. Honolulu: University of Hawai’i Press. Serrano, Miguel. 1963. The Serpent of Paradise. The Story of an Indian Pilgrimage. Lon- don: Rider & co. Shakabpa, Tsepon Wangchuk Deden. (Derek F. Maher, trans.) 2010. One Hundred Thou- sand Moons: An Advanced Political History of Tibet, 2 vols. Leiden: Brill. Sharma, B.R. 1990. “The Institution of the Village Gods in the Western Himalayas.” In Himalayan Environment and Culture. N.K. Rustomji and Charles Ramble (eds.) Shimla: Indian Institute of Advanced Studies; 131–140. , 2003. “Sāṅchā Vidyā: A Traditional Astro-Tāntric System of Knowledge.” In Where Mortals and Mountain Gods Meet: Society and Culture in Himachal Pradesh. Laxman Thakur (ed.) Shimla: Indian Institute of Advanced Studies; 355–370. Sharma, B.R. & A.R. Sankhyan. 1996. People of India: Himachal Pradesh vol. xxiv. Anthro- pological Survey of India. Delhi: Manohar. Sharma, Kamal Prasad. 2001. Maṇimahesh Chambā Kailāsh. New Delhi: Indus. Sharma, Mahesh. 1995. “Concentric Rings of Pilgrimage: Local, Regional and Subconti- nental Linkages.” Studies in Humanities and Social Sciences ii.i; 97–109. , 1996. “Marginalisation and appropriation: Jogis, Brahmins and Sidh shrines.” ieshr 33.1; 73–91. , 2001. The Realm of Faith: Subversion, Appropriation and Dominance in the Western Himalaya. Shimla: Indian Institute of Advanced Studies. , 2004. “State formation and cultural complex in western Himalaya: Chamba genealogy and epigraphs—700–1650c.e.”ieshr 41.4; 389–432. , 2008. “Puranic Texts from Kashmir: Vitastā and River Ceremonials in the Nīlamata Purāṇa.” South Asia Research 28.2; 123–145. , 2009. Western Himalayan Temple Records: State, Pilgrimage, Ritual and Legal- ity. Leiden: Brill. , 2012. “State, Pastures and Rice-fields: The Gaddi Shepherds of Colonial Hima- chal Himalayas (North India).” Man in India 92.1; 13–35. 498 bibliography

Sharma, Y.D. 1955–1956. “Past Patterns of Living as Unfolded by Excavations at Rupar.” Lalat Kala 1–2; 121–129. Shashi, S.S. 1979. The Nomads of the Himalayas. Delhi: Sundeep Prakashan. Shastri, Lobsang. 2002. “Activities of Indian Panditas in Tibet from the 14th to the 17th Century.” In Tibet Past and Present: Tibetan Studies i. the Proceedings of the 9th International Seminar for Tibetan Studies, Leiden 2000. Henk Blezer (ed.) Leiden: Brill; 129–145. , 2003. “The Fire Dragon Chos ‘Khor (1076ad).” In The History of Tibet vol. ii. The Medieval Period. The Development of Buddhist Paramountcy. Alex McKay (ed.) London: RoutledgeCurzon; 174–184, (first published in Tibetan Studies: Proceedings of the 7th Seminar of the International Association for Tibetan Studies, vol. 1, Graz 1995. Ernst Steinkellner et al., [eds.] 1997. Wien: vöaw; 873–882). , 2009. “Jalandhara in the Eyes of Tibetan and Trans-Himalayan Pilgrims.” The Tibet Journal 34.2; 3–34. Sherring, C.A. 1974. Western Tibet and the Indian Borderland. Delhi: Cosmo (first pub- lished as Western Tibet and the British Borderland: the sacred country of Hindus and Buddhists …, 1906 [not 1916 as indicated in reprint.]) London: Edward Arnold. Sherring, M.A. 1859. The Indian Church during the Great Rebellion. London: James Nisbet. , 1868. Benares, the sacred city of the Hindus in ancient and modern Times. London: Trubnor. , 1872–1879. Hindu tribes and Castes 3 vols. London: Trubner. Shiraishi, R. 1996. AsceticisminBuddhismandBrahmanism:AComparativeStudy. Tring: The Institute of Buddhist Studies. Shneiderman, Sarah. 2002. “Embodied Ancestors: Territory and the Body in Thangmi Funerary Rites.” In Territory and Identity in Tibet and the Himalayas, piats 2000: Tibetan Studies: Proceedings of the Ninth Seminar of the International Association for Tibetan Studies, Leiden 2000. Katia Buffetrille & Hildegaard Diemberger (eds.) Leiden: Brill; 233–252. Shulman, David. 1980. Tamil Temple Myths: Sacrifice and Divine Marriage in the South Indian Saiva Tradition. Princeton: University Press. Sihlé, Nicolas. 2009. “The Ala and Ngakpa Priestly Traditions of Nyemo (Central Tibet): Hybridity and Hierarchy.” In Buddhism Beyond the Monastery: Tantric Practices and their Performers in Tibet and the Himalayas. Sarah Jacoby and Antonio Terrone (eds.) Leiden: Brill; 145–162. Siklós, Bulcsu. 1993. “Datura Rituals in the Vajramahabhairava-Tantra.” Curare 16: 71–76. , 1996. The Vajrabhairava Tantra: Tibetan and Mongolian Versions, English Translation and Annotations. Tring: Institute of Buddhist Studies. Singh, Gurdial. 1955. “Three Months in Upper Garhwal and Adjacent Tibet.” Alpine Journal lx; 257–269. bibliography 499

Singh, Jogishwar. 1989. Banks, Gods and Government: Institutional and Informal Credit Structures in a Remote and Tribal Indian District (Kinnaur, Himachal Pradesh) 1960– 1985. Stuttgart: J. Steiner Verlag. , 1990. “A brief survey of village gods and their moneylending operations in Kinnaur district of Himachal Pradesh; along with earlier importance of trade with Tibet.” In Wissenschafts-geschichte und gegenwärtige Forschungen in Nordwest- Indien. Lydia Icke-Schwalbe and Gudrun Meier (eds.) Dresden: Staatliches Museum für Völkerkunde; 244–255. Singh, R. Raj. 2006. Bhakti and Philosophy. Plymouth; Lexington Books. Singh, Raghubir. 1982. “Pilgrims return to Kailas, Tibet’s sacred mountain.”Smithsonian May; 94–103. Singh, Ranvir. 1979. “Raktavarn Glacier and Ascent of Unnamed Virgin Peaks.” Himala- yan Journal 37; 81–87. Singh, S.P. 1995. “Origin and growth of the Institution of Pilgrimage.” In Pilgrimage Studies: Sacred Places, Sacred Traditions. D.P. Dubey (ed.) Allahabad: The Society of Pilgrimage Studies. Singh, Yunas. 1917. “On the Roof of the World: Visits to Tibet.” Chronicles of the London Missionary Society (2 parts) April 1917; 66–69: May 1917; 81–84, (reprinted in 2003.The History of Tibet, vol. 3. The Modern Period 1895–1959. The Encounter with Modernity. Alex McKay [ed.] London: RoutledgeCurzon; 501–508). Sircar, D.C. 1973. The Śākta Pīṭhas. Delhi: Motilal Banarsidass, 2nd revised edition. , 1977. Studies in the geography of ancient and medieval India. Delhi: Motilal Banarsidass, (first published, 1971). Skinner, Thomas. 1833. Excursions in India: including a walk over the Himalaya Moun- tains, to the sources of the Jumna and the Ganges, 2 vols. London: Richard Bentley (second edition). Smith, Brian K. 1986. “Ritual, Knowledge and Being: Initiation and Veda Study in Ancient India.” Numen 33.1; 65–89. Smith, D.H. 1968. Chinese religions. London: Weidenfeld & Nicolson. Smith, Gene. 2001. Among Tibetan Texts: History and Literature of the Himalayan Pla- teau. Boston: Wisdom Publications. Smith, W.L. 1980.The One-Eyed Goddess: A Study of the Manasā Mangal. Stockholm: Almqvist & Wikall. Snellgrove, David. 1959. “The Notion of Divine Kingship in Tantric Buddhism.” Numen. (Supplement: La regalita Sacra / The Sacral Kingship: Contributions to the Central Theme of the viiith International Congress for the History of Religions, Rome, April 1955.) Leiden: Brill; 204–218. , 1976. The Hevajra Tantra: A Critical Study. London: oup (first published, 1959). , 1982. “Buddhism in North India and the Western Himalayas: Eleventh to Thirteenth Centuries.” In The Silk Road and the Diamond Path: Esoteric Buddhist Art 500 bibliography

on the Trans-Himalayan Trade Routes. Deborah Klimburg-Salter (ed.) Los Angeles: ucla Arts Council; 64–80. , 1987. Indo-Tibetan Buddhism: Indian Buddhists and their Tibetan Successors. London: Serindia. Snellgrove, David, and Hugh Richardson. 1968. A Cultural History of Tibet. London: Weidenfeld & Nicolson. Snelling, John. 1990. The Sacred Mountain: A Complete Guide to Tibet’s Mount Kailas (revised edition). London/The Hague: East-West Publications. Snyder, Gary. 2007. Passage Through India. Emeryville: Shoemaker and Hoard (first published, 1972). Söhnen, Renate, & Peter Schriener. 1989. Brahmapurāṇa: Summary of Contents, with Index of Names and Motifs. Wiesbaden: Otto Harrassowitz. Sorensen, Soren. 1904. An Index to the Names in the Mahabharata. London: Willliams and Norgate. Spengen, Wim van. 2000. Tibetan Border Worlds: A Geohistorical Analysis of Trade and Traders. London/n.y.: Kegan Paul International. Sperling, Elliot. 1990. “Hulegu and Tibet.” Acta Orientalia Academiae Scientiarum Hun- garicae xliv.i–ii; 145–157. , 2003. “Some notes on the early ’Bri-gung-pa Sgom-pa.” In The History of Tibet vol. ii. The Medieval Period. The Development of Buddhist Paramountcy. Alex McKay (ed.) London: RoutledgeCurzon; 373–392 (first published in Silver on Lapis: Tibetan Literary Culture and History. Christopher I. Beckwith [ed.]. 1987. Bloomington: The Tibet Society; 33–53). Spink, Walter M. 2005. Ajanta, vol. 1, History and Development. Leiden: Brill. Spitzer, A. 1973. “The Historical Problem of Generations.” American Historical Review 78.5; 353–385. Srinivas, M.N. 1989. The Cohesive Role of Sanskritization and Other Essays. Delhi: oup. Staal, Frits. 1963. “Sanskrit and Sanskritization.” Journal of Asian Studies 22.3; 261– 275. , 1982. “The Himalayas and the Fall of Religion.” In The Silk Road and the Diamond Path: Esoteric Buddhist Art on the Trans-Himalayan Trade Routes. Deborah Klimburg-Salter (ed.) Los Angeles: ucla Arts Council; 38–64. , 1982a. “What is Happening in Classical Indology?—A Review Article.” Journal of Asian Studies xli.ii; 269–291. , 1990. “The Lake of the Yaksa Chief.” In Indo-Tibetan Studies; Papers in honour and appreciation of Professor David L. Snellgrove’s contribution to Indo-Tibetan Stud- ies. Tadeusz Skorupski (ed.) Tring: The Institute of Buddhist Studies; 275–292. , 2003. “Three Mountains and Seven Rivers.” In Indian Culture and Buddhism. Felicitiation Volume for Prof. Musashi Tachikawa. Shoun Hino and Toshihiro Wada (eds.) Delhi: Motilal Banarsidass; 1–24. bibliography 501

Stablein, M. 1978. “Textual and Contextual Patterns of Tibetan Buddhist Pilgrimage in India.” Tibet Society Bulletin 12; 7–38. Stearns, Cyrus. 2010. The Buddha from Dolpo: A Study of the Life and Thought of the Tibetan Master Dölpopa Sherab Gyaltsen. Ithaca: Snowlion (first published, 1989). Stein, Rolf A. 1972. Tibetan Civilisation. Standford: University Press (first published as La Civilisation Tibétaine. 1962. Paris). , 1979. “Religious Taoism and Popular Religion from the Second to Seventh Centuries.” In Facets of Taoism: Essays in Chinese Religion. Holmes Welch and Anna Seidel (eds.) New Haven and London: Yale University Press; 53–82. , 1990. The World in Miniature: Container Gardens and Dwellings in Far Eastern Religious Thought. Stanford: University Press (first published in French, 1987). , 1995. “La soumission de Rudra et autres contes tantriques.” Journal Asiatique 283.1; 121–160. Steinmann, B., 1996. “Mountain deities, the invisible body of society—A comparative study of the representations of mountains by the Tamang and the Thami of Nepal, the Lepcha and Bhutia of Sikkim.” In Reflections of the Mountain: Essays on the His- tory and Social Meaning of the Mountain Cult in Tibet and the Himalaya. Ann-Marie Blondeau & Ernst Steinkellner (eds.) Wien: vöaw; 179–218. Stewart, E. 1897. “Account of the Hindu Fire-Temple at Baku, in the Trans-Caucasus Province of Russia.” jras 29; 311–315. Stoddard, R.H. 1966. “Hindu Holy Sites in India.” University of Iowa Ph.D. thesis. Strachey, Henry. 1848. Narrative of a journey to the lakes Cho Lagan, or Rakas Tal, and Cho Mapan, or Manasarowar, and the valley of Pruang, in Tibet, in September and October, 1846. Calcutta: Baptist Mission Press, (also Journal of the Asiatic Society of Bengal, July-August-September, 1848). Strickman, Michael. 2002. (Bernard Faure, ed.) Chinese Magical Medicine. Standford: University Press. Stutchbury, Elisabeth. 1999. “Perceptions of Landscape in Karzha: ‘Sacred’ Geography and the Tibetan System of ‘Geomancy.’” In Sacred Spaces and Powerful Places in Tibetan Culture: A Collection of Essays. Toni Huber (ed.) Dharamsala: lwta; 154–186 (first published in 1994. The Tibet Journal 19.4; 59–102). Sumption, Jonathan. 1975. Pilgrimage: an image of medieval religion. London: Faber. Sundaranand, Swami. 2001. Himalaya: Through the Lens of a Sadhu. Gangotri: Tapovan Kuti Prakashan. Suryavanshi, Bhagwan S. 1986. The Geography of the Mahabharata. New Delhi: Rama- nand Vidya Bhawan. Susumu, Otake. 2007. “On the Origin and Early Development of the Buddhāvataṃsa- kasūtra.” In Reflecting Mirrors: perspectives on Huayan Buddhism. Imre Hamar (ed.) Wiesbaden: Otto Harrassowitz; 86–107. Sutherland, Gail Hinich. 1992. Yakṣa in Hinduism and Buddhism: The Disguises of the Demon. Delhi: Manohar (first published 1991. n.y.: suny). 502 bibliography

