<<

Research Designs for Hawaiian

Research Designs for Hawaiian Archaeology

Agriculture, Architecture, Methodology

Thomas S. Dye, editor

Special Publication 3

Society for Hawaiian Archaeology All rights reserved.

Copyright © 2010 by for Hawaiian Archaeology.

Published in 2010 in the United States of America by Society for Hawaiian Archaeology, P.O. Box 23292, , HI 96823. Contents

List of Figures vii

List of Tables ix

1 Watershed: Testing the Limited Land Hypothesis Robert J. Hommon 1

2 Traditional Hawaiian Surface Architecture: Absolute and Rel- ative Dating Thomas S. Dye 93

3 Lady Mondegreen’s Hopes and Dreams: Three Brief Essays on Inference in Hawaiian Archaeology Dave Tuggle 157

Index 185

v

List of Figures

2.1 Map of the Hawaiian Islands ...... 96 2.2 Oblique schematic of Kaneaki Heiau ...... 98 2.3 Plan of site 50–10–04–22268 ...... 104 2.4 Interior of the U-shape enclosure at site 50–10–04–22268 . 105 2.5 Panoramic view of site 50–10–04–22268 ...... 106 2.6 Hypothetical stratigraphic section ...... 110 2.7 Bayesian calibration yields interpretable results ...... 113 2.8 Deduction and induction ...... 115 2.9 Plan of site 50–10–04–22119 ...... 121 2.10 Dated enclosure at site 50–10–04–22119 ...... 122 2.11 Plan of site 50–10–04–22201 ...... 124 2.12 Plan of site 50–10–04–22248 ...... 125 2.13 Plan of site 50–50–17–1089 ...... 127 2.14 Plan of site 50–50–17–1088 ...... 128 2.15 Estimated ages of construction events ...... 132 2.16 Plan of H¯apaiali‘i Heiau ...... 134 2.17 Posterior probability distributions for H¯apaiali‘i Heiau . . . 136 2.18 Generation length model ...... 139

vii

List of Tables

1.1 Expansion scenarios in the Kohala Field System ...... 12 1.2 Hawaiian Islands depopulation rates ...... 54 1.3 Tahiti: Estimated depopulation rates, 1769–1863 ...... 55 1.4 Marquesas: Estimated depopulation rates, 1840–1926 ... 55 1.5 Hawaiian Islands: Ancient population estimates ...... 56 1.6 Ratio of Phase 1 and 2 14C dates ...... 59 1.7 Cognate names of chiefs, Hawai‘i and New Zealand . . . . . 65

2.1 Architectural component descriptors ...... 102 2.2 14C dates from beneath surface architecture ...... 122

ix

1

Watershed

Testing the Limited Land Hypothesis

Robert J. Hommon National Park Service, retired

He ali‘i ka ‘aina;¯ he kauwa¯ ke kanaka. The land is a chief; man is its servant. (Land has no need for man, but man needs the land and works it for a livelihood.)

Pukui (1983, 62) Mohala¯ i ka wai ka maka o ka pua. Unfolded by the water are the faces of the flowers. (Flowers thrive where there is water, as thriving people are found where living conditions are good.)

Pukui (1983, 237) About thirty years ago there was much talk that geologists ought only to observe and not theorise; and I remember some one saying that at this rate a man might as well go into a gravel-pit and count the pebbles and describe the colours. How odd it is that anyone should not see that all observation must be for or against some view if it is to be of any service!

Charles Darwin, responding in a September 18, 1861 letter to a “great naturalist’s” view that “[t]he mistake is, that Darwin has dealt with origin. Why did he not put his facts before us, and let them rest?” Darwin and Seward (1903)

1 2 Robert J. Hommon

Preface

• In the centuries following the establishment and growth of the first successful Hawaiian colony by voyagers from central East Polynesia, small groups of people, probably led by junior chiefs, established colonies in far-flung locations throughout the main islands that were best suited for reliable, sustained economic production.

• These initial colonies, numbering about 30, formed the salubri- ous cores of chiefdoms, later to become constituent districts of large polities.

• The colonies grew, expanding settlement laterally along coast- lines and transforming inland ecosystems into agricultural com- plexes along valleys and un-dissected slopes.

• In leeward regions, notably on the island of Hawai‘i, extensive inland agricultural development may have been spurred by the arrival, shortly before the fifteenth century, of a new staple crop, the sweet potato (Ipomoea batatas).

• In the final stage of growth, population doubled to the maximum ancient size within a short time, approximately the fifteenth century. During this period of economic expansion, the aristoc- racy benefited from rapidly increasing revenues in the form of food, manufactures, and luxury goods for themselves and their retinues.

• By the sixteenth century, rain-fed was being prac- ticed increasingly in marginal zones where the frequency and severity of drought-caused crop failure varied inversely with rainfall.

• The Hawaiian population stabilized and may have begun to decline during the late sixteenth century, possibly because agri- cultural expansion was approaching the limits of arable land.

• As the limits of agricultural land were approached, food short- ages and possibly population overshoot led occasionally to lo- calized famine.

• Chiefs in power during the relatively rapid transition from gen- eral expansion to economic stress experienced diminishing con- tributions in taxes and corvée labor. Watershed: Testing the Limited Land Hypothesis 3

• By the beginning of the seventeenth century, in response to eco- nomic shortages and uncertainty, governing chiefs of large and powerful multi-district polities were applying a newly-developed political , conquest warfare, to augment their revenues by capturing other polities complete with their resident producers.

• In the context of competing, belligerent polities, leaders increas- ingly came to be selected for their political and military skills regardless of hereditary status ascribed by genealogical rank and sacred character.

• By the eighteenth century, internal economic, social, and po- litical dynamics of the competing Hawaiian polities had trans- formed them into primary incipient states in every significant re- spect, including large populations, endogamous socioeconomic classes, and centralized governments. These governments, sanc- tioned by state religions, headed by kings, and organized into three or more strata of offices occupied by members of the hereditary nobility, exercised their authority to collect taxes, raise armies, wage true conquest warfare, and construct public works.