Sutherland, Peter. 1998. “Travelling Gods and Government by Deity: An Ethnohistory of Power, Representation and Agency in West Himalayan Polity.” Oxford University D.Phil. thesis. , 2004. “Local Representations of History and the History of Local Represen- tation: Timescapes of Theistic Agency in the Western Himalayas.” ebhr 25/26; 80– 118. Svoboda, Robert E. 1986. Aghora. At the Left Hand of God. Albuquerque: Brotherhood of Life Publications. Swearer, Donald K., Sommai Premchit, & Phaithoon Dokbuakaew. 2004. Sacred Moun- tains of Northern Thailand and their legends. Silkworm Books: Chiang Mai. Sweet, Michael J. (trans.) & Leonard Zwilling (ed.) 2010. Mission to Tibet. The Extraordi- nary Eighteenth Century Account of Father Ippolitio Desideri, s.j. Boston: Wisdom. Takeuchi, Tsuguhito, Burkhard Quessel & Yasuhiko Nagano. (eds.) 2011. Research Notes on the Zhangzhung Language by Frederick W. Thomas at the British Library. (Bon Studies 14) Osaka: National Museum of Ethnology. Tambiah, Stanley J. 1976. World Conqueror and World Renouncer: A study of Buddhism and Polity in Thailand against a Historical Background. Cambridge: cup. . 1985. “The Galactic Polity in Southeast Asia.” In Culture, Thought, and Social Action: An Anthropological Perspective. Stanley Tambiah. Cambridge ma: Harvard University Press; 252–286. Tapovan, Swami. (Thapovanam, Gurudev). 1984. Hymn to Badrinath. Bombay: Central Chinmaya Mission Trust. , 2001. Kailas Yatra. Mumbai: Central Chinmaya Mission Trust (first published in Malayalam, 1927). , 2003. Wanderings in the Himalayas. Mumbai: Central Chinmaya Mission Trust (first revised edition, 1990). Tautscher, Gabriele. 2007. Himlayan Mountain Cults: Sailung, Kalingchok, Gosainkhund. Territorial Rituals and Tamang Histories. Kathmandhu: Vajra Publications. Taylor, Kathleen. 2001. Sir John Woodroffe, Tantra and Bengal. ‘An Indian Soul in a European Body’?. Richmond: Curzon. Teltscher, Kate. 2006. The High Road to China: George Bogle, The Panchen Lama and the First British Expedition to Tibet. London: Bloomsbury. Templeman, David. 1994. “Internal and External Geography in Spiritual Biography.” The Tibet Journal 19.3; 63–76. , 1997. “Buddhaguptanātha: a Late Indian siddha in Tibet.” In Tibetan Studies, Proceedings of the 7th Seminar of the International Association for Tibetan Studies, Graz 1995, vol. ii. Helmut Krasser, Michael Torsten Much, Ernst Steinkellner, Helmut Tauscher. (eds.) Wien: vöaw; 955–965. , 2002. “Iranian themes in Tibetan Tantric Culture: The Ḍākinī.” In Religion and Secular Culture in Tibet. Tibetan Studies ii. piats 2000. Tibetan Studies: Proceedings bibliography 503

of the Ninth Seminar of the International Association for Tibetan Studies, Leiden 2000. Henk Blezer (ed.) Leiden: Brill; 113–128. , 2008. “Becoming Indian: A Study of the Life of 16–17th century Tibetan Lama, Tāranātha.” Monash University Ph.D. Thesis. , (ed.) 2010. New Views of Tibetan Culture. Caulfield: Monash University Press. Thakur, Laxman. 1989. “A Note on the Nandi Inscription and the Date of Mani Mahe- sha Temple, Bhramaur.” In The Western Himalayan Kingdom: History and Culture of the Chamba State. V.C. Ohri and A.N. Khanna (eds.) Delhi: Books and Books; 155– 160. , 2001. Buddhism in the Western Himalaya: A Study of the Tabo Monastery. Delhi: oup. , 2002. “Exploring the Hidden Buddhist Treasures of Kinnaur (Khu Nu): A Study of the Lha Khang Chen mo, Ribba.” In Buddhist Art and Tibetan Patronage: Ninth to Fourteenth Centuries, piats 2000: Proceedings of the Ninth Seminar of the International Association for Tibetan Studies, Leiden 2000. Deborah Klimburg Salter & Eva Allinger (eds.) Leiden: Brill; 29–44. , (ed.) 2003. Where Mortals and Mountain Gods Meet: Society and Culture in Himachal Pradesh. Simla: Indian Institute of Advanced Study. , 2011. “Rin chen Bzang po’s Footsteps: Re-Evaluating the contribution of the Great Translator through folklore and archaeology.” In Buddhist Himalaya: Studies in Religion, History And Culture. Proceedings of the Golden Jubilee Conference of the NamgyalInstituteofTibetology,Gangtok,2008:Volume1:TibetandtheHimalaya. Alex McKay & Anna Balikci-Denjongpa (eds.) Gangtok: nit; 209–218. Thakur, M.R. 1997. Myths, Rituals and Beliefs in Himachal Pradesh. Delhi: Indus. , 2003. “Some Cultural Legends and Traditions of the Kulu Valley.” In Where Mortals and Mountain Gods Meet: Society and Culture in Himachal Pradesh. Laxman S. Thakur (ed.) Simla: Indian Institute of Advanced Study; 321–333. Thakur, U. 1973. “The Holy Places of North India as Mentioned in the Skanda-Purana.” Purana 15.2; 93–120. Thapar, Romila. 1978. Ancient Indian Social History: Some Interpretations. New Delhi: Orient Longman. , 2000. “The Householder and the Renouncer in the Brahmanical and Buddhist Traditions.” In Cultural Pasts: Essays in Early Indian History. Romila Thapar, Delhi: Oxford University Press, (first published, 1981. Contributions to Indian Sociology, 15.1–2; 272–298). , 2000. “The Mouse in the Ancestry.” In Cultural Pasts, Essays in Early Indian History. Romila Thapar. Delhi: oup; 797–806 (first published in Amrtadhārā: Prof. R.N. Dandekar Felicitation Volume. S.D. Joshi (ed.) 1984. Delhi: oup; 427–434). Thar, Tsering. 2002. “Shar rdza Hermitage: A New Bonpo Center in Khams.” In Khams pa Histories: Visions of People, Place and Authority. piats 2000: Tibetan Studies: Proceed- 504 bibliography

ings of the Ninth Seminar of the International Association for Tibetan Studies, Leiden, 2000. Lawrence Epstein (ed.) Leiden: Brill; 155–172. , 2009. “Mount Ti se (Kailash) Area: The Center of Himalayan Civilisation.” In Bon The Everlasting Religion Of Tibet: Tibetan Studies in Honour of Professor David L. Snellgrove, New Horizons in Bon Studies 2. Samten G. Karmay and Donatella Rossi (eds.) (Special edition of) East and West 59.1–4; 25–29. Thomas. F.W. 1935–1963. Literary Texts and Documents Concerning Chinese Turkestan, 4 vols. London: Royal Asiatic Society. , 1957. Ancient Folk-Literature from North-Eastern Tibet. Introduction, Texts, Translations and Notes. Berlin: Akademie-Verlag. Thompson, Phyllis. 2002. Sadhu Sundar Singh: A new biography of the remarkable Indian holy man and disciple of Jesus Christ. Secunderabad: Om Books (first published, 1994. Carlisle). Thurman, Robert, and Tad Wise. 2000. Circling the Sacred Mountain: A Spiritual Adven- ture Through the Himalayas. n.y.: Bantam books: (first published, 1999). Tichy, Herbert. 1938. Tibetan adventure: travels through Afghanistan, India and Tibet. London: Faber & Faber. Tiso, Francis V. 1994. “The Rdo Rje ‘Chang Rnam Thar in the Bka’ Brgyud Gser ‘Phreng Genre.” In Tibetan Studies: Proceedings of the 6th Seminar of the International Asso- ciation for Tibetan Studies Fagernes 1992, 2 vols. Per Kværne (ed.) Oslo: icrhc; vol. 2; 884–888. Tolia, R.S. 1994. British Kumaun-Garhwal. An Administrative History of a Non-regulation Hill Province. Gardner and Traill Years (1815. a.d.–1835a.d.). Almora: Shree Almora Book Depot. Tomas, Andrew. 1974. Shambhala: Oasis of Light. London: Sphere Press. Touw, Mia. 1981. “The Religious and Medical Uses of Cannabis in China, India and Tibet.” Journal of Psychoactive Drugs 13.1; 23–34. Trautman, Thomas. 1971. Kautilya and the Arthasastra: A Statistical Investigation of the Authorship and Evolution of the Text. Leiden: Brill. , 1997. Aryans and British India. Berkeley/London: University of California Press. Tripathi, B.D. 2007. Saddhus of India: The Sociological View. Varanasi: Pilgrims’ Publish- ing (first published, Bombay, 1978). Tsuda, Shinichi. 1974. The Samvarodaya-Tantra: Selected Chapters. Tokyo: Hokuseido Press. , 1978. “A Critical Tantrism.” The Memoirs of the Toyo Bunko 36; 167–231. , 1982. ‘“Vajrayosidbhagesu Vijahara’: Historical Survey from the Beginnings to the Culmination of Tantric Buddhism.” In Indological and Buddhist Studies: Volume in Honour of Professor J.W. de Jong on his Sixtieth Birthday. L.A. Hercus et al. (eds.) Canberra: Australian National University; 595–616. bibliography 505

Tucci, Giuseppi. 1940. Travels of Tibetan Pilgrims in the Swat Valley. Calcutta: Greater India Society. , 1949. Tibetan Painted Scrolls, 2 vols. Roma: Libreia Dello Stato. , 1955. “The Sacral Character of the Kings of Ancient Tibet.” East and West 4; 197–205. , 1956. Preliminary Report on Two Scientific Expeditions in Nepal. Rome: IsMEO. , 1962. Nepal: the discovery of the Malla. London: George Allen & Unwin (first published in Italian, 1960). , 1970. The Theory and Practice of the Mandala. New York; Samuel Wiesner. , 1971. “A propos Avalokiteśvara.” In Opera Minora, 2 vols. Giuseppi Tucci. Rome: G. Bardi; 489–509. , 1971a. “Himalayan Cīna.” In Etudes tibétaines dediées à la mémoire de Marcelle Lalou, Paris: Libraire d’Amerique et d’Orient; 548–552. , 1977. “On Swāt, the Dards, and Connected Problems.”East and West 27.19–104. , 1980. TheReligionsofTibet. (Geoffrey Samuel: trans.) London: Routledge Kegan Paul (first published in German, 1970). , 1988. Rin-Chen-Bzaṅ-Po and the Renaissance of Buddhism in Tibet Around the Millenium (Indo-Tibetica ii). New Delhi: Aditya Prakashan (first published in Italian, Rome, 1932). Tucci, Giuseppi and Eugenio Ghersi. 1935. Secrets of Tibet: Being a Chronicle of the Tucci Scientific Expedition to Western Tibet (1933). London/Glasgow: Blackie & Son. Turin, Mark. 2002. “Ethnonyms and other-Nyms: Linguistic Anthropology among the Thangmi of Nepal.” In Territory and Identity in Tibet and the Himalayas. Katia Buf- fetrille and Hildegard Diemberger (eds.) Leiden: Brill; 253–270. Turner, Victor. 1973. “The Center Out There: Pilgrim’s Goal.” History of Religions 12; 191–230. Turner, Victor, and E. Turner. 1978. Image and Pilgrimage in Christian Culture. n.y.: Columbia University Press. Urban, Hugh. 1995. “The Remnants of Desire: Sacrificial Violence and Sexual Transgres- sion in the Cult of the Kapalikas and in the Writings of Georges Bataille.”Religion 25; 67–90. , 1995. “The Strategic Uses of an Esoteric Text: The Mahānirvāṇa Tantra.” South Asia 18.1; 55–81. , 2003. Tantra. Sex, Secrecy, Politics and Power in the Study of Religion. Berkeley: University of California Press. Uray, Géza. 1988. “Ñag.ñi.dags.po: A Note on the Historical Geography of Ancient Tibet.” In Orientalia Iosephi Tucci memoriae dicata vol. 3. G. Gnoli & L. Lanciotti (eds.) Rome: IsMEO; 1503–1510. , 2003. “Queen Sad-ma-kar’s Songs in the Old Tibetan Chronicle.” In The History of Tibet. Vol. 1, The Early Period to c. ad 850. The Yarlung Dynasty. Alex McKay (ed.) 506 bibliography

London: RoutledgeCurzon; 245–272, (first published 1972. Acta Orientalia Hungarica 25.3; 5–38). Urbanska-Szymoszyn, Anna. 2011. “Tranforming Tibetan Icon. Chinese Impact and Global Implications on the Picture of Mount Kailas.” The Tibet Journal 36.4: 17–39. Valantasis, R. 1995. “Constructions of Power in Asceticism.” jaar lxiii.iv; 775–822. Vargas, Ivette. 2010. “Legitimising demon diseases in Tibetan medicine: the conjoining of religion, medicine, and ecology.” In Studies in Medical Pluralism in Tibetan History and Society. Sienna Craig, Mingji Cuomu, Frances Garrett, and Mona Schrempf (eds.) Halle: iittbs; 379–404. Varma, Rommell. 1988–1989. “Meru and Kailasa: Reflections on a Tangible ‘Sacred Space’.” London University (soas) m.a. thesis. Veer, Peter van der. 1988. Gods on Earth: The Management of Religious Experience and Identity in a North Indian Pilgrimage Centre. London: Athlone. Vitali, Roberto. 1996. The Kingdoms of Gu.ge Pu.hrang: According to the mNga’.ris rgyal. rabs by Gu.ge mkhan.chen Ngag.dbang grags-pa. Dharamsala: lwta. , 2003. “A Chronology (bstan rtsis) of Events in the History of mNga’ ris Skor gSum (Tenth-Fifteenth Centuries).” In TheHistoryofTibetvol.ii.TheMedievalPeriod. The Development of Buddhist Paramountcy. Alex McKay (ed.) London: Routledge- Curzon; 53–89. Vogel, J.Ph. 1926. Indian Snake Lore. London: Probsthain. Vohra, Rohit. 1983. “History of the Dards and the Concept of Minaro Traditions among the Buddhist Dards of Ladakh.” In Recent Research on Ladakh: History, Culture, Sociology, Ecology. D. Kantowsky & R. Sander (eds.) Köln: Weltforum Verlag; 51– 80. , 1989. The Religions of the Dards in Ladakh. Investigations into their pre-Buddhist ‘Brog-pa Traditions. Luxembourg: Skydie Brown International. Vreeland, Herbert H. 1957. Mongolian Community and Kinship structure. New Haven: Human Relations Area Files. Waddell, L.A. 1895. The Buddhism of Tibet, or Lamaism: with its mystic cults, symbolism and mythology, and in its relation to Indian Buddhism. London: W.H. Allen. Wadia, D.N. 1968. “The Himalaya Mountains: Its Origin and Geographical Relations.” In Mountains and Rivers of India. B.C. Law (ed.) Calcutta: National Committee for Geography; 36–40. Wakefield, E.W. 1996. Past Imperative: My Life in India 1927–1947. London: Chatto & Windus. Waller, Derek. 1990. The Pandits: British Exploration of Tibet and Central Asia. Lexington: University of Kentucky Press. Walter, Michael. 2009. Buddhism and Empire: The Political and Religious Culture of Early Tibet. Leiden: Brill. Walter, Michael, & Christopher Beckwith. 1997. “Some Indo-European Elements in bibliography 507

Early Tibetan Culture.” In Tibetan Studies: Proceedings of the 7th Seminar of the International Association for Tibetan Studies, Graz 1995, vol. ii. Ernst Steinkellner (ed.) Wien: vöaw; 1037–1054. Walton, H.G. (ed.) 1910. Dehradun. A Gazetteer. Allahabad; Government Press. , (ed.) 1910. British Garhwal. A Gazetteer, being Vol. xxxv of the District Gazetteers of the United Provinces of Agra and Oudh. Allahabad: Government Press. , (ed.) 1911. Almora: A Gazetteer, being Vol. xxxvi of the District Gazetteers of the United Provinces of Agra and Oudh. Allahabad: Government Press. Wangdu, Pasang, and Hildegard Diemberger. (ed./trans.) 2000. dBa’ bzhed: The Royal Narrative Concerning the Bringing of the Buddha’s Doctrine to Tibet. Wien: vöaw. Warner, Karel. 1989. “The Longhaired Sage of rv 10.136: A Shaman, a Mystic or a Yogi?”In The Yogi and the Mystic: Studies in Indian and Comparative Mysticism. Karel Warner (ed.) Richmond: Curzon Press; 33–53. Wasson, R. Gordon. 1968. Soma, divine mushroom of immortality. The Hague/n.y.: Har- court. Waterhouse, David. 2004. “Brian Henry Hodgson—A Biographical Sketch.” In The Ori- gins of Himalayan Studies: Brian Henry Hodgson in Nepal and Darjeeling 1820–1858. David M. Waterhouse (ed.) Abingdon: RoutledgeCurzon; 1–25. Watson, Francis. 1961. “Pilgrims to Badrinath.” Geographical Journal xxxiv.vii; 421–428. Wayman, Alex. 1990. “Messengers, What Bring Ye?” In Indo-Tibetan Studies; Papers in honour and appreciation of Professor David L. Snellgrove’s contribution to Indo- Tibetan Studies. Tadeusz Skorupski (ed.) Tring: The Institute of Buddhist Studies; 305–322. Wedemeyer, Christian. 2001. “Tropes, Typologies, and Turnarounds: A Brief Geneology of the Historiography of Tantric Buddhism.”History of Religions 40; 223–259. Wessels, C.J. 1924. Early Jesuit Travellers in Central Asia. The Hague: Mouton. White, David Gordon. 1996. The Alchemical Body: Siddha Traditions in Medieval India. Chicago: University Press. , 1997. “Mountains of wisdom: On the interface between Siddha and Vidyādhara cults and the Siddha orders in medieval India.”International Journal of Hindu Studies 1.1; 73–95. , 1998. “Transformations in the Art of Love: Kāmakalā Practices in Hindu Tantric and Kaula Traditions.”History of Religions 38.2; 172–198. , (ed.), 2001. Tantra in Practice. Delhi: Motilal Banarsidass (first published, Princeton, 2000). , 2003. Kiss of the Yogini: “Tantric Sex” in its South Asian Contexts. Chicago: University Press. , 2009. Sinister Yogis. Chicago/London: University of Chicago Press. Wilford, Francis. 1801. “On Mount Caucasus.” Asiatick Researches 6; 455–539. Wilson, Andrew. 1876. The Abode of Snow: Observations on a journey from Chinese 508 bibliography

Tibet to the Indian Caucasus through the upper valleys of the Himalaya. London: Blackwoods, 2nd edition (first published, 1875). Wilson, H.H. 1958. Religious Sects of the Hindus. Calcutta: Susil Gupta (first published as A Sketch of the Religious Sects of the Hindus. 1862. London: Trubner). Winkler, Ken. 1990. A Thousand Journeys: The Biography of Lama Anagarika Govinda. Longmead, Shaftesbury: Element Books. Witzel, Michael. 1980. “Early Eastern Iran and the Atharvaveda.”Persica ix; 86–128. , 1993. “Nepalese Hydronomy: Towards a history of settlement in the Himalayas.” In Nepal, past and present: Proceedings of the Franco-German Conference at Arc-et- Senans, June 1990. Gérard Toffin (ed.) Paris: cnrs; 217–266. (accessed 24 October 2009 at ww.people.fas.harvard.edu/~witzel/mwbib.htm). , 1995. “Early Indian history: Linguistic and textual parameters.” In Language, Material Culture and Ethnicity: The Indo-Aryans of Ancient South Asia. George Erdosy (ed.) Berlin/n.y.: Walter de Gruyter; 85–125. , 1995a. “Rgvedic history: poets, chieftains and polities.” In Language, Material Culture and Ethnicity: The Indo-Aryans of Ancient South Asia. George Erdosy (ed.) Berlin/New York: de Gruyter; 307–352. , 1995b. “Early Sanskritization. Origins and Early Development of the Kuru State.”Electronic Journal of Vedic Studies, 1.4; 1–26. , 1997a. “The Development of the Vedic Canon and its Schools: The Social and Political Milieu.” In Inside the Texts Beyond the Texts: New Approaches to the Study of the Vedas. Michael Witzel (ed.) Harvard Oriental Series: (Opera Minora, vol. 2); 257–345. , 1999. “Aryan and non-Aryan Names in Vedic India: Data for the linguistic situa- tion, c. 1900–500b.c.” In AryanandNon-AryaninSouthAsia:Evidence,Interpretation and Ideology. J. Bronkhorst & M.M. Deshpande (eds.) Harvard: Harvard Oriental Series (Opera Minora vol. 3); 337–404. , 2004. “The Rgvedic Religious System and its Central Asian and Hindukush Antecedents.” In The Vedas: Texts, Language and Ritual. Arlo Griffiths & Jan Houben (eds.) Groningen: Forsten; 581–636. , 2014. “Textual Criticism in Indology and in European philology during the 19th and 20th centuries.”Electronic Journal of Vedic Studies 21.3; 9–91. Wojtilla, G. 1984. “Notes on Popular Saivism and Tantra in Eleventh Century Kashmir (A study on Ksemendra’s Samayamatrka).” In Tibetan and Buddhist Studies commemo- rating the 200th anniversary of the Birth of Alexander Csoma de Koros, vol. 2. L. Ligeti (ed.) Budapest: Akademai Kiado; 381–389. Wright, Rita. 2010. The Ancient Indus: Urbanism, Economy and Society. Cambridge: cup. Wujastyk, Dominic. 2002. “Cannabis in Traditional Indian Herbal Medicine.” In Ayur- veda at the Crossroads of Care and Cure: Proceedings of the Indo-European Seminar bibliography 509

on Ayurveda held at Arrábida, Portugal, in November 2001. A. Salema (ed.) Lisbon and Puna: Centro de História de Além-Mar, Universidade Nova de Lisboa; 45–73. Wylie, Turrell V. 1962. The Geography of Tibet According to the ‘Dzam-gling-rgyas-bshad. Rome: IsMEO. , 1964. “Mar-pa’s Tower: Notes on Local Hegemons in Tibet.”History of Religions 3.2; 278–292. Yamamoto, Carl. 2009. “Vision and Violence: Lama Zhang and the Dialectics of Political Authority and Religious Charisma in Twelfth-Century Tibet”, University of Virginia Ph.D. thesis. (For published version, see Vision and Violence: Lama Zhang and the Politics of Charisma in Twelfth-Century Tibet. 2012. Leiden: Brill.) Yadav, Krishna. 1988. The Guru in the Valley of the Gods. Chandigarh: Valley of the Gods Publications. Zeisler, Bettina. 2011. “For Love of the Word: A New Translation of Pelliot Tibétain 1287, The Old Tibetan Chronicle, Chapter One.” In New Studies of the Old Tibetan Documents: Philology, History and Religion. Yoshiro Imaeda, et al. (eds.) Tokyo: Tokyo University of Foreign Studies; 97–216. Zysk, Kenneth G. 1991. Asceticism and healing in ancient India: medicine in the Buddhist monastery. n.y.: oup.

Cited Classical Indic Texts in Translation

The Atharvaveda Bloomfield, Maurice. 1897. Hymns of the Atharva-veda. Together with Extracts from the Ritual Books and the Commentaries. Oxford: Clarendon. Whitney, W.D. 1971. Atharva-veda-samhitā, 2 vols. New Delhi: Motilal Banarsidass (first published, Harvard 1905).

The Arthashastra Rangarajan, L.N. (ed./trans) 1992. Kautilya. The Arthashastra. New Delhi: Penguin Books, (first published, 1987).

The Abhidharmakosa Pruden, Leo M. (trans.) 1990. Abhidharmakosabhasyam of Vasubandhu. Berkeley: Asian Humanities Press (translated from the French; Louis de La Vallé Poussin. 1971. L’Abhidharmakośa de Vasubandhu. Bruxelles: Institut belge des hautes études chi- noises). 510 bibliography

The Jatakas [Originally published in 6 volumes by Cambridge University Press; series editor E.B. Cowell: available at www.sacred-texts.com.] The Jataka. vol. i: R. Chalmers (trans.) 1895. The Jataka. vol. ii: W.H.D. Rouse (trans.) 1895. The Jataka. Vol. iii: H.T. Francis & R.A. Neil (trans.) 1897. The Jataka. vol. iv: W.H.D. Rouse (trans.) 1901. The Jataka. Vol. v: H.T. Francis (trans.) 1905. The Jataka. Vol. vi: E.B. Cowell & W.H.D. Rouse (trans.) 1907.

The Mahabharata J.A.B. van Buitenen, (ed./trans.). 1973–1978. The Mahabharata. 3 vols. Chicago: Univer- sity Press.

The Mahāvastu Jones, J.J. (trans.). 1952.The Mahāvastu. 3 vols. London: Luzac & co.

Puranas Brahma Purāṇa. 3 vols. (translated by a committee) Delhi: Motilal Banarsidass; 1985– 1986. Brahmāṇḍa Purāṇa. 4 vols. G.V. Tagare (ed./trans.), Delhi: Motilal Banarsidass; 1983. Kūrma Purāṇa. 2 vols. G.V. Tagare (ed./trans.), Delhi: Motilal Banarsidass; 1982. Matsya Puranam. B.D. Basu (ed./trans.), Allahabad: The Panini Office; 1916. Skandapurāṇa. G.V. Tagare (ed./trans.), Delhi: Motilal Banarsidass; 1992. Siva-purana. 4 vols. Arnold Kunst & J.L. Shastri (eds./trans.), Delhi: Motilal Banarsidass; 1970. Varāha Purāṇa. S.V. Iyer (trans.) Delhi: Motilal Banarsidass; 1985. Vāyu Purāṇa. 2 vols. G.V. Tagare (ed./trans.) Delhi: Motilal Banarsidass; 1987.

The Rajatarangini Stein, Marc Aurel, (ed./trans.). 1900. Kalhana’s Rajatarangini, a chronicle of the Kings of Kasmir. 2 vols., Calcutta: Constable and Co. Pandit, Ranjit Sitaram, (ed./trans.). 1935. Rājaṭarañgiṇī: the saga of the kings of Kaśmīr / translated from the original Samskrt of Kalhaṇa and entitled the River of Kings with an introduction, annotation, appendices, index, etc. Allahabad: Indian Press.