These were major elements of the model of ancient Hawaiian history detailed in my 1976 dissertation and amplified in a 1986 paper (Hommon 1976, 1986). The model, describing the rise of what are now termed archaic states, was based on the scant then-available body of data that was often, in retrospect, of questionable accuracy, reliability and applicability. Nevertheless, three decades of subsequent research often have served to test and frequently to support the model. The value of such model-building in generating testable hypotheses seems clear. This paper focuses on a hypothesis arising from principal ele- ments of the model. The Limited Land Hypothesis can be considered a revised and testable version of the filled land notion that has been discussed by researchers in Hawai‘i for more than thirty years. This hypothesis concerns what appears to have been a watershed process in ancient Hawaiian history. I propose that the broad-spectrum trans- formation of Hawaiian in the last centuries before Western contact in 1778 was to a significant degree associated with a marked reduction in the rate of long-term agricultural expansion as natural and cultural limits on arable land were reached. 4 Robert J. Hommon

The primary purpose of this paper is to urge the members of ’s archaeological community to incorporate rigorous tests of the Limited Land Hypothesis into future research. The body of the paper is divided into three parts:

Part 1 introduces the Limited Land Hypothesis and defines its terms.

Part 2 summarizes an array of features of Hawaii’s indigenous history that have supported the construction of the hypothesis and that may be further illuminated as future research lends support to the hypothesis.

Part 3 outlines research procedures that can be applied in the field and the lab to test the hypothesis.

Part 1. Introducing the Hypothesis

The Limited Land Hypothesis states:

By ad 1550 ancient Hawaiian agricultural expansion slowed significantly as it approached effective limits on staple pro- duction imposed by available , sociopolitical factors, and natural variables including rainfall and soil fertility.

This hypothesis pertains to primary agricultural expansion, that is, the process of transforming previously uncultivated land into cropland. Expansion refers to areal augmentation as distinguished analytically from intensification, which refers to the process of in- creasing production per unit of land already under cultivation. Under some circumstances, these processes can be observed in isolation, as for example expansion in the form of initial clearing and planting in a shifting cultivation regime, involving no long term investment or increased intensification such as the addition of terraces in an already established agricultural complex. More commonly the two processes probably took place simultaneously in ancient Hawai‘i. For example, establishing new lo‘i (pond-fields) along a permanent stream involved both areal expansion and construction of terraces, canals and other long-term improvements that together constituted landesque capital intensification (Blaikie and Brookfield 1987; Kirch 1994). Watershed: Testing the Limited Land Hypothesis 5

The hypothesis can be tested in any irrigated, rain-fed, or flood- water agricultural complex where the expansion process can be an- alyzed archaeologically. In principle it can be tested at any scale, ranging from a single cluster of lo‘i to the archipelago as a whole. The term ancient refers to the time span from the establishment of the first successful colony in Hawai‘i in about ad 800 (Athens et al. 2002, 57) to about ad 1800, selected arbitrarily for present purposes to approximate the time of effective transformative interaction with the non-Polynesian world.

Terminus Ante Quem

The chronological framework presented here is not intended as a general cultural sequence but rather is designed specifically for testing the Limited Land Hypothesis. This framework consists of agricultural expansion phases 1 and 2, the spans of which may vary from area to area. Agricultural expansion was a multi-scalar process, taking place at local to archipelagic scales simultaneously but at different rates and at different times. Expansion phase 1 often appears to have begun around ad 1300 in areas where staple production was dependent on rainfall and at least a century earlier in irrigable areas. The hypothesis proposes the year ad 1550 as the terminus ante quem, that is, the latest possible date marking the transition from the relatively rapid agricultural expansion of phase 1 to the significantly slower rate of phase 2. In most regions of Hawaii expansion seems to have continued until about ad 1800 (Allen 2001a, c; Kirch et al. 2004; Ladefoged et al. 2003; Mulrooney and Ladefoged 2005; Tomonari-Tuggle 2006).1 The chronology of the hypothesis is based in part on a revised version of the 14C frequency distribution curve constructed by Dye and Komori (1992a, b) and considered by them to be isomorphic with the archipelagic population curve. Among the 599 14C samples that were used in the 1990s to develop the Dye-Komori curve were some that were not controlled for provenance and composition factors including in-built age or the old-wood problem. In consultation with Tom Dye (see Dye 2000) , the revised version of the Dye-Komori curve used here advances the peak population value 100 years, from about ad 1450 to ad 1550 in an attempt to approximate a correction for

1For added perspective, counting by generations averaging 20 years (Hommon 1976; Stokes 1933), Kamehameha (ca. 1750–1819) was a member of generation 50, reckoned from the approximate birth year (ad 770) of a postulated founder of the first successful Hawaiian colony. 6 Robert J. Hommon these uncontrolled data. Between approximately ad 1550 and ad 1800 this revised curve indicates a very slow decline in population. In the discussion that follows if proxy data indicate a population exhibiting a low positive to low negative growth rate it is referred to as having near zero growth. A secondary reason for selecting ad 1550 as the approximate year of transition to a markedly slower rate of expansion has no cultural significance, but rather derives from a peculiarity of the radiocar- bon calibration curve. This chronological quirk can be summarized by considering hypothetical examples dating from three successive periods of 50 radiocarbon years each between 160±60 and 310±60 bp. Samples dating between approximately 260±60 and 210±60 ra- diocarbon years bp would yield calibrated ad ranges spanning more than five times as long a period as would sample pairs bracketing the immediately preceding or succeeding 50 radiocarbon-year periods. As a result of this quirk, a graphic plot of a small set of calibrated date ranges from sites occupied between ad 1400 and ad 1800 is likely to exhibit two clusters. The early end of the earlier cluster begins in the ad 1400s; that of the later cluster, in the ad 1600s. The apparent gap between the two clusters may vary from 100 to 200 years (Allen 2001c; Burtchard and Tomonari-Tuggle 2004; Coil and Kirch 2005; Dega and Kirch 2002; Dixon et al. 2002; Ladefoged et al. 2005; McCoy 2005; Tomonari-Tuggle 2006; Tuggle 1997).2 Researchers should be wary of assigning cultural significance to such a radiocarbon data gap without first identifying significant ar- chaeological data that would support such a determination, as do Coil and Kirch (2005, 79). For present purposes, the radiocarbon gap happens to facilitate testing of the Limited Land Hypothesis by conveniently sorting most evidence into early and later groups.