The Ramayana Goldman, Robert P. (ed./trans.). 1984–1994. The Ramayana of Valmiki: an epic of ancient India. Princeton, n.j.: Princeton University Press. Vol. 1: Bālakāṇḍa / Introduction and translation by Robert P. Goldman; annotation bibliography 511

by Robert P. Goldman and Sally J. Sutherland. Princeton, n.j.: Princeton Univer- sity Press. 1984. Vol. 2: Ayodhyākāṇḍa / introduction, translation, and annotation by Sheldon I. Pol- lock; edited by Robert P. Goldman. Princeton, n.j.; Guildford: Princeton Univer- sity Press. 1986. Vol. 3: Araṇyakāṇḍa / introduction, translation, and annotation by Sheldon I. Pol- lock; edited by Robert P. Goldman. Princeton, n.j. (usa); Oxford: Princeton Uni- versity Press. 1991. Vol. 4: Kiṣkindhākāṇḍa / introduction, translation, and annotation by Rosalind Lefeber; edited by Robert P. Goldman. Princeton, n.j. (usa): Princeton University Press. c.1994.

Unpublished Documents

Kailas Manasarovar Yatra: Instructions for Yatris. (Information hand-out for Indian pilgrims. Produced by the Indian Ministry of External Affairs [East Asia Division], 1996 edition). Undated report (circa 1948) by British Political Officer, Captain R.K.M. Saker, “On the Future of the British Trade Agency”; (copy in the possession of the author, courtesy of Mrs A. Saker). Oriental and India Office Collections: cited documents. National Archives of India: cited documents. Index

Abhidharma/kośa 76, 79, 114, 120, 122, Anglo-Chinese Trade Regulations, 1893 405 124–126, 244n62, 318, 320, 353, 428 Anglo-Nepal war 142, 212 Abu, Mount 207, 433 Anglo-Russian Convention, 1907 405 Achaemenid empire 30, 51 Anglo-Sikh war 154, 179 A-chuk Tsering 335 Anglo-Tibetan Convention, 1904 387, 390, Adam (Biblical) 412 393, 404 Adhi Kailas, Mount 149, 190–195, 199, 201, Anotatta (lake) 85, 115–116, 120–124, 251, 334 126–128 Aesthetics (of landscape) 109–110, 393 See also Anavatapta, Manasarovar, and Afghanistan 87, 93, 381, 384 Rakas Tal Agartha Network 412 Arab sources 375, 382 Aghori (sect) 69, 243 Āranyakas 42, 44 Ahura Mazda (deity) 30 Arbuda 97 Airavati (river) 79 Arhats 115, 119 Ajanta 90 Arjuna 56–57, 61–64, 136, 222, 224 Ajay Pal, 203, 207, 208n34, 209 Arnold, Edward 397 Akhbar, Emperor 214 Arthaśāstra 94n13, 170, 289, 308–309 Alaka 89, 206 Ārya/Aryan 34, 444 Alakananda (river) 55, 66, 74, 80, 84, 141, 200, See also Indo-European 203–204, 206–207, 216, 218, 246 Arwa (river) 246 Alavaka, King 115, 116n11 Arya Samaj 227, 414 Alchemy/alchemical substances 100–105, Asita 120 112, 169, 205, 289, 302, 329–330, 429 Askot 383, 386 Alexander the Great 50–51 Asoka, Emperor 49, 66n2, 81, 122 Ali, S.M. 75 Astapada, Mount 129–131 Allahabad 11, 431 aśvamedha 40, 67 See also Prayag Atharvaveda 19, 39, 41–43, 47, 67, 306 Almora 68, 92, 101, 103, 137, 140–142, 145, Atharvavedic 19, 42, 46, 52–53, 68–69, 94n9, 149n1, 190, 201, 203n13, 207n28, 383, 95–96, 104, 195, 203n13, 226, 310 388–389, 393, 396–397, 407, 419–420, Ati Muwer (deity) 371 450 Atlantis 411–412, 423 Alpine Club, The 1n1 Atiśa 290, 296, 307 Amdo 345, 361 Atkinson, Edwin 137–139, 141–144, 384, 394, Amnye Machen, Mount 334, 337, 350, 368, 401 370, 410n2 ātman 42 Anagorika Govinda, Lama 269, 419–422, 423, Aurangzeb, Emperor 381 425–426, 447 Avalon, Arthur ananda (suffix) 231n24 See Woodroffe, Sir John Anavatapta (lake) 116, 121, 125, 318–319, Avataṃsaka Sūtra 125, 308 321–322, 437 Avesta 29–30 Anavatapta (mountain) 81n65 Ayodhya 12, 113n89, 152n8, 172n90 See also Anotatta, Manasarovar, and Rakas āyurveda 46 Tal Anavataptagäthä 121 Ba Phnom 27 Andrade, Antonio de, 203, 209, 376 Baba Kailas Angkor kingdom 27 See Adhi Kailas index 513

Bacon, Francis 377 Bhairavakula Tantra 281 Badari Bhairon (deity) 205n22, 220 See Badrinath Bhairongati (confluence) 215, 220 Badra (river) 80 Bharata (continent of) 74, 76 Badrinath (place) 55, 57, 66–67, 85, 92–93, Bharati, Ananda 207 98, 103–104, 111, 135, 137, 140–141, 143, Bhega 167 188, 201–203, 204n19, 205–200, 210n44, Bhimagiri, 187n70 211–213, 215–217, 219, 221–222, 225, 235, Bhimakali (deity) 184–186, 187n70 239, 244–246, 372, 376, 389n53, 392, 395, Bhutan 331–332, 338, 358, 438 430, 435 Bhutanese monasteries in Ngari 332n103 Badrinath (Sri)/Devta Badrinath (deity) 178, Bhutias 93, 194, 199, 200n4, 203n13, 382 184–185, 206, 435 Bihari Giri 172 Baekdu, Mount 25–26, 78, 427 Bindusara (lake) 80, 84, 87, 210 Bajpai, S.C. 183 Biographies (spiritual) Baku 103n52 See rnam thar Balahhadra Giri 172 Birth of the War God 385 Bamian 87, 93n6, 381 Bishop, Peter 402–403 Bamyan Blavatsky, Helena Petrovna 414–416, 420 See Bamian Blezer, Henk 341, 352 Bandits (in Kailas region) 101, 237, 338, 406 Blue Annals 292 “Barkha Tarjum” 391 Bogle, George 215, 377n10 Barth, Frederik 443 Bollee, Willem 130 Baspa Valley 176–177, 184–185 Bön 7, 8, 14, 15, 19, 79, 92, 152, 168, 182, Bathing (sacred/ritual) 122, 176, 187, 189, 317 244n62, 273, 275, 277, 281, 284, 286, Baz Chand, King 92, 98, 202, 203n13, 207n28, 291–293, 295, 299, 302, 304, 314–316, 324, 430 334, 337–363, 365–369, 372, 426–429, Bazhed (sBa bzhed/dBa’ bzhed) 283 442n40, 446, 448 Beas (river) 36n46 Definition of 339–343, 346 Beckwith, Christopher 344, 348 Modernism, absence of 339 Bell, Sir Charles 217, 335, 400 Nativism 340n5, 342, 343n19, 413 Bellezza, John 130, 343, 348, 350, 361, Suggested influences on 342, 349n40 366–368 Brahma (deity) 55, 58, 74, 135, 139–140, 159, Belton, Pauline 326 370, 402, 404, 438 Benares 12, 38n55, 72, 128, 135, 188, 208n36, Brahma Purāṇa 72, 83 245, 362, 388, 395, 416, 431 brahman (‘universal soul’) 35, 55 Benares Sanskrit College 381, 388 Brahmani/Bharmani Devi (deity) 159, 166 Bengal 144, 377 Brahmans (general) 35–36, 38–39, 43, 48, Bernier, Francois 381 53–54, 62–63, 93, 95, 110, 151, 170, 211–212, Besant, Annie 399, 415 214, 226, 228–229, 241, 362, 381, 417, 433, beyul 180 444 Bhadrakali (deity) 155, 171, 186n62 Brahman caste 11, 17, 42, 46, 65, 68–69, Bhadrawah 152, 154–157, 160, 162n46, 167, 133, 198, 220, 240–241, 335 184, 190 Brahman priests 12, 16, 40–43, 48, 134, 151, Bhagirath(i), King 60, 83n72, 209–210 170–171, 186, 188, 195, 215, 217 Bhagirathi (deity) 211 South Indian Brahman priests (in Bhagirathi (river) 79, 141, 200, 204, 212, Himalayas) 9, 211n44, 228, 215–219, 225, 246, 430, 446 231–232, 236, 361 Bhagwan Sri Hamsa 191, 417–418, 422, 424 Brahmanical system/tradition Bhairava (deity) 69, 205n22, 306, 312, 336 See Hinduism 514 index

Brahmanical schools 41 Burton, Richard 378–379, 384, 414 Brahmanical immigration into Himalayas Bushahr 151, 167, 169, 174, 177–180, 182, 133, 202 184–189, 198, 206, 217, 284, 359 Brahmāṇḍa Purāṇa 74, 84 Bushahr-Tibet Trade Treaty 178, 180 Brāhmaṇas 42–43, 47 Bushahru Naga (deity) 185 See also Śatapatha Brāhmaṇa Buston 106 Brahmapura 159 Brahmaputra (river) 1, 59, 77, 81–83, 320, Cakrasaṃvara 376, 400 See Demchok Brahmaur 108n75, 112, 152, 159–160, 164, 169 Cakrasaṃvara Tantra 97, 106, 107n68, 281, Brahmo Samaj 144 296, 305, 307, 312–314, 330, 426 Britain/British 6, 7, 8, 16, 82, 110, 132, 138, 140, Cakravarman, King 286n50 142, 146, 150, 152, 174, 178–179, 181, 190, Calcutta 238 196, 204–205, 212, 216, 218, 224, 232, 235, Cambodia 27, 362 237, 336–337, 377, 415 Cannabis 101, 106–109, 170, 187, 233, 242–243, British policy towards Kailas-Tise 383, 386 336–337, 377–380, 384, 387–396, 405 Capuchin 375 British Trade Agency in Gartok 310n22, caryā 318, 326 336n116, 386–390, 404–408 Cassels, W.S. 406 British Garwhal 190, 204 Caucasus, Mount 381 British Resident 161 Celtic mythology 411, 419 Bru (clan) 354 Central Asia 10, 13, 85, 108n72, 277, 317, Btsan (deities) 365, 367–368, 372 342–343, 370, 438 Buddha (Sakyamuni) 2, 26, 27, 45, 114–115, Central Asian Cultural Complex 348, 119–120, 124, 127–128, 305, 307–308, 349n40 317–318, 325, 336, 404 Chait Singh, Raja 208n36, 431 Buddhacisation 153, 189, 275, 304, 315–317, Chaksu (river) 334, 337, 351, 361, 364, 449 See Paksu Buddhaghosa 114 Chamba 103–104, 111–112, 150, 154, 158–161, Buddhaguhya 282–283, 310 163–166, 168–171, 184, 187–188, 196, 209, Buddhaguptanath 208n36 330–331, 348, 372, 429–430–431 Buddhism, 4, 14, 17, 26, 27, 49, 53, 94, 96, 99, Chamba Gazetteer 162 112, 128, 152, 155, 168–169, 180, 186, 207, Champawat 202 217n74, 273, 275, 278, 285–286, 290–293, Chand dynasty 138, 145, 202–203, 210n44 296, 313–315, 325, 327, 329, 336, 339, 344, Chandapur Kailas 228 346, 352, 355–356, 358, 361, 371, 426, 428, Chandler, David 362 442 Chandra Parbat, Mount 228 Buddhist sources 2, 13, 48, 66, 73, 79, Changchub Gyaltsen 323 114–115, 120–121, 124–125, 127–128, 320, char dam 211 355, 365 Charang 176, 179–180, 188nn73–74 Buddhist pilgrimage 7, 48n16, 49, 302, Charpatnath 164, 166 358 Charpatnath (temple) 164 Formation of Tibetan sects in 296–298 Charpatnath tradition 172 Western Buddhism 420 Chatreśvara 67 Buddhists (general) 85–86, 92, 98, 109n77, Chatterji, B.R. 207 114, 115, 130, 176, 178, 179–181, 183, 188, Chaurasi (temple complex) 159–160, 167 192, 282, 302, 329, 337, 350–352, 360, chelas 364, 413, 446 See spirit mediums Buffetrille, Katia 274, 435 Chenab 36n46, 79 index 515