Agricultural Expansion

Testing the Limited Land Hypothesis requires a basic understanding of the dynamics of agricultural expansion. Research at an array of scales ranging from the entire archipelago through the Kohala Field System, spread across 21 ahupua‘a, to small pond-field (lo‘i) complexes shows that expansion proceeded in two general ways. The first of these can be called discontiguous expansion, a process in which cultivation is initiated in a location that is spatially separate

2The relatively short gap in the ‘Ewa Plain data (Tuggle 1997, 18, Fig. 10) probably is attributable to the larger size of the data set. Watershed: Testing the Limited Land Hypothesis 7 from previously cultivated land. The distance of separation tends to vary directly with the size of the research area. In the second process, contiguous expansion, each new increment of cultivated land is established adjacent to one or more existing plots. These two processes appear to have proceeded simultaneously in some instances and sequentially in others. The archaeological data in a given research area may support the division of the agricultural expansion processes into as many as three major stages, called here primary, secondary, and final. A portion of a research area in which one of these stages predominates can be called a primary, secondary, or final expansion zone. In some areas evidence may be sufficiently detailed to delineate sub-stages and sub-zones (e.g. Ladefoged and Graves 2007, 76–80). Generally speaking, cultivation of an area was initiated during the primary stage, increased in extent during the secondary stage, and extended to maximum limits during the final stage. Primary expansion zones were probably selected because they were considered by the Hawaiian planters to be superior to other nearby locations with regard to factors such as growing conditions or proximity to coastal settlements. Conversely, final zones appear to have been developed last largely because they were considered less desirable than nearby primary and secondary zones. Two research areas provide useful examples of agricultural expan- sion in this discussion. The rain-fed Kohala Field System of northwest Hawai‘i Island consists of a reticulated pattern of stone walls and trails that defines about 3,892 ha of roughly rectangular fields. The lo‘i system of Wailau Valley on the shore of Moloka‘i includes about 20 ha of irrigated pond-fields. Available information from these two field systems indicates that a research area is likely to exhibit multiple zones pertaining to each expansion stage. For example, sepa- rate secondary zones were often added adjacent to two or more sides of multiple primary zones. Ladefoged and Graves (2000), elaborating on earlier analyses (Cordy and Kaschko 1980; Kaschko 1973; Kirch 1984; Rosendahl 1972a), have determined that the Kohala Field System’s field boundaries were built in three successive, relatively-dated phases, which they term Temporal Units (TUs) 1, 2, and 3. These TUs, identified by Ladefoged and Graves (2000) chiefly on the basis of wall and trail abutments, appear to be broadly equivalent to the three expansion stages defined here. In the discussion that follows, we will assume that the area of 8 Robert J. Hommon cultivated land of each expansion stage was proportional to the length of its walls.3 The primary stage in a research area is identified chiefly by evi- dence of initial discontiguous expansion, that is, the establishment of major cultivation in primary zones distant from other contemporary cultivated complexes. It may be possible to identify characteristics such as high soil fertility that attracted initial settlement in these zones. On the archipelagic scale, primary expansion is evident in the establishment, between the ninth and twelfth centuries, of about 30 small, maximally discontiguous colonies in the salubrious cores of what eventually became the Hawaiian political districts such as Kona and H¯am¯akua (Hommon 1986, 67). The primary stage in the multi-community Kohala Field System is here assumed to be represented by the field walls built during TU1, of which 90 percent are concentrated in two large discontigu- ous complexes and one smaller one within the boundaries of eight of the research area’s 21 ahupua‘a as they existed in the nineteenth century (Ladefoged and Graves 2000). Each of these three clusters, which seems to satisfy the above definition of primary zone, proba- bly represents the inland component of one or more of the ancient local communities roughly equivalent to later ahupua‘a. These pri- mary zones may have been selected because they offered the best locally available growing conditions, though other factors, such as proximity to already existing shoreline settlements also may have been important (Ladefoged and Graves 2000, 439). The secondary zones in the Kohala Field System evince expansion into the spaces between the primary zones, along the uphill and down- hill margins of these zones, and into the territories of seven ahupua‘a at the north and ends of the research area that yielded little or no evidence of primary activity (Ladefoged and Graves 2000, 433, Fig. 7). During the final stage, like the second, expansion continued into previously underdeveloped spaces along the boundaries of already developed zones (Ladefoged and Graves 2000, 435, Fig. 8). Most notably, the final stage saw major expansion into the southernmost six ahupua‘a, which, according to Mulrooney and Ladefoged (2005, 49), were developed after 1550. Another element of the second and final

3This is a conservative assumption. It could be argued alternatively that on average a field bounded by primary stage (TU1) walls tended to be larger than one bounded by secondary and final phase walls, whose closer spacing tended to limit field size. Watershed: Testing the Limited Land Hypothesis 9 stages was the process of in-filling in which new field boundaries were added in spaces between preexisting ones. In-filling is usually considered a process of intensification rather than expansion. Recently, McElroy (2007, 150–164) has adapted the structural anal- ysis techniques developed for the rain-fed Kohala system to the irri- gated system of Wailau Valley. Evidence of three expansion stages were found in each of seven lo‘i complexes and two stages in another. The primary stage was represented in each complex by from one to three discontiguous primary zones, each of which contained from two to more than 30 lo‘i. The secondary expansion stage in the Wailau complexes entailed the addition of new sets of lo‘i in secondary zones contiguous to one or more primary zones. The sets of lo‘i in the final zones were contiguous with primary and/or secondary zones. On a larger scale, radiocarbon data have enabled McElroy (2007, 238, Fig. 7.13) to assign the founding of each of 13 recorded lo‘i systems in Wailau Valley to one of three temporal units (TUs). As in most other Hawaiian valleys the irrigable flood plains of Wailau tend to be relatively broad near the shore and to narrow with the distance inland. The expansion sequence along Wailau Stream, the larger of the two in the valley, began in the broader plains near the shore, and, after 1400, continued farther inland during TU 2 and TU3. Though this sequence may agree with expectations, the process of agricultural expansion was neither consistently linear nor unidirectional. We have already seen that expansion tended to resemble a patchwork rather than a line. In addition, the initial development along the smaller Kahawai‘iki stream in the eastern part of the valley was some 500 to 1,000 m inland, followed between ad 1400 and 1650 by expansion downstream.