Chhota Kailas Dane, Louis 404–405 See Adhi Kailas Dan-gun 25 China/ese 13, 20, 25–27, 30–34, 38, 40–41, 61, Darada 78, 81, 215, 276, 314, 334, 336, 353, 358, See Dards 398, 427, 449 Darchen (monastery) 332 Communist China 150, 218, 328, 332, 407, darcho 191, 255 421, 450 Dards 51, 290, 333–334 Influence on Tibetan mountain sacralities Darjeeling 335 279, 309, 434, 436n17 Darma Gyaltsen 307–308n15 Chini 174, 178, 180n39, 189 Darma Sonam, Tokden 299–300 Chitkul 176n13, 177, 180, 184, 186n64 Darma valley 190–192, 334 Christian 1, 11, 16, 41, 138, 140, 152, 174, 388, Daśnāmī 113, 172, 240, 431 389n53, 404, 412, 440 Dattatreya (rsi) 139 Cīnas 51 Dattātreya (deity) 417 Circumambulation 3, 55, 157, 175–176, 179, Datura 106–107 182, 187, 189, 274, 284, 300, 302, 309, 315, David-Neel, Alexandra 416, 420, 424 325, 327–328n88, 330, 332, 337, 359, Davidson, Ronald M. 68, 286, 311, 442 428 Dehradun 204 See also parikrama Dehradun Gazetteer 202 Clark, Mathew 113, 431 Deities, assembly of 56, 121, 187, 370 Clarke, John 100 Delhi 209, 358 Cloud-Messenger, The 88, 131, 403 Demchok (Heruka) 1, 2, 7, 102, 114, 284, 295, Colonial state 298, 300–301, 305–307, 312–314, 316, 329, See British 331, 336, 361, 364, 365, 428–429, 446, Communitas 440–444 449 Congress (party) 399 See also Mandala—Cakrasamvara Cosmic order 89 mandala Cosmology 3, 13, 20, 26, 32–33, 44, 71n26, Deo Tal (lake) 246 76–77, 90, 116, 124–126, 128, 160, 167–168, Desideri, Ippolito 376 172, 219, 224, 278, 319, 325, 345, 372, 427 Devaprayag Creatures, associated with “four rivers” See Prayag 80–81, 123, 126, 320–321, 353–354, 356 Devibhumi 239 Cremation grounds/cemeteries 86, 96, Devtās 178, 184 162–163, 170, 243, 246, 306, 312, 331, 336, Dhami, Major K.S. 200 432 Dhanchho 165 Cuchama and Sacred Mountains 419, 421 Dharamsala (H.P.) 169 Cultural transmission 443–444 dharamsalas, plans for establishment at Curzon, Lord 378n13, 399–400 Kailas 407 Cutler, Nathan 4n8 dharma 77, 89, 241, 382 Dharma la (pass) 105, 190 Daba 204n16 Dharmamaṇḍala sūtra 282, 310 “Jongpon” of 291 dhuni 243 Daily Mail 386 dikṣita 39 ḍākinīs (female spirits/deities) 185n59, 306, Dirghasattra 55 330, 336 Discrimination of the Three Vows 318 Dalai Lamas (general) 16, 178, 186n64, District Commissioner, Almora 388 207n30, 323, 337, 405, 410n2, 412, 423 Dodital 220 5th Dalai Lama 208n36, 327–328, 370 Doi Ang Salung, Mount 26, 427 Dampa Tsang 300 Dolanji 183 516 index

Dolma (deity) 98n31 Gaddi 10n, 154n14, 165–168, 197–199, 433 Dorje Ghuya Sangpo 368 Gaddi deities 167n75 Dowman, Keith 99 “Galactic Polities” 150–151, 172, 311, 427 Drakpa Gyaltsen 298, 307 Gampopa 296, 324 Drepung 181n41, 323, 329 Gandhala, Mount 330 Dri chu/Yangtse (river) 345 Gandhamadana, Mount. 57–58, 79n61, 80, Drigungpa (Kargyu) 297, 299–302, 323 116–119, 123, 125–127, 135, 295n77 Drolma La (pass) 98n31, 302 See also Gurlha Mandhata Druid(ic) 381 Ganden 323 Drukpa (Kargyu) 297, 300, 328, 330–331 Gandhi, Mahatma M.K. 236 Duling 186 Ganga Dullu 103n52, 151n6 See Ganges Dumont, Louis 43 Ganga chu (river) 82, 84 Dunhuang (caves, sources) 276, 278, 281, Gangama (deity) 152n8, 210–211, 434 310n25, 342, 346, 350 Ganges 3, 15, 19, 36, 38, 55, 57, 59, 66–67, Durga (deity) 135, 188n70 77, 79–80, 82–83, 85, 88–89, 93, Dvagspo 296 101, 110, 121, 124–125, 127, 139–140, Dyczkowski, Mark 99, 441 143, 177, 196, 199–200, 204, 209–210, Dzogchen 281–282, 355–356, 370, 413 212, 215–218, 220, 224–225, 233, 238, 244–247, 320, 356, 376, 381, 408, 430, East India Company 16, 137, 178, 212, 215, 217, 435, 446 377 Rivers making up the 218–219 Eliade, Mircea 5 Gangotri 5, 15, 137, 141, 143, 201–202, 203n14, Elliot, T.S. 418 205–206, 209–213, 215–217, 219–224, Ellora 90 228–229, 231–232, 235, 238, 241, 244, 246, Ephedra 105 323, 380, 395, 430, 432 Epics Gangotri glacier 261 See Mahabharata and Rāmāyaṇa Gangotri mahātmya 243 European See also Sri Gangotri Kshetra Mahātmya See West Gangotri math 212–213n52 Evans-Wentz, W.Y. 419–421, 424 Gangs Tise Everest, Mount 418 See Kailas Eygpt, ancient 381, 411, 419 Gansser, Augusto 52n34, 191, 246 gar log (tribes) 290 Fa-hsien 124 Garden of Eden 78 Falk, Nancy 63 Gartok Ferishta, Mhd. Karim 215 See British Trade Agency in Gartok Firearms licences for pilgrims, effect of Garuda 155–156 406–408 “Garuda Lords” 160 Four Rivers (assoc/w Manasarovar) 65, Garutman 42 75–83, 123, 125–127, 205, 308n15, Garwhal 66, 68, 92, 135, 137, 150, 185, 201–203, 319–320, 353, 356, 445–446 207, 217–218, 348, 395 Four Vedas Gaumukh 15, 82, 131n61, 141, 199–200, 207, See Vedas 209–210, 213–215, 217, 229, 241, 244, 435, Fraser, J.B. 187, 216, 220–221 446 Freemasonary 411 Gaumukh Glacier in modern period Frontier protectors 163, 433n9 244–245 Frontiers 20, 150, 152–153, 179, 188, 234, Gaurikund (Manimahesh Kailas lake) 159, 240n55, 433 162, 165–166 index 517

Gaya (Budh) 72, 431 Guge/Guge-Purang 92, 177–180, 184, Ge Hong 28 206–207, 277–278, 284, 290, 299, Gekhod (deity) 293, 350n43&45, 369– 301, 323, 327–328, 331–332, 355, 371 429–430 Geluk(pa) (sect) 7, 181n41, 182, 323, 325, Guide-books (dkar chag) 6, 20, 21, 102, 327–329, 331, 356, 408 125–126, 194, 237, 292–293, 335, 343, Gendun Chöphel 182 355–356, 365, 368. Geneologies See also mahātmya See Royal Geneologies of Ngari Guide to India 182 Genesis (Biblical) 78 Gujjar 10n Geological Survey of India 236 Gum-gum nala (ford) 217–218 Gerard, Alexander 169, 175, 181, 187 Gunji 193–194 Gerard, Lt. Patrick 174 Gupta dynasty 49, 88 Ghata (temple) 156–157 Gurkha 142, 151, 178, 187, 190, 202–204, Ghose, Atal Bihari 144 210–211, 382, 430, 434 Ginsburg, Alan 421 See also Nepal Giri (sect) 69, 108n75, 111, 113, 172, 187n70, 195, Governor of Srinigar 216 199, 215, 222, 231n24, 240, 417n24, 429, Gurla Mandhata, Mount 57n48, 116, 127, 237 431 See also Gandhamadana Glaciers (wider meaning of) 67n7 Guru Nanak 169–170 Glover, Maurice 175 Gurugem (monastery) 359–360 gnas ri 274, 276, 334 Gurung, Kalsang 344 The 8 gnas ri mountains of modern Tibet Guru-shishya relationship 242 334 gzhi bdag Goetz, Hermann 431 See yullha Gokarna 209, 224 Gold 50–51, 101–102, 104, 285, 297 Haider, Mirza 375 Goraknath 103n52, 111, 208n34 Han dynasty 28–29 gotras 46, 68, 71, 202 Hanoman (deity) 170 Götsangpa Gonpo Dorje 102, 300–302, 313, Haoma 29 316–317, 330, 428, 441–442, 448 See also soma Gould, Sir Basil 407 Harā, Mount 29 Government of India (British) 388, 397–399, Harahvaitī 29 404, 419 Haramouk, Mount 215n62, 449n56 Government of the United Provinces 388 Haridwar 205, 211, 214, 216, 242, 389n53, 395, Govinandagiri 240 434 Gragspa, King 300 Harrer, Heinrich 245 Granoff, Phyllis 131 Harsar 159, 162–163, 433 Gray, David 305–306, 312, 314 Harshacharita 162n44 “Great Game” 336 Harsil 210, 217–218 Greece 438 Havell, E.B. 106, 144 gṛhamedhin Brahmans 38 Hazra, Rajendra Chandra 70, 73 Gri gum, King 280n26 Hearsey, Captain Hyder Jung 216, 381–382 Grunendahl, Reinhold 57, 245 Hedin, Sven 4n8, 214, 236–237, 375n3, Gshen (gShen) 346 376, 398–404, 408, 410, 413, 417–418, Gshenchen Luga 341 422–425, 432, 447 Gterma Antiquity of Kailas pilgrimage in works of See Terma 401–402 Gtsang Smyon Heruka 324–325 Powerful supporters of 399–400 518 index

Reliance on Atkinson, Sherring etc, Hindu Kush 51, 348 401–402 Hindustan-Tibet road 178 Writings of 400, 402 History of the Punjab Hill States 156 Heesterman, Jan 38n57, 39–40, 41n68, 222 Hiuen Tsang 66, 177 Henbane Hodgson, Brian Henry 204n20 See Alchemical substances Hodgson, J.A. 211, 214, 216–217, 229 Henotheism 59, 372 Hor 357–358, 361 Herbert, J.D. 174, 195, 216 Hovhannes Joughayetsi 190 Hero(es) 63–64, 89, 136, 234, 292, 322, Huber, Toni 101, 105, 276, 310, 314, 319, 322, 324–325, 327, 427, 434 329, 334 Herodotus 30, 50–51 Hevajra Tantra 97 Illustrated Weekly of India 421 “Hidden Lands” In the Forbidden Land 385 See Beyul Inden, Ronald 65, 71 Hilton, James 425 India, pre-modern/Indic 18, 19, 144, 156, 178, Himachal(a), Mount 139–140, 144–145 182, 196, 286, 310, 323, 358 See also Himalayas British rule in 181, 217–218, 387, 398, 407, Himachal Pradesh 183n1, 370 430 Himalayas 4, 9, 12, 14, 34–37, 40, 44n80, Independent (ie.; post 1947) 17, 19, 102, 45–46, 49–54, 56, 58–59, 61–63, 66–70, 151, 160, 226, 238, 407 75, 82, 85–88, 90, 92–93, 98–99, 101–102, As Tibetan “Holy Land” 289–290, 107–110, 113–115, 117–119, 122, 126–132, 296–297, 329, 359 141–142, 150–152, 155, 159, 178, 184, Indian (general) 140, 143, 149, 238, 383, 393, 195–196, 198, 202, 205, 221–222, 224–225, 409, 415 232, 239, 244, 301, 306, 317–318, 323, 356, Indian traders to Tibet 388, 391, 359, 363, 375–376, 378, 380, 384, 413–415, 432n7 422, 429–431, 433, 443, 450 Indian Army 149n3, 189n74, 190, 193 Himavat, Mount 74, 80, 90, 125, 126, 135, 160, Indian External Affairs Department 451 321 India Office Library 138 Himavatkhanda 136 Indo-Europeans 35, 37–39, 40n65, 51–52, 63, Hindus (general) 61, 87, 92–93, 132, 151, 178, 106, 197, 244n62 214, 236, 238, 364, 386–387, 402 See also Arya/Aryan Hindu kings/kingship 95, 151, 156 Indo-Tibetan border 217, 235, 400 Hindu sources 2, 10, 53–54, 70–73, 75, 77, Indra (deity) 35, 74, 305 79, 86, 121–122, 127, 133–134, 140–141, 206, Indus (river) 1, 30, 36n46, 37–38, 43, 51, 221, 224–225, 226, 320, 393, 422 59, 75, 77, 79–81, 83, 125, 245, 320–321, See also Sanskrit sources 375–376 Hindu Pilgrimage 3, 12, 47–50, 53–54, Indus valley 34, 36n47, 82, 439 64, 86, 92, 136, 166, 173, 188, 196, 201, Intelligence Department, British Military 209, 212–213, 222, 227, 229, 235, 381–384, 400 386–393, 405, 408, 431–432, 445, 451 Iran Hinduism 4, 11, 12, 13, 17, 43, 45, 95, 97, See Persia 133–134, 145, 152, 155, 168–169, 180, Islam 6, 19n37, 41, 107, 132–133, 138, 150, 185–186, 193–195, 217, 227, 232, 302, 317, 157–158, 202, 387, 389n53 339, 361, 372, 389, 392–393, 417, 442, 446 Most sacred place in 2, 8, 25, 139, 382, Jogī 170–172 387, 390, 393, 395, 404, 410 Jadh Ganga 141, 177, 200, 202, 204, 216–219, Hinduisation 158, 189 225, 245–246, 430 See also Sanskritisation Jadhang 218 index 519

Jads 9n14, 10n, 93, 192n84, 200n4, 218, Aesthetic appeal of 110, 422 221n83, 228n11, 246, 433 As a toponym 436 Jains 20, 48, 53, 73, 129, 131 European construction of 375–413, Jalandhara 97–98, 103, 306 421–425, 447 Jambu tree 319 First modernist account of 397 Jambudvipa 76, 115, 320 Jain deity 131 Jammu 154, 167 Kailas “Abode of Śiva” model 158, 168, 173, Jamuna (river) 36n46, 55, 59, 77, 79, 88, 124, 189, 195, 197, 239, 361–363, 430 141, 204, 215, 220, 221 Kailas Temples 89, 90–91 Japan/ese 398, 408 Meru-Kailas association 219 Jātakas 115–117, 120–121, 123 Milarepa at 294–295 jaṭilā 37 Modern pilgrimage to 190, 408, 450–452 Jesuits 177, 203, 214, 375, 377, 409 Number of visitors to 383–384, 406–407 Jha, Makhan 48 Ruined Buddhist monastery at 276n13 Jhelum 36n46 Universal spirituality, construction of Jigten Gönpo 126, 260, 295, 297, 299, 318, 321, 1–2, 418, 422–423, 425–427 324, 442–443, 447 Kailas Manasarovar (book) 237–238, 421 Jin dynasty 25 Kailas mountains’ association with funerals Jitman-ji 155–157, 256 432 Jñānārṇava Tantra 97–99 Kailash ashram/math 227, 240–241 Jorkanden (peak) 174 Kālachakra Tantra 308 Joshi, Dharmanand 141 Kalash religion 348n36 Joshi, Maheshwar 66 Kali (deity) 155 Joshimath 66, 201 Kali (river) 66, 140, 150, 190, 203 Joshimath math 208n36 Kali yuga 130, 133, 225 Jugnao 183n51 Kalidasa 88, 91, 109, 131, 385, 403, 418, 428, Jumla 151n5, 201 447 Jumlee festival 193 Kālikā Purāṇa 96–97 Jvalamukhi 98, 103–104, 112, 140–141, 151n6, Kalinda pass 231 169, 180, 188, 201n7, 302, 329–331, 429, kalis (local deities) 185n59 435, 440 Kalpa Jyolanka 191 See Chini Jyolingkong 190–191 Kamru 177–178, 180, 184–185, 206, 432 jyotirlingas 430 Kangxi emperor 26 Kangyur 307 Kadampa 296, 323 Kāpālikas (sect) 68n12, 69, 111 Kailas (Tise) 1–10, 12–13, 15–16, 18–21, 25, Kaplas Kailas 149, 154–159, 162, 166–168, 173, 27–30, 32, 37–38, 43–46, 49–54, 56–60, 184–185, 197–198, 249, 334 63–93, 97–100, 102–106, 109–121, 123–126, Kaplas Kund 154, 157–158 128–131, 133–146, 149, 165, 180, 182, 184, Kapstein, Mathew 281, 283 186–187, 189–191, 193–194, 197–199, Kargyu(pa) (sect) 7, 15, 181, 295–296, 298, 203, 206, 210, 223–225, 227–231, 233, 299, 303, 305, 307–308, 312–316, 319–320, 235–238, 243–246, 248, 273–276, 322–330, 353, 360–363, 368, 408, 278–279, 281–284, 290–293, 295–296, 428–429, 440 298–301, 304–305, 308, 312–313, 315–316, Sub-sects of 297 319–320, 323–328, 330–337, 339–340, Kargyu Wangchuk, 48th Desi of Bhutan 332 343–345, 349–372, 375–376, 380, Karma (Kargyu) 300 382–410, 411–416, 418–419, 421–422, Karmapa, 16th 327n80 425–437, 440, 445–446, 448–452 Karmapa, 3rd 293 520 index

Karmarupa 97, 101n44, 306 “Koothumi” 415 Karmay, Samten 274, 333, 346, 350 Krśna (deity) 177–178 Karnali (river) 1, 59, 77, 81–83, 140, 219, 356, kṣatriya 35, 42, 51, 53, 58 446 Kubera 56, 58, 60, 73–75, 83–84, 86, 89, 120, Karnaprayag 212 135, 165, 206, 239, 371–372 Karu Druwang 356–357 Kubjika (sect) 100 Kashmir 51, 68, 85, 97, 102, 108n72, 150–151, Kubjikāmata Tantra 97–98 154–156, 159, 171, 177, 179, 285, 331, 381, Kugti 152, 160, 162–163, 218, 330 395, 400, 449n56 Kulacūḍāmaṇi Tantra 97 Kasi raja 139 Kulu 68, 150, 152n8, 153n10, 159, 177, 306, 330 Kasi/Kashi Kumaon 66, 92, 137–138, 140, 142, 145, 150, See Benares 201–202, 215, 348, 383, 395, 430 Kathmandu 136, 178, 313n37 Kumaon Vikas Mangal 451 Katyuri (dynasty) 66–67, 176, 201 Kumbh Mela 11, 442n38 Kaula (sect) 100 Kunāla Jātaka 119 Kaurava (clan) 56 Kunga Gyaltsen 298 Kawa Karpo, Mount 334, 337 Kuninda (dynasty) 66–67, 201 Kawaguchi Ekai 398–399, 408, 416, 418 Kunlun (Mount) 78, 336n116, 405, 435 Kedara(nath) 98, 111, 135, 137, 141, 143, 146, Kunu 284, 315 201n7, 204n19, 205, 208, 210n44, 211, 213, Kuru 36, 43, 210, 309 219, 221–222, 225, 389n53, 395 Kuruksetra 36, 135, 171, 210, 309 Kedarakhanda 136–137, 141, 210, 224, 232 Kushal, Molly 167 Kelangnaga 161–163, 168, 173, 197, 435 Kushan empire 244n62, 342n14 Kelasa Kuti 190–194 See Kailas Kværne, Per 182, 370–371 Kelasakuta (peak) 115 Ketumati (river) 118–119 Labchi 294, 298–300, 313, 315, 323, 334 Kham(pas) 181, 345, 357, 361 Ladakh 7, 20, 51, 152, 177–178, 181, 275n8, 328, Khaskar 175–176, 189 330–332, 406, 430 Khon (family) 330 Lahaul 104, 160, 162–163, 169, 177, 330–331, Khyunglung 345, 359n79 384 Khyung-sprul 182–183, 188, 351, 357n71, Lama Zhang 297–300, 303 358–361, 443 Lambigad glacier 200 Kim Jong-il 26 Landor, Henry Savage 385–386, 397 King, Richard 133 Laos 27 Kinnars 51, 119–120, 176 Laxmi-Narayan 57, 69, 166, 171, 198, 209 Kinnaur 51n30, 151, 169, 176–184, 187, 193n88, Laxmi-Narayan temple, Chamba 161, 166, 171 206, 331, 348, 359, 416 Lha ri (“Soul-mountain”) 277–278, 339–340, Refugees settled at Brahmaur 163 349–350 Religious world of 179–180, 184–187 Lha btsan Spelling of 174n1 See Btsan Language/dialect 177, 188 Lhasa 2, 7n, 93, 172, 179, 181, 208n36, 282, 298, Kinnaur (Kinner) Kailas 149, 169, 174–199, 329, 337–338, 380, 387, 391, 398, 400, 229, 250–251, 334, 351, 385, 416, 428, 430 405, 410, 429–430 Kiras 160 Li Gotama 421 Kiratas 51, 67 lingam 162, 164, 165n61, 166 Klu 7, 169n80, 275, 279, 315–317, 344, 353, 437 Lio Porgyal, Mount 385 Kongpo 346 Lipu Lekh (pass) 140, 190, 194, 396, 450 Kongpo Bönri, Mount 334, 345, 350 Liverpool University 227 index 521

Livingstone, David 378 Malalasekara, G.P. 115, 116n11, 120, 123 Lo Bue, Erberto 310 Malla kingdom 151, 201–202, 315 Local religion 9, 14, 34, 151–153, 157, 161, Malla, K.P. 136 173, 179–180, 184–185, 189, 193, 196, 211, Mana 216, 246–247, 438 219–220, 241, 275, 286, 291, 295, 302, 317, Mana La (pass) 85, 203, 216–217, 230, 346–348, 349n39, 350–351, 361, 372, 413, 245–246, 376, 438 422, 427, 430, 432, 434 Manas(a) (multiple meanings) 58, 438 Annual ritual 156–158, 161, 183–184, Manas(a) (deities) 438 191–192, 210–211, 219–222 Manasa, Mount 58–59 Lokāloka (mountains) 76 Manasakhanda 103, 136–143, 145, 191, 210, London Missionary Society 388, 404n102 224, 232, 394, 401, 403, 447n53 Lopez, Don 420 Manasarovar 1, 3, 6, 7, 8, 9, 50, 56–60, 66–68, Lost Horizon 425 74–77, 80, 82–87, 97–98, 100, 103, 112, 116–120, 122, 127, 135, 140, 144, 151n6, 168, Ma sangs (deity/ies) 344 187, 190, 193, 202, 205, 210, 215–216, 236, Macdonald, A.W. 315, 444 247, 275–276, 278–279, 281–282, 284, Mada (lake) 84, 87 290–292, 300, 302, 304, 315–319, 321, Madden, Captain 175, 187 327–329, 339, 343–345, 349–353, 360, Madros (deity) 7, 131, 275, 299n95, 301, 372, 375–376, 381–383, 386, 391, 401–402, 315–317, 364, 439 404–405, 413, 418, 427, 429–430 Madros (lake) 353, 437 As a toponym 437–438 Magadha 49, 67 Monasteries around Manasarovar Magi 30–32, 39, 41, 226, 448 329n89 Mahābhārata 5, 45–46, 48–49, 51, 52–54, “manasa-sannyasi” 243n58 56–57, 59–64, 66–68, 84, 87, 93, 114, Manchu 135–136, 197, 206, 238, 243–244, 247, 402, Manda/Mandakini (lake) 80, 123, 437 428, 438 Mandakini (river) 74, 80, 83–84, 127, 141, 146, Mahākala 302 205, 207, 219 Mahāmeru maṇḍala 7, 34, 150, 308–315, 329, 365, 372, See Meru 449 Mahānirvāṇa Tantra 143–145 Cakrasamvara maṇḍala 302, 303, 308, Mahāpuraṇas 316–317, 329, 335, 361–362, 449 See Puraṇas Mandala of Sri Kailas 244 Mahāraj Om Naga Giri 240 Mandara, Mount 130 Mahā-Rudra-Manimahesha 164 Mani Bhadri (deity) 206–207 Mahāsiddhas (84) 159, 166, 169, 208, 290, Manimahesh Kailas 149, 158–170, 172–173, 414n12 180, 184–185, 189, 196–198, 222n85, Mahatmas 410, 414, 417, 422, 424–425 249–250, 334, 363, 430 Mahātmyas 63n79, 72, 98, 103, 136, 141, 172, Modern Tibetans to 169 184, 188, 191, 213, 224, 229, 243, 390, 431 Manimahesh (temple) 160, 164–165 Mahāvastu 119–120, 124 Manimahesh Śiva 158 Mahāvira 45, 129–131 Manipat Shah, King 203 Mahayana-Tantric 278 Manjusri (deity) 282 Mahāyoga Tantras 283, 306 Manosilatala (peak) 115–116 Mahendra, Mount 59, 210 mantras 41n68, 52, 193–194, 242 Maheshwar/Mahesvara 163–164, 184, Manusmṛti 70 302n113, 307, 312–314, 316 Mapang (Klu/Naga deity) 192, 356 Mahi (river) 79 Mapang (lake) Mair, Victor 30 See Manasarovar 522 index

Mapang Gulach (deity) 192–194 Mountains 27–29, 31–34, 59, 112, 168, 173, 313, Mapham (lake) 433 See Manasarovar See also names of individual mountains Marco Polo 410 Mountains as frontiers 28, 163, 180, 186, Marlungpa Wangchuk Sengye 299, 324 189, 191, 276, 332–333 Marpa 292, 295–296, 305, 318, 320, 324 Mountains as abode of the dead 175, 180, Martin, Dan 354 189, 432 Martyn, J.A.K. 228 Mountaineering, objections to 220 Mathi Devi (deity) 180 Use in oaths 280n24 Maṭh 100 Mountain gods (general) 27, 33 Matsya Purāṇa 75, 79–80, 83–84, 127, 210 See also names of individual deities Mauryan empire 49 Mucalinda (lake) 117–120, 123, 131 Me Parang (deity) 192n84 mudrās 242 Mecca 1, 387, 389n53 Mughals 113, 132, 145, 149–150, 152, 160, 177, Meghadūta 196, 375, 377, 381, 431 See Cloud-Messenger Mukhtinath 103n52, 151n6, 313n37 Melchizedek teachings 412 Mukhwa 210, 211n45, 217n74, 220 Menri (monastery) 183, 356, 358 munis 17, 53 Meri (deity) 370n38 Muslim Meru 3, 13, 26, 27, 44, 58–60, 65, 74–79, 82, See Islam 84, 85n84, 87, 90, 116, 124–126, 128, 130, Mussoorie 384 135, 160, 219, 307, 312, 345, 349, 354, Muz(h) Tagh Ata, Mount 36n48, 244, 254, 434–435 354 Modern Indian Mount Meru 228, 245 Mysore, Maharaja of 407 Meru (W.B. Yeats’ poem) 418 Merugiri 98 Nāgas 7, 9, 12, 118, 122, 130, 151, 155–158, 160, Meruvarman 159 165, 173, 180, 185–186, 188, 198, 220, 284, Mesopotamia 411 361, 413, 432, 437 Milarepa 7, 109, 119, 292–294, 296, 302, 304, Naga lakes 7, 9, 122, 149, 157, 168, 186, 220, 313–316, 318, 320, 324–325, 327, 352, 361, 247, 275, 317, 427, 432–433, 437 426, 442 Naga (renunciate order) 207, 240 Ministry of Rites 31 Naga stones 221 Missionary (strategies, etc) 9–10, 14, 19n38, Naga tribes 120 152–153, 158, 182n48, 183, 195–196, 198, Nag-Pal, King 154–157 355, 409, 433, 443–444, 448 Nalanda 296 Missionaries, Christian 379, 404 Nalika, Mount 117–119 Mithra (deity) 29 Nam Lugyel, King 298 Mo-tam, Mount 362 Nanda (deity) 207n28 Modernism/ity 6, 16, 133, 139, 144, 196, Nārada Bhakti sūtra 232 224–225, 230, 240, 243, 316, 343, 359, Narayan ashram 194 388, 392 Narayan(a) 55–57, 180, 185 Mongols/ia 215n67, 276, 301, 306, 319, 323, Narayana (sect) 370 Narendranaga 204 Frontier mountains in 333 Narmada (river) 78 Mongol-Tibetan army 328 Naro Bönchung 293, 352, 442 Monserrate, Antonio de 375–376 Naropa 290, 292, 295–296, 305, 318, 324 Moorcroft, William 217, 228–229, 381–382, Nāth (sect) 68n11, 69, 100, 104, 111–113, 154n13, 405 166, 183n51, 187, 196, 201n7, 205n22, Morang 183, 186n64 206–209, 222, 428–430, 434, 440 index 523