The Late Expansion (LEX) Index

Testing the Limited Land Hypothesis in a research area requires a formal definition of what constitutes a significant reduction in the mean annual rate of agricultural expansion after ad 1550. For this purpose, we will assume that, as with most measures of economic growth, a reduction in the rate of agricultural expansion by at least 50 percent probably had marked socio-economic repercussions in ancient Hawai‘i and therefore satisfies the significance criterion. The equation for continuously compounded growth rate, r , is

ln n − ln n r = 2 1 (1.1) t2 − t1 10 Robert J. Hommon

where n1 = area at time t1, n2 = area at time t2, (with t2 > t1), and ln is the natural logarithm (Keyfitz 1966, 581). Calculating the Late Expansion (LEX) Index for a given data set yields a number that compares the mean annual expansion rate of expansion phase 2 (post-1550) with that of expansion phase 1 (pre- 1550). The LEX Index is calculated by dividing the expansion rate of phase 2 by that of phase 1. For example, if archaeological data indicate that the mean annual rate of expansion before 1550 was 2 percent and that after 1550 it was 1 percent, then the LEX Index would be 0.5, which is the threshold value for confirming the Limited Land Hypothesis. The smaller the LEX number for a particular research area, the greater the reduction in the post-1550 expansion rate, and therefore the stronger the support for the hypothesis. A LEX Index greater than 0.5 shows that the phase 2 rate of expansion was more than half that of phase 1 and thus disconfirms the hypothesis. A LEX value of 1.0 would signify that the rate of expansion in phase 1 continued unchanged in phase 2. The LEX Index of a delineated research area can be calculated given values of two temporal and three spatial variables. The five variables are as follows:

1. The maximum area under cultivation. Unless evidence indicates a withdrawal of land from cultivation, we can assume that the maximum extent of cultivation is equal to the total area of a research area bearing evidence of agricultural use.

2. The date of maximum expansion. Research in rain-fed areas including the Kohala Field System and Kahikinui District, Maui indicates that expansion continued until Western contact, ap- proximately ad 1800 (Ladefoged et al. 1998; Vitousek et al. 2004).

3. The area under cultivation at initial expansion. In most cases, evidence for the initial development of cultivation will be dif- ficult to identify because of its limited lateral extent and the likelihood of disturbance by subsequent development. For pur- poses of comparison we can assume that expansion was initiated with a specific proportion, such as one percent, of the eventual maximum area under cultivation.

4. The date of initial expansion. The consensus is that the major expansion of rain-fed systems such as the Kohala Field System Watershed: Testing the Limited Land Hypothesis 11

was under way by about ad 1400 (Allen 2001c; Carson 2005; Coil and Kirch 2005; Hartshorn et al. 2006; Hommon 1976; Kirch et al. 2004; Ladefoged and Graves 2000; Ladefoged et al. 1998, 2003; McCoy 2005; Mulrooney and Ladefoged 2005; Tomonari-Tuggle 2006). However, major development in rain-fed areas appears to have been preceded by expansion on a small scale beginning around ad 1300, within perhaps a century of the introduction of the sweet potato to Hawai‘i. Irrigation agriculture probably has a longer history in Hawai‘i than does the dry-land cultivation of sweet potatoes, so the sequence of initial expansion and major development in a given irrigable valley is likely to predate that in the rain-fed regions by one to two centuries. The 14C data in wet valleys are often inadequate to demonstrate whether or not irrigated systems were developed significantly earlier than rain-fed systems. In Wailau Valley, Moloka‘i, however, the 14C record reported by McElroy (2007, 202, Table 7.1) indicates that more than seventy percent of the investigated lo‘i systems were under cultivation by ad 1400.

5. The area under cultivation in 1550. The LEX Index varies in- versely with this quantity.

As evidence is accumulated, variable 5 often will be the last to be determined and thus will tend to be the key to testing the hypothesis in a given agricultural complex. Values for the other four variables of- ten can be drawn from previous research, as in the following example, which demonstrates the process of calculating the LEX Index using data from the Kohala Field System. Table 1 presents four hypothetical expansion scenarios for the Ko- hala Field System. According to Ladefoged and Graves (2000, 438, Ta- ble 1), the area that contains agricultural walls within the 21 ahupua‘a of the field system totaled 3,892 ha, which we will take to represent the maximum extent of cultivated land, variable 1. Available infor- mation indicates that the date of maximum expansion (variable 2) was ad 1800. We will estimate that in ad 1300 one percent of the maximum figure, or 39 ha, was under cultivation at initial expansion (variable 3), and that the date of initial expansion (variable 4) was ad 1300. Holding these four variables constant, the four scenarios in Ta- ble 1.1 illustrate how the LEX Index varies inversely with the hypo- thetical area under cultivation in ad 1550. The first two scenarios 12 Robert J. Hommon

Table 1.1 Hypothetical expansion scenarios in the Kohala Field System

Scen- Cultivated Land Area Mean Annual LEX ario Area (ha) P1∗:P2† Growth Rate Index ad 1300 ad 1550 ad 1800 P1 P2

1 39 1,350 3,892 1:2.3 1.42 0.42 0.3‡ 2 39 837 3,892 1:3.7 1.23 0.62 0.5§ 3 39 584 3,892 1:5.7 1.08 0.76 0.7¶ 4 39 389 3,892 1:9 0.92 0.92 1.0k