Nawang Namgyal, Shabdrung 331–332 Padari (pass) 154, 167 Nazorean Essenes, Order of 413 Padmasambhava 102n49, 166, 169, 179, 182, Nebesky-Wojkowitz, René 366–367, 370 281, 286, 296, 324, 326, 351, 359, 372, 376, Neel Kanth Kailas, Mount 149n1 443 Nehru, Jawaharlal 238 Pagmo Drupa (founder) 297–298, 324 Nepal 20, 82, 93, 97, 101, 136, 151, 172n90, 201, Pagmo Drupa (Kargyu) 297, 301, 324 210n44, 212, 315, 358, 398, 451 Pahari 19 See also Gurkha Paksu (river) 79–80, 83, 125, 320 Neru/Sineru Pal dynasty 204n16 See Meru Pali (sources) 19, 86, 114–115, 118–120, 125, Neru Jātaka 124 127–128, 130, 247, 435, 437 Neumaier, Eva 172 Pallava dynasty 90 New Age 236, 336, 342, 361, 404, 411–412, 414, Pamirs 51, 75, 244, 354, 381 419, 421, 426, 446 Pan-Asian 17, 25, 27, 32, 96 Ngari (western Tibet) 53, 67, 140, 142, 160, Pan-Babylonianism 5n 168n76, 177, 179, 181, 183, 193–194, 197, Pan-Hindu 219, 408 200n4, 201–202, 204, 205, 276–279, Pan-Indic/Indian 112, 158, 171, 173, 188–189, 281–283, 287–288, 290, 301, 305, 314, 196, 199, 201, 207, 209, 211, 214, 224, 317, 322–324, 328, 332, 337, 345, 348, 354, 229–230, 231n24, 239, 241, 362, 431, 433 357–359, 361, 380, 382n27, 386, 397, 401, Pan-Tibetan 327 416, 428, 430, 434 Panchen Lama 208n36, 215, 322, 327–328, 19th century decline of 338 410n2, 430–431 Nīlamata Purāṇa 72 Pandakila (fort) 192 Nilang 141, 177, 202, 217–218, 221, 228n11, 229 Pandava (clan) 56, 62, 154n13, 166, 186n64, “Nine black mountains” 125, 320, 350n41 192, 193n88, 224, 247, 443 Niti La (pass) 203 Pande, Badri Datt 137 Norbu Ling (monastery) 357 Pandit Lal Singh Upadhya 441n36 Nyanchen Tanglha, Mount 334, 350, 365, Pandu Parbat (deity) 192 368, 370 Pannaka (deity) 122 Nyima Bumsal 357–358 Pant, Rudrapatta 137, 140, 142 Nyingma(pa) (sect) 182, 282–283, 288, 296, parikrama 144, 176, 179–180, 183, 186, 326, 330–331, 341, 355, 429 189–190, 401, 405 Nyipangtse (deity) 370 Parmar dynasty 203, 207 Parvati (deity) 139, 143, 193, 258 Oddiyana Parvati (lake) 190, 193 See Uddiyana Pāśupatas 68–69, 93, 111–113, 135, 222, 362, Ögedai Khan 301 428 Old Tibetan Chronicles 345 Patanjali 418 Ölmolungring 244n62, 351–355, 357, 361, 428 Peking 215 Omei Shan, Mount 331, 368n23 Peri (deities) 185n59 ‘Opening’ (a pilgrimage) 301–302, 330, 423, Persia 25, 29–35, 38, 40–41, 100n39, 353, 427 428 Phnom Kulen 27 Orgyenpa Rinchen Pal 181–182, 330, 350n43 Phou Si 27 Orissa 72, 117n14 phyi dar (“Later/Second Dissemination of Ortner, Sherry 340 Buddhism”) 278, 281–285, 289, 305, 315, Osmaston, Major Gordon 200, 228 317, 336, 341, 343–344, 349, 351, 354–355, Oxus (river) 75, 79–80, 244n62 365, 370, 372 Pilgrimage, by Royalty 171, 196, 202, 315 Padam Singh, King 359 Pilo/Spilo/Spello 183–184 524 index

Pinch, William 15, 431 262 pipīlika 51 Raldang (peak) 174–175, 186–188, 385 Pir Naga (deity) 185 Ralung 297 pīṭhas 96–99, 103n52, 165, 187, 301, 306, Rama (deity) 58, 61n72, 62, 64, 83, 117n13, 313n37, 317, 330–331, 430 136, 222 Plutarch 30 Ramakrishna mission 194 Poetry, Sanskrit court 131 Rāmāyaṇa 5, 45–46, 52–54, 58–61, 63–64, Potala (in Lhasa) 402 66, 68, 83, 86–87, 93, 114, 117, 139, Potala, Mount 282 197, 238, 243–244, 247, 394, 402, Powari 176 428 Pradyuma Shah, King 203–204 Ramble, Charles, 356–357, 366 Pratap Singh, King 163 Ramesvara(giri?) 172 Praum Puri 215–216n68 Rampur 178, 185, 187, 212, 358 Prayag 59, 88–89, 98, 171, 211, 218–219 Raper, Captain Felix 216 Prince of Wales Museum Bombay 421 Rasas 100, 166 Princely states 132, 140–141, 204, 212, 230 Ravana (deity) 58, 86, 165, 371 Ptolemy 51, 375 Ravi (river) 36n46, 79 Puh 385 Rawain district 204 Pulliramalaya 97 Rawling, Captain Cecil 386, 387 Punjab 34, 36–37, 78, 151, 166 Rechungpa 293, 325 Punjab Hill States 150, 160, 174, 178 Rekong Peo 174, 189n74 Purāṇas 49, 65, 69, 70–74, 78, 84, 87, 93, 96, Rennell, Major James 215, 381 98, 119, 134, 136, 138–139, 141, 165, 222, Renunciate(s) 2, 4, 6, 7, 14–18, 25, 26, 33, 226–227, 232–233, 238, 247, 320, 339, 38–40, 43, 45–47, 52–53, 62–63, 69, 77, 381, 402, 428, 438 87, 89, 93, 95–96, 98–101, 103–113, 117, Purang 202, 298 119–120, 128, 134, 145, 164–165, 168–172, See also Taklakot 174, 187–189, 194–195, 198–199, 201, Purang district 277, 300 203n13, 206–207, 208n36, 212–213, Purangir(i) 215 221–223, 225, 227, 232–234, 239–243, Purang-Yartse kingdom 323 281–282, 286, 288, 291–292, 304, 306, Purity/impurity 31, 38–42, 46, 53, 93–95, 110, 315, 322, 326, 329–331, 362, 383, 386, 392, 170, 186n62, 195, 222, 288 395, 405, 408, 417, 427–431, 440–441, Purnagiri 97 444, 448 purohita 42, 95 As archetypes 111, 240–241, 448 Pushkar 54–55, 62 British dealings with 212, 215, 383n32 Displays of magical power 300 Qing dynasty 25 Links to Independence movement 226, “Queen Mother of the West” 78 230, 236 Patronage of 112, 132, 170–171, 182, 188, Rabindranath Tagore 421 203n13, 221, 226, 230n21, 231, 241, 282, Raj Rajeshwari (deity) 208 288–304, 315, 322–323, 329, 351 rajaguru 166, 330 Self-mortification 241n57 Rajaju (deity) 192 Trading networks 100–101, 104–105, 112 Rajastan 54, 97n26, 207, 214 Reting 296 Rājatarangiṇī 160n37, 177 Rewalsar 169, 181–182, 358 Rakaposhi, Mount 148n1 Ṛgveda 13n22, 35–39, 42, 44n80, 47n12, 51, Rakas Tal 1, 82, 84–86, 103, 116, 118–119, 140, 305, 308, 435 318, 321, 356, 376, 382 See also Vedas Raktavarn(a) Bamak (glacier) 200, 245, Ribba 176n13, 179 index 525

Ridges (wider meaning of) 67n7 361–363 rimed 183, 337, 355–356, 359 Sakya(pa) (sect) 296, 298, 300–301, 307, 319, Rinchen Zangpo 177, 179–181, 189, 285–286, 323, 325, 328, 330 291, 307, 351, 407, 428, 443 Sakya (monastery) 296 Ripa Nagpo 303 Sakya Pandita 3, 298, 323, 368, 446, 449 Rishikesh 204, 218, 227, 242 Critique of Kargyu Kailas-Tise equation rnam thar (religious biographies) 109n78, 317–322, 325, 368 324–325 samadhi 166, 170 Roberts, Peter 325–326 Samantabhadra 368 Rock, Joseph 410n2 Samaveda 41 Rodseth, Lars Thomas 443 Sambara/Samvara Roerich, Nicholai 416–417, 420, 424 See Demchok Rosicrucian Order of the Golden Dawn 417 Samdup, Kazi Dawa 144, 416, 419, 420n33, Rowe, Bradley 452 426 Roy, Ram Mohan 144 Samuel, Geoffrey 356 Royal Geneologies of Ngari (mNga’ ris rgyal Saṃvarodaya Tantra 97 rabs) 284–285, 323, 355 Samye 298 Royal Geographical Society 236, 378–379, Sanderson, Alexis 94, 107n70 399 Sangye Tshalpa 299 ṛṣis/rishis 35–36, 38–39, 44–46, 48, 53, 87, Sankaracharya 12, 113, 138, 206, 443 140, 143, 222, 224–227, 230–233, 235, 243, Sanskrit 19, 70–71, 133, 137, 140, 145, 334, 414n12, 437–438, 448 403 See also renunciates Sanskrit sources 11, 12, 88, 114–116, 118, 133, Rudok 330, 370, 396n70 135, 144–145, 206, 224–225, 234, 240, 243, Rudra (deity) 36–37, 53, 165–166, 192n84 247, 290, 307, 320, 404, 432 Rudraprayag 141, 203, 205, 211, 218–219 Sanskritisation 19, 92–93, 113, 134, 145, Russia 7, 26, 336, 382, 388, 398 151–153, 155–160, 163–165, 168, 170–172, 180, 192–196, 202, 203n13, 207, 209, 211, Sacred geography/landscape/space 2, 5, 7, 8, 219–223, 225, 257, 430, 433 14, 15, 27, 28, 37, 47, 49, 50, 54, 71–72, 81, Sapni (palace of the Guge Queen) 180 96–98, 108, 109n78, 112, 116, 118, 121–128, Sarahan 178, 184–185, 188n70 142, 145, 168, 173, 176, 187, 194, 197, 205, Sarasvati (order) 230n23, 240 207, 208n36, 211, 213, 219, 225, 231, 304, Sarasvati (river) 36n46, 37, 54, 59, 210, 216, 314, 320–322, 336, 350, 361–362, 408, 218, 245–246 427, 431, 446 Sarat Chandra Das 415 Various constructions of 325–328, 331, Sarasvati (deity) 155 224–225, 350 Sarayu (river) 37, 58 sacrifice, 29–32, 158, 166, 191–192, 346 Sarju (river) 79 Vedic 40–41, 43, 47, 54–55, 106, 309 Sarnath 81 Sadhu Sundar Singh 174n3, 412 sarovar [Skrt: = lake] 58 Sagara, King 60 Sarvatathāgatatattvasaṃgraha Tantra 307 Sahilavarman, King 160, 166 Śatapatha Brāhmana 371 Śaivite 15, 27, 68–69, 75, 82, 98, 107–108, 111, Sati (deity) 96 113, 158–159, 163–164, 170, 173, 176, 180, Satipaṭṭhāna Sutta 125 186, 188–189, 194–195, 198–199, 205, Satopanth bamak (glacier) 216 207–208, 219, 222, 239, 243, 329–330, Satopanth Tal 216, 246–247, 263, 430, 432 392, 408–409, 429 Schary, Edwin 413–414 Śaivite Tantras 306 Schlagintweit brothers, 384 Saivite response to Kargyu construction Schopen, Gregory 72 526 index