∗P1 = expansion phase 1, ad 1300–1550. †P2 = expansion phase 2, ad 1550–1800. ‡ Strong confirmation of the hypothesis. § Confirmation of the hypothesis at the threshold level. ¶ Disconfirmation of the hypothesis. k Disconfirmation of the hypothesis with unchanged expansion rate. would confirm the Limited Land Hypothesis; the second two would disconfirm it. Scenario 2 shows that the threshold LEX Index of 0.5 would be reached if 837 ha, or 21.5 percent of the maximum cultivated area, were being cultivated in ad 1550. A full confirmation of the hypothesis requires determining the LEX Index of a research area. However, if project funding or time limitations do not allow full confirmation, a preliminary or minimal confirmation requires demonstrating only that the LEX Index is equal to or less than 0.5. Testing for preliminary confirmation can be tightly focused on particular sections that have been identified as likely areas of phase 1 cultivation. Identification of at least two levels of expansion zone (for example, primary and secondary) in effect stratifies the sample and greatly reduces the effort necessary for hypothesis testing. The primary zones of the Kohala Field System as identified above are a case in point. Currently available absolute dating results are insufficient to determine the ages of these zones. However, given the widely-accepted view that large-scale expansion in rain-fed complexes began around ad 1400, it seems very likely that the area brought un- der cultivation during the century and a half between 1400 and 1550 extended to at least the 21.5 percent of the total field system calculated in scenario 2 of Table 1.1. Under the conservative assumption that the area of cultivated land of each expansion stage was proportional to the length of its walls, absolute dating analyses demonstrating that the primary zones, which comprise 20 percent of the system, were Watershed: Testing the Limited Land Hypothesis 13 under cultivation by 1550 would alone nearly suffice to confirm the hypothesis. Minimal confirmation would require dating an additional 1.5 percent (63 ha) of secondary zone lands to phase 1. A full con- firmation of the hypothesis might be accomplished with additional testing in secondary zones, since data from the southern part of the system indicate that the final zones date from phase 2. It is important to emphasize that Scenario 2 in Table 1.1 demon- strates that the phase 2 rate of expansion was significantly reduced even though expansion during phase 2 brought fully 3.7 times as much land under cultivation as that developed in phase 1. The key to under- standing this conclusion, which may at first seem counter-intuitive, is to observe that the process of agricultural expansion proceeded exponentially. Because agricultural lands supplied most of the caloric needs of the Hawaiian population, the process of agricultural expan- sion was related in often complex ways to the exponential growth of population throughout the Hawaiian sequence. Just as populations consistently grow or decline exponentially rather than arithmetically (Sutton and Molloy 1989, 32), so Hawaiian agricultural expansion must have proceeded exponentially by virtue of its intimate connection with population growth.

Part 2. Supporting the Hypothesis

To justify the effort and expense required for archaeological hypothe- sis testing as described in Part 3, the Limited Land Hypothesis should be shown to be plausible. Part 2 presents natural, archaeological, and ethnohistorical evidence for wide ranging changes in Hawaiian culture under way by ad 1550 that seem to be associated with a significant decline in agricultural expansion as postulated by the hypothesis. Part 2 has three divisions. The first division responds to the view, contrary to the Limited Land Hypothesis, that Hawaiian agricultural expansion had not approached significant limits before Western con- tact. The second division discusses aridity, soil infertility, and other environmental factors that limited agricultural expansion. These discussions have benefited from recent multi-disciplinary work by Hawaiian Biocomplexity Project participants in Kahikinui, Maui and the Kohala Field System, Hawai‘i Island (Hartshorn et al. 2006; Kirch 2003; Kirch et al. 2004; Ladefoged and Graves 2000; Ladefoged et al. 1998, 2003; Mulrooney and Ladefoged 2005; Vitousek et al. 2004). 14 Robert J. Hommon

The third division refers to indirect lines of evidence for a reduced rate of agricultural expansion following ad 1550. These include tra- ditional famine foods and methods of population control, reduced production, near-zero population growth, the economic roots of warfare, and the emergence of Hawaiian archaic states.

“Only the best of the arable land”: Responses to Arguments for Unreduced Expansion

Contrary to the filled land view expressed in the Limited Land Hypoth- esis has been the notion, pertaining primarily to irrigated agriculture, that before Western contact

the Hawaiian Islands were not completely settled. Only the best of the arable land, capable of cultivation by the gardening methods practiced by the natives, was actually utilized. Compared with areas like Peru and the Philip- pines which were terraced up to high altitudes, land and water resources in Hawai‘i were by no means fully devel- oped. (Handy and Handy 1972, 280)

In his study of Halele‘a District, , Earle remarks on the “un- derutilization of agricultural resources” and goes on to state:

Even the alluvial soils were greatly underutilized as shown by the later, tremendous expansion in irrigation agricul- ture for rice. The inattention to kula farming and the restriction of irrigation farming cannot be explained on technological grounds . . . Also, as seen by the cut stone aqueduct in Waimea, Kaua‘i, and as seen by the exten- sive complexes on the Napali coast, Kaua‘i, the Hawaiian agriculturist had the considerable technological sophistication necessary to extend irrigation agriculture (Handy and Handy 1972, 280–282). Yen (1973, 79) empha- sizes that Hawaiian irrigation agriculture, in contrast to other Oceanic areas like New Caledonia, was “simple” (non- intensive) on the basis that valley slopes and ridges were not regularly terraced. (Earle 1978, 163–164)

Earle elaborates further on later land use.

The total area of irrigated land [in Halele‘a District] in 1850 is estimated to have been 115 ha. In comparison, by using Watershed: Testing the Limited Land Hypothesis 15

longer ditches with lesser slopes, the Chinese rice farmers had an estimated 526 ha in production at the turn of the century. (Earle 1978, 106)

Responses to the Handys’ and Earle’s views can be summarized as follows:

1. Handy and Handy’s statement that “[o]nly the best of the arable land . . . was actually utilized” (Handy and Handy 1972, 280) is contradicted by archaeological evidence (Allen 1987a; Bennett 1931; Earle 1980; Hommon 1970; Hommon and Barrera 1971; Kirch 1994; Mills 2005). This point is discussed further in the section “Filled Alluvial Lands” on page 19.

2. Earle’s comparison of taro lo‘i lands with rice-growing areas is based on the extensive records of the Mahele¯ of the mid- nineteenth century. According to my calculations (Appendix A), by 1850, the time of these land records, the Hawaiian pop- ulation had been reduced by more than 80 percent. Though the magnitude of population loss in Halele‘a is unknown, it was probably substantial, so I suggest that the extent of land under cultivation before contact was significantly greater than that recorded in the Mahele¯ documents. If the abandonment of Halele‘a agricultural land between ad 1778 and ad 1850 was proportional to a population loss of 80 percent then the lo‘i at contact roughly matched the extent of the later rice fields. While rice farmers expanded Halele‘a agricultural lands by building longer ditches, some of their lands had formerly been planted in taro, as noted by Handy (1940, 72) and by Handy and Handy:

The flats in which rice was planted by the Chinese had been the taro lo‘i of the Hawaiians, amply irrigated by ditches from the Hanalei River. (Handy and Handy 1972, 420)

3. Limits on suitability of land for cultivation as perceived by ancient Hawaiians may not be readily obvious to us. The “inat- tention to kula farming” mentioned by Earle, and Handy and Handy’s assertion that “only the best of the arable land” was used should be considered in light of the findings of the Bio- complexity group (Kirch 2003), as described on page 18 in the “Slopes Without Systems” section, suggesting that leaching of soil nutrients tended to preclude extensive agriculture on the 16 Robert J. Hommon

slopes of all the islands other than Hawai‘i and western Maui (Hartshorn et al. 2006, 11092).