Se (etymology of) 365 165–168, 172, 175, 187, 189–190, 193–198, Sects (definition of) 68 209, 213, 258, 281, 362, 364, 369, 372, Sengchen Lama 415 381–382, 385, 394, 402, 412, 415, 428, 438, Schwarzgruber, Prof. Rudolph 200 446, 449 Schwieger, Peter 330 “Six Yogas of Naropa” 296, 330 Sengge Namgyal 331–332 Śiva Purāṇa 75 Sera (monastery) 323 Śivabhumi 165, 167n76, 173, 198, 222, 239, 449 Shah, Emperor Bahadur 152 Sivling (Gangotri peak) 229, 244 Shambala 308, 410n2, 415–416, 417n22, 425 Sivling (Kinnaur peak) 175, 251 Shang dynasty 27, 31n29 Skanda Purāṇas 58n56, 72, 75, 134–138, 209 Shangri-la 180, 425 Skinner, Thomas 212, 215 Shardza Tashi Gyaltsen 356–357 Slave trade 203n15 Sharma, Mahesh 81, 164, 166, 169, 196 Snellgrove, David 342, 349n40 Shenrab (gShen rabs) 340, 342, 344, 352, 354, Snelling, John 397, 452 368, 412 snga dar (“Earlier/First dissemination of Sherring, Charles 139, 237, 404, 409–410, 413, Buddhism”) 278, 342 422–425, 432, 447 “Snow(y) mountain(s)” 36, 78, 126, 195, 312, Construction of a modern Kailas 387– 318n50, 320–321, 382n25, 436 397, 401, 403, 408, 447 Snyder, Gary 421 Intention to stimulate pilgrimage soma 35–36, 47, 105–106 390–391, 393, 396–397 See also Haoma Reliance on Atkinson 394 Somagiri, Mount 58–59 Sectarian understanding of Hinduism Song, Mount 28 392 “Soul-mountain” Sherring, Rev. Mathew 388 See Lha ri Shigatse 215, 410n2, 416 Sounige (deity) 184 Shipki La (pass) 385 Southern Tibet 402 Siddha Satyanath 207 Spa (clan) 357–358 siddhas 55, 74, 84, 95, 100, 104, 111, 122 Spa Nyima Bumsal siddhi 222, 306, 442 See Nyima Bumsal Sikhs 20, 177, 179 Speke, John 384 Sikkim 416, 420 Spirit mediums 157, 162, 333 Simla Convention 1914 217 Spiritualism 411 Simla 174 Spiti-Lahaul 177 Sindhu Sri Chand 169–170 See Indus Sri Gangotri Kshetra Mahatmya 213, 228, Singh, Jogishwar 177, 185 243 Sinja Sri Kailas 5, 149, 199–201, 216, 219, 223, See Jumla 228–229, 231, 235, 238, 244, 247, 252–253, Sino-Indian Accord 1954 407 334 Sino-Indian border closure 181, 195, 407, Renunciate visits to 233 450 Sri Kailas glacier 200, 245 Sino-Indian war 1962 450 Sri 108 Krishnangiri-ji Maharaj 108n75 Sita (river) 80, 83, 125 Sri Manimahesha 164 Sita (deity) 58, 61n72, 79, 117n13 Sri Purohit Swami 417–418 Sita-Ram (sect) 243 Sri Swami Dhanraj Giri Maharaj 227, 240 Śiva 1, 2, 4, 19, 37, 45, 53, 60, 68–69, 73, 75, 82, Srid pa’i sman (deity) 366 86, 88, 96, 99–100, 108–109, 112, 134–136, Srinagar (Garwhal) 79n60, 141, 203, 207, 430 139–140, 143–145, 149, 157–158, 161–162, Srinagar (Kashmir) 79n60 index 527

Srinivas, M.N. 19, 194–195 Syrian ru 105 Staal, Frits 37–38, 85–86, 116, 120 Tai, Mount 452n60 “Standing stones” 193, 432 Tajikstan 244n62, 353 Standzin, Prince 332 Taklakot 92, 202 Stein, R.A. 276 See also Purang Strachey, Henry 137, 190–191, 382–384, 386, Taklakot Dzongpons 192n86 405 Taksang Repa Ngawang Gyatso 330–331, 360 Strachey, Richard 137, 384 Tambiah, Stanley 150, 311, 427 Strachey, Sir John 137, 140–142, 191 tangka 259, 311 Subramanian Swami 450 Tangyur 283 Sudarsana, Mount 58–59 Tantra (system) 11, 19, 93–96, 104, 145, 442 Sudassana/kuta 123–124 Systemisation of 311 Sumana 122 Reaction against 287n53, 288n56, 289n59 Sumeru Tantras (texts) 94, 142–144, 209, 221, 276, See Meru 286, 296, 307, 441 Sumeru pīṭha math 208n36 Tantric(s) 19, 69, 73, 86, 98–100, 102–103, Survey of India 216, 224, 448 105–107, 109n76, 114, 167, 242–243, Sutlej 1, 36n46, 51, 59, 66, 77, 81–83, 85, 119, 281–282, 287, 302, 310, 319, 322, 336, 351, 177–179, 187, 285, 321, 331, 385 354, 362 Suvannagiritala, Mount 117 tapas 35, 46n10, 62–63, 77, 93 Svarnabhumi 76, 435 Tapovan (above Gangotri) 210, 228 Swami Dayananda (Giri) Sarasvati 226–227, Tapovan (above Joshimath) 66 230, 232, 242, 414 Tara (deity) 302 Swami Harivallabh 381n23 Taranatha 208n36 Swami Jnananandagiri 227, 229, 236–238, Targo, Mount 334, 350 240, 265 Tarim (river) 75, 80n61, 244n62 Swami Pranavananda 83–84, 85n83, 102, 105, Tashilumpo 327 137, 191, 193–194, 227, 236–238, 266, 359, Tavatimsa heaven 120, 124 407n114, 409, 421–422 Taylor, Kathleen 144 Contacts and networks 236 Tazig 244n62, 340, 352–353, 356 Influence of Swami Jnanananda 236 Tecate/Cuchama, Mount 419 Modernist 236–237 Tedong valley 176, 180 Visits to Kailas-Manasarovar 236 Tehri Garwhal 142, 151n6, 204, 217, 225, 235 Swami Purnananda Sarasvati, 226, 230 Tehri, Raja of 227 Swami Sri 1008 Narayan 194 Temisgang, Treaty of 332 Swami Sunderanand 108n75, 212, 231–232, Terma (gter ma) 326n27, 330, 341 242, 244, 268 Terry, Edward 214 Contacts and networks 231 Thag La (pass) 203, 245 Swami Tapovan[am] 3n, 108n75, 212, Thailand 26 219–220, 227–235, 237–238, 241–244, Thakur Jai Chand 388 267, 329, 359–360, 447 Thakur, Laxman 180 Contacts and networks 229, 231 Thakur, R. 175 Influence of Swami Jnanananda 229–230 thakurs 178 Modernist 230–231 Thapa, General Amar Singh 210–211, 221 Spiritual inclinations 230, 234 The Abode of Snow 384 Visionary reading of landscape 233 The Great Plateau 386 Swami Virjananda Sarasvati 226, 230, 232 The Holy Mountain 417–418, 425 Swat 34, 40n65, 51, 180 The Sacred Mountain 397 Syr Daria (river) 75 The Tibetan Book of the Dead 419 528 index

The Way of the White Clouds 421–422 Trisong Detsen, King 154, 283, 355 Tibet’s Great Yogi Milarepa 419 Tsang 290, 323, 354, 361 Theosophy 144, 227, 399, 410, 411, 414–417, Tsangpo 419–420, 424 See Brahamaputra Thok Jalung 50 Tsanpa Tholing 181, 188, 206–207, 219, 244, 290, See Brahma 300–301, 327, 329, 359 Tsanpa Gyare Yeshe Dorje 297 1076 chos kor at 290 Tsaparang 203, 217–219, 376 Abbot of 284 Tsari (mountain) 101, 105, 298–299, 313, 316, Thomas, F.W. 138 323, 331, 334, 337, 358 Three Years in Tibet 398 Tsewang Rigzin 370n37 Tibet/ans 1, 2, 7, 8, 13, 18, 50, 93, 101, 109–110, Tshalpa (Kargyu) 297, 299–300 112, 125, 146, 149–151, 167, 169, 178–183, Tsong Khapa 323 185, 188–190, 192–193, 195–196, 202, 204, Tucci, Giuseppi 14n25, 273, 277, 335, 366, 206–207, 209, 212, 215–216, 218, 233–235, 370, 407, 424n48&49 245, 273, 276, 278–281, 285, 289–290, Turkic/Turko-Mongol 276, 434 296, 298, 306–307, 314, 317, 322, 323, Turner, Victor 440 325, 328, 333, 340–341, 352, 365, 368, 375–376, 379, 383, 391, 395, 397, 399, Ü-Tsang 402–407, 413, 415, 416, 451 See Tsang Tibeto-Burman language 179, 192n86 Uddiyana 97, 180–181, 306, 330–331, 347, Tibet-Ladakh-Mughal war 178 435 Tibetan Book of the Dead, The 144 Uma (deity) 165 Tibetan Guidebooks to Tise-Kailas (list of) Upaniṣads 42–43, 133, 229, 232, 418 335 Uprethi, Ganga Datta 141 Tibetan invasions of Bhutan 331 Urgyen Gyatso 415, 416n21 Tibetan sources 3, 6, 8, 10, 19, 92, 146, 154, Urgyen Wanchuk, King 358 204, 275–276, 279, 325, 328, 345, 353, 422 Uruvela Kassapa 121 Tibetan studies 15, 279, 334–336, 346 Uttarādhyayana-cūrṇi 131 Tidang valley Uttarakashi 220 See Tedong valley Uttarakuru 76, 121 Tieffenthaler, Joseph 214 Tilopa 296, 324 Vaidyuta, Mount 84 Timur 214 vaimukh 242 Tīrthaṅkaras 129–130 Vaiśnavite 18n35, 55–56, 57, 69, 82, 139, 156, Tirthapuri (Pretapuri) 102–103, 112, 302, 330, 171n85, 185, 198, 219, 392 429 Vajramahābhairava Tantra 106 tīrtha-yatra 47 Vajravarahi (deity) 312 tīrthikas 92 Vajrayogini (Dorje Pagmo) (deity) 102 Tise (Ti se) Vaksu (river) See Kailas See Paksu Tise, etymology of 365 Vāmakeśvara Tantra 97 Tise dkar chag 302 Vamka, Mount 117–119 Tokden (rtog ldan) 15 vaṃśavalīs (Royal Chronicles) 74, 80, 84, 133, Toponyms 90, 103n52, 118, 127–128, 131, 159, 155, 166, 171, 177 191, 206–207, 318, 429, 434–439 Varaha, Mount 59 Traill, George 204–205 Varanasi Transhimalaya 401 See Benares Triloknath 111–112, 166n72, 181, 330 Varman dynasty 160–161 index 529 varṇāśramadharma 46n9 380, 385–386 vāstuśāstra 159 Western visitors to Tibet, 1984 onwards Vasubhandu 76, 114, 125 451 Vasudhara falls 216, 246 Western Himalayan Cultural Complex 9, Vasuki Purāṇa 156n21 151n6, 153–154, 158, 161, 165, 168–169, Vasukinag 155–158, 162, 167–168, 256 184–185, 188–189, 191, 195, 220, 222, 275, Vayu Purāṇa 72–73, 80, 83–84, 106 279, 317, 348–349, 427, 432–434, 444, Vedanta 11, 19n38, 144, 196, 229–230, 392 446 Vedas 5, 12, 34, 37, 39, 41, 46, 53–54, 56, 65, Western Tibet 78, 86, 105–106, 133, 197, 210, 225–227, See Ngari 232, 243 Western Tibet and the British Borderlands Veer, Peter van der 431 264, 390, 394, 397–398 Vessantara Jātaka 117–118, 131 White marble statues 162 Vessantara, King 117–119, 122 White, David Gordon 69, 100, 111–112, 311 Vigne, G.T. 154 Wilderness, and Kings 62–63 Vikramasila (monastery) 290 Wilford, Captain/Major 138n23, 381 Vīṇāśikha Tantra 362 Williamson, Frederick 332 Vinaya Piṭaka 120, 290 Wilson, Andrew 384–385 Vindhaya (mountains) 88, 120, 139 Witzel, Michael 348 Vipula, Mount 117, 119 Woodruff, Sir John 144, 416, 419 Visions/Visualisation 99, 104, 108–109, 198, World-centre 3, 44, 78, 129, 308–309, 428 224, 232–233, 238–239, 275, 301, 304, World-mountain 8, 9, 26, 32, 77, 81, 126, 273, 314–315, 325, 341, 359 278, 307, 317, 427 Visionary authority 238–239, 243, 329, World-religions 9n15, 14, 18n36, 27, 32–33, 360 81, 151, 153, 155, 168–169, 182–183, 189, Visnu 45, 55, 82, 135, 139–140, 156, 232, 392, 220, 222, 275, 334, 340, 348, 360n85, 446 365, 427–428, 432, 434–435, 442, 443, Viṣṇu Purāṇa 74, 79–80 446n52, 449, 452 Vitali, Roberto 277n15, 298, 301, 316, 323, 348, Wright, Rita 439 355, 367 wu 31–32, 41, 448 Vourukaša 29 Wu-t’ai shan, Mount 331 vrātya 39–40, 47, 52–53, 56, 63, 67, 94n10, 113, Wylie system 19 195, 444 Yajurveda 41, 47 Waddell, L.A. 335 yakṣas 60, 74, 83, 88, 115, 118, 120, 371 “Walk of Yu” 28n12 Yama (deity) 180 Walter, Michael 278, 348–349 Yamuna Wangdi, Governor of 332 See Jamuna Webb, Lt. William, 216 Yamunotri 211 Welchen (deity) 371 Yarkhand (river) West(ern)/Westerners 3, 6, 12, 33, 82, See Tarim 104, 120, 133–134, 145, 149, 175, 180, Yarlung dynasty 93n8, 277–278, 285–287, 225–227, 232, 238–239, 284, 293–294, 316, 333–334, 339, 343–345, 350, 319, 334–337, 340, 444 355 Western landscape aesthetics 110, 386, Assassination of last King of 288 425 Yeats, William Butler 417–419, 424–425 Western understandings of Tibet 379– Yeshe Gyaltsen 355, 368 380, 385, 397, 402, 408, 410, 422 Yeshe Ö, King 282, 284–288, 303, 351, 355 Western visitors to Tibet, 19th century Ordinance of 287 530 index yoginīs 71n28, 90, 96, 183n51, 306, 312 Zhang-zhung 8, 156, 183, 244n62, 277, Younghusband, Sir Francis 378n14, 380, 333–334, 339, 341–346, 348–352, 399–400, 402, 403n97 355–356, 359–361, 368–369, 413, 427, “Younghusband Mission” 386–387, 404, 429, 434 408, 410 Cultural Complex of 348 Yuan dynasty 323 Religious world of 346–347 yue 27 Zhang-zhung snyan rgyud 281, 355 yullha 9n15, 182–183, 189, 274–275, 284, 315, Zhou dynasty 27 333, 359, 364–365, 368, 370, 429 Zhouli 31 Yunas Singh 404 ‘Zone of the Consecrated Warrior’ 222, 452 Yungdrung gutse, Mount 353 Zoroastrians 29, 33 Zorowar Singh Kahluria, General 382–383, Zanskar 172, 384 388 Zarathustra 30