4. The absence of high slope terracing as evidence of agricultural underdevelopment is much less significant than Handy and Handy believed if considered in a geological context. The ter- rains into which the terrace systems of Peru, the Philippines, and New Caledonia were cut are considerably more ancient, and therefore more weathered and more amenable to terracing than are the steep, adamantine basalt mountainsides that define the youthful valleys of Hawai‘i.

5. Necessity may be the mother of invention, but she is never a single parent. In the regions with hillside terracing mentioned above, the significantly longer settlement histories, the internal interaction of various ethnic groups, and the relatively frequent external contacts increased the likelihood of indigenous and im- ported technological innovation as compared with Hawai‘i, with its geographic isolation, short cultural sequence, and relatively brief period of maximum population density.

6. The postulated rejection of technological innovations such as the dressed stone aqueduct (of which K¯ık¯ıaola at Waimea is the sole example) or elevated irrigation ditches does not in itself demonstrate that expansion did not approach “real” or perceived limits of usable land. Hawaiians may have favored dealing with such limits in other ways, such as population con- trol, as discussed below. It is also possible that the K¯ık¯ıaola aqueduct was a recent innovation that had not yet had a chance to disseminate to other regions when Vancouver documented it in 1792 (Mills 2005; Vancouver 1798).

7. Competing priorities might also have precluded adoption of innovations such as the stone aqueduct even though their tech- nology and benefits were appreciated. For example, perhaps Halele‘a residents rejected the idea of opening up new lands to cultivation by means of a new system with a high-intake aqueduct because it would have fostered community discord by disrupting traditional water distribution rights in the valley. Watershed: Testing the Limited Land Hypothesis 17

Environmental Limits on Agricultural Expansion

Aridity

Any evidence that a boundary of an ancient agricultural complex corre- sponds to a known or suspected natural barrier to reliable production may support the hypothesis. Such evidence would tend to indicate that agricultural activity had expanded to the maximum extent possi- ble along that boundary. The more extensively such boundaries match natural barriers, the more likely it is that the rate of expansion in the complex had slowed or ceased. Expansion of sweet potato cultivation to zones offering minimal soil moisture might indicate that better watered lands were already in use. Beginning in the 1970s, intensive surveys in dry, leeward regions identified lower slope zones with sparsely distributed sites situated between the much denser coastal habitation concentrations and extensively-developed inland agricultural zones. These barren or intermediate zones were consistently attributed to annual rainfall insufficient to support dry-land crops (Hommon 1980; Smith and Schilt 1973), an observation that is supported by general horticultural sources (Kay 1973; Norman et al. 1984; Purseglove 1968). The fact that the lower boundary of the Kohala Field System tends to approximate the 20 in. isohyet was observed more than 35 years ago (Murabayashi 1970; Newman 1970) and has been confirmed during recent extensive research (Ladefoged and Graves 2000; Ladefoged et al. 2003). In some cases, the combined effects of natural conditions limited agriculture. For example, in Kahikinui

[t]he rapid drop-off in settlement density <400 m eleva- tion is directly related to the aridity and low water-holding capacity of the soil; this combination would have reduced soil moisture to levels below that required for sweet potato growth . . . The fact that a sizable population did perma- nently use Kahikinui’s resources, despite its marginality for agriculture, testifies to the demographic and economic pressures building within Hawaiian sociopolitical systems in the centuries immediately before European contact. (Kirch et al. 2004, 9940) 18 Robert J. Hommon

Drought Risk

The closer sweet potato cultivation approached the 20 in. annual rainfall isohyet the greater the risk of crop failure due to frequent and severe drought (Hommon 1976; Nakamura 1933; Ripperton and Hosaka 1942). In any given complex or region, the fact that farmers expanded cultivation into the zone where failure was known to be frequent is consistent with the notion that better watered lands were no longer available for exploitation, which in turn signals a significant reduction in the rate of agricultural expansion. Examples of exploitation of high risk zones are to be found in the Kohala Field System (Ladefoged et al. 1998, 868, Fig. 5), Kawela, Moloka‘i (Weisler and Kirch 1985), and probably the Kekaha kalana (major sub-district) of Kona (Dye et al. 2002).

Leached Soils

The upper elevation boundary of evidence of rain-fed agriculture in the Kona Field System (Newman 1970, 127–128), the Kohala Field Sys- tem (Ladefoged et al. 1998; Vitousek et al. 2004), and Kahikinui tends to be at an altitude of about 800 m (2625 ft.). Recently the Hawai- ian Biocomplexity group has found that while too little rain largely determined the lower elevation boundary of rain-fed crop lands, too much rain can be a limiting factor as well. Evidently, the 800 m upper boundaries of cultivation in Kohala and Kahikinui correspond to re- duced soil fertility resulting from loss of soil nutrients due to leaching, as well as higher soil acidity and anoxia (Hartshorn et al. 2006; Kirch et al. 2004). As in the case of lower boundaries determined by aridity, the fact that agriculture had extended all the way to these upper boundaries imposed by low fertility indicates a significant reduction in the rate of agricultural expansion in these regions.

Slopes Without Systems

Any discussion of limits on agricultural expansion must address the question of why large regions of the islands were never subjected to extensive development. The limits on lands suitable for agricultural development are in large part defined by the areas that exhibit little or no evidence of such activity. The undeveloped barren or interme- diate zones that lie adjacent to the Kona and Kohala Field Systems have been attributed to aridity. Aridity itself does not explain why large-scale integrated systems like these of leeward Hawai‘i Island Watershed: Testing the Limited Land Hypothesis 19

apparently never developed in other islands’ leeward regions that resemble those of Kona and Kohala in both slope and rainfall. On the basis of investigations in the rain-fed agricultural complexes on soils formed on the geologically young Hawai‘i Island and eastern Maui Vitousek et al. (2004, 1668) have reached the conclusion that “low soil fertility precluded the development of large-scale intensive dryland agricultural systems on stable upland surfaces on the older islands of the Hawaiian archipelago.” Or, as Hartshorn et al. (2006, 11092) put it, “[g]ently sloping, stable surfaces on the older volcanoes ( Maui to Kaua‘i) are too infertile because of long-term nutrient leaching.” Other factors being equal, the older the region geologically, the less fertile the soil. A seeming exception that actually supports this rule is the 900 ha field system of Kalaupapa, Moloka‘i. Though Kalaupapa peninsula is connected to the 1.5 million-year-old East Moloka‘i volcano, it was formed by the eruption of the 330,000 year-old Kauhako¯ volcano (Clague 1998; McCoy 2005). The soils formed on the Kauhako¯ flows are young enough not to have lost their fertility to leaching. On the archipelagic scale then, soil infertility due to age appears to have been the major barrier that limited agricultural expansion. As a result, huge expanses of the Hawaiian islands evidently can be eliminated when we seek to answer the question of whether the rate of agricultural expansion in Hawai‘i had slowed. Accordingly, research should focus on geologically young lands, as well as regions with rejuvenated soils (see the section “Flood-Water Systems” on page 20).

Filled Alluvial Lands

Unlike the often fuzzy boundaries of dry-land systems, those of irrigated lo‘i systems tend to be sharply defined by slope and elevation relative to water sources (Allen 1987a; Earle 1980; Hommon 1970; Hommon and Barrera 1971; Mills 2005; Riley 1975). Of Kaua‘i, Bennett wrote:

The impressive of the agricultural terracing is its tremendous extent. In the valleys in which little distur- bance has gone on, particularly the Napali section, the maximum of tillable soil was utilized. Even a 10 foot square of soil among the great mass of lava rock will be cultivated. (Bennett 1931, 21) 20 Robert J. Hommon

Kirch (1994, 254) observed that “in the well-watered western Hawai- ian Islands, most of the broad alluvial floodplains had been modified for pondfield irrigation by the time of European contact.” As in dry-land areas, the tendency for lo‘i systems to extend to the lim- its imposed by natural features including cliffs, steep slopes, and stream beds is consistent with an expansion process that had slowed significantly before Western contact. Revised frequency distribution curves of 14C samples from agricultural sites, mainly from O‘ahu, show a spike around ad 1550, suggesting that by that time primary forest-clearing, which presumably produced more carbon than did maintenance activities, may have been completed (Allen 1992; Dye 1994; Williams 1992). The results of this analysis are consistent with the Limited Land Hypothesis. If additional problem-oriented research as outlined in Part 3 were to produce similar results, they would provide strong confirmation for the hypothesis. Once, Anahulu Valley, O‘ahu appeared to be a major exception, perhaps one that disproved the rule. Research published by Kirch and Sahlins (1992) seemed to indicate an absence of intensive use of the valley before Western contact. A lack of development in a valley so well suited for the growing of irrigated taro suggested a history of agricultural development radically different from that of other studied Hawaiian regions. Recently, Dega and Kirch have revisited the results of Kirch’s and Sahlins’ earlier research and have decided that, contrary to previous conclusions, Anahulu had been well settled earlier but that

one outcome of [the post-contact] extensive reworking of the mid-valley landscape was the obscuring of the surface evidence for residential and agricultural practices deriv- ing from the century or so immediately before European contact. (Dega and Kirch 2002, 112)

Flood-Water Systems

Agricultural systems of ancient Hawai‘i are conventionally classified as wet or dry, that is irrigated or rain-fed. A third kind of complex, seldom given its due—the flood-water system—was distinct from either wet or dry though it combined features of both. In his study of the Indians, Hack (1942, 26) defined flood-water fields as those “watered by surface runoff.” Watershed: Testing the Limited Land Hypothesis 21

In flood-water farming the crops are watered by the runoff of an area much larger than the area actually cultivated. The moisture is stored in the soil between floods and provides a constant water supply for the plants except during long droughts. (Hack 1942, 26)

In Hawai‘i the archaeological evidence for flood-water systems is found on or adjacent to slight to moderate slopes and small inter- mittent water channels. These systems consist of earthen terraces and stone diversion walls, dams, mounds, and alignments that appear to have served to guide, impede, and spread surface water and sus- pended sediment during and following rainstorms. The constituent structural features tend to be somewhat amorphous and were evi- dently improvised to fit each local micro-environment. Numerous flood-water systems were identified along the colluvial slopes of Makaha Valley, O‘ahu (Hommon 1969, 1970). Among other exam- ples of flood-water systems are those reported at Kawela (Weisler and Kirch 1985, 138–139) and Kalaupapa (McCoy 2005, 340), Moloka‘i and an area at the boundary of Kona and Kohala districts, Hawai‘i Island (Rosendahl 1972b, 86–91). Several of the non-irrigated complexes mapped and described in H¯alawa Valley, Moloka‘i by Riley (1975) and some of the unpaved terrace complexes on Nihoa and Necker (Emory 1928) also probably were flood-water systems. With the exception of Kalaupapa, the older soils of these areas generally tend to be leached of soil nutrients, as discussed above. However erosive processes, in particular mass wasting of cliff faces such as those of Makaha, are continually depositing newly exposed material still relatively rich in nutrients. By constructing flood-water systems on colluvial slopes, Hawaiian farmers not only supplemented limited rainfall with slope-wash, but also trapped rich alluvium in which to grow their crops. Any dry-land system that features terraces or other water man- agement structures is to a degree a flood-water system, even if well supplied by direct rainfall, but in the strict sense the term most use- fully applies to systems which depended for success on managed surface flow from a relatively large catchment area. The advantages of increased water supply and richer soils in flood- water systems were probably accompanied, especially on steeper slopes, by increased maintenance requirements and greater risk of crop-loss due to flood damage during unusually heavy rainstorms. The extensive development of flood-water systems tends to support 22 Robert J. Hommon the Limited Land Hypothesis in that it indicates that areas better suited to lower-risk reliable agricultural production were no longer available for development.

Indirect Evidence for Limits on Agricultural Expansion

Famine Foods

Traditional knowledge of a variety of wild plants identified as famine foods tends to support the notion that commoners recognized that, at least in some areas, food shortages were relatively frequent and unpredictable (Handy and Handy 1972; Kamakau 1961; Malo 1951; Schmitt 1970a). The maintenance of such knowledge suggests that shortages could not be alleviated readily by relocating families or communities to more productive areas (Hommon 1976, 293–294), a restriction that is consistent with agricultural expansion that had approached the limits of arable lands. Widespread knowledge of famine foods may well be a consequence if, as the evidence alluded to above indicates, increasing populations had expanded into lands where crop failure was brought on by frequent drought such as those described in Maui’s Kula District (Kamakau 1961, 23–24) or on Ni‘ihau (Portlock 1789; Vancouver 1798). This evidence is not presented to imply that famine was necessarily either frequent or widespread or was a significant population control process in ancient Hawai‘i (see the section “Population Control” on page 24).

Adze Production

According to McCoy (1990, 1999) the period of peak production at the Mauna Kea adze quarries spanned the fifteenth and sixteenth centuries. The fact that this period closely approximates estimated phases of rapid population growth (Dye 1994; Dye and Komori 1992b) and agricultural expansion into virgin lands (Allen 1992) suggests that beginning around ad 1400 the adze-making on Hawai‘i Island was ramping up to meet the burgeoning demand for for land clearing as well as building and other craft work (Handy and Handy 1972; Hommon 1986, 1992). With regard to the Limited Land Hypothesis, the apparent reduced demand following the sixteenth century is consistent with diminished land-clearing activities during the last two centuries before Western contact as limits of arable land were approached. Watershed: Testing the Limited Land Hypothesis 23

Population Decline

Dye and Komori (1992b), following Rick (1987), have developed popu- lation growth estimates for ancient Hawai‘i based on the observation that, given certain assumptions about the deposition and sampling of archaeological carbon, “a frequency distribution of 14C age deter- minations [can be considered] isomorphic to a population curve.” If we accept this isomorphism, the 14C-constructed curves based on various data sets from all islands except Ni‘ihau and Lana‘i consis- tently indicate that population tended to grow relatively slowly at first, then more rapidly, and then approximated zero until contact (Allen 1992; Dye and Komori 1992b; 1992; Streck 1992; Tuggle 1997; Williams 1992). As noted in Part 1, for the present study the point at which population may have ceased to grow, has been shifted forward 100 years from approximately ad 1450, as reported by Dye (1994), to ad 1550. Holding other variables, such as marine-food productivity, con- stant we should expect that the ancient Hawaiian rate of population growth would vary directly with agricultural productivity. The cessa- tion of population growth during expansion phase 2 (ad 1550–1800), as indicated by the revised Dye-Komori population curve, is consistent with a deceleration of agricultural expansion during this period.

A Population Growth Rule of Thumb

One way to assess the plausibility of the Limited Land Hypothesis in a research area where the area of cultivated land has not yet been determined is to compare 14C-based evidence for population growth rates in expansion phases 1 and 2 with hypothetical population fig- ures that represent positive growth in phase 1 and near-zero growth in phase 2. The application of what is called here the population growth rule of thumb to five research areas, including Kealakekua and Waimea, Hawai‘i; Kahikinui, Maui; Kalaupapa, Moloka‘i; and the ‘Ewa Plain, O‘ahu is discussed in Appendix B. The results of this test indicate near-zero population growth dur- ing expansion phase 2 in all five areas. If we assume that agricultural expansion tended to vary directly with population growth, then the 14C data provide provisional support for the Limited Land Hypothesis in these areas. 24 Robert J. Hommon

Population Control

Demographic models of ancient Hawai‘i have consistently exhibited near-zero population growth during expansion phase 2 (Dye 1994; Dye and Komori 1992b; Hommon 1976; Kirch 2007). Though localized food shortages and pre-Cook introduction of Eurasian disease as hypothe- sized by Dye (1994, 13–15) may have contributed to population decline, there is no evidence of an archipelago-wide health catastrophe. Rather, I suggest that as agricultural expansion approached nat- ural and culturally-imposed limits, Hawaiian commoners may have achieved near zero-growth by intensifying population control prac- tices. According to Pukui et al.,

[t]he wish for descendants, so typical of Hawaiians, and present, does not mean the Hawaiian tradition upheld uncontrolled child-bearing. Large families were wanted— but babies should be spaced out. Hawai‘i had measures designed to control conception and ways of thought to induce conception. Both made use of prayer and ritual, herbs and potions, and all but forgotten lore handed down in families. Where traditional birth-spacing is concerned, we know more about cultural attitudes than specific means. (Pukui et al. 1972, 2:100)

One of these means was evidently the extended nursing of infants, which reduced the number of births by lengthening the interval be- tween them (Pukui et al. 1972, 2:100–101). Though there is no direct evidence for this practice in Hawai‘i before Western contact, it has been widely observed among ranging from hunter-gatherer bands to modern nation-states. In his comparative study of states ap- proximating the size and socio-political complexity of ancient Hawai‘i, Trigger (2003, 310–311) cites instances of extended nursing as a birth- control device among the Yoruba of West Africa and the people of ancient Egypt, Mesopotamia and the Valley of Mexico.

Roots of War

A variety of reasons including the accumulation of prestige, rebellion against oppression, revenge, love of battle and even boredom have been cited as triggers for war in ancient Hawaii (Fornander 1969; Kamakau 1961; Pukui et al. 1972). Ultimately however, major wars in Hawai‘i as elsewhere were led by elites to maintain or project Watershed: Testing the Limited Land Hypothesis 25 power over land and other resources by controlling large-scale behavior. This view is underlined by the unsubtle epithet of Ku-k¯ ¯a‘ili- moku, the hereditary war-god of Hawai‘i Island: “Ku¯ island-snatcher” (Pukui and Elbert 1973). Although stories of earlier chiefs tell of occasional raids and bat- tles, according to traditional accounts and genealogical dating it is not until about ad 1600 that conquest warfare appears in Hawaii (Hommon 1976, 153–160, 315–331). Wars of conquest were waged to capture and extend control over entire districts or islands, complete with resident producers, a goal that entailed replacing and sometimes sacrificing defeated rulers. I propose that the prevalence of conquest warfare during expansion phase 2 was in large part a response to the disruption of economic growth brought about by the hypothesized phase 2 reduction in the rate of agricultural expansion. War, seen from the perspective of the Limited Land Hypothesis, was effectively a continuation of agricultural expansion by other means. Ancient Hawaiian chiefs tended to experience variations in pro- duction of the lands they administered most directly in the form of the relative abundance of food and other necessities (e.g., mats, kapa, malo and pa‘¯ u¯) and prestige goods contributed as taxes and offerings by the commoners on those lands. The quantity, quality, and consistency of such revenues constituted the wealth that enabled chiefs to support retinues and to satisfy superiors and subordinates, thereby maintaining prestige within the elite community. For some 750 years, beginning with the first successful colony around the beginning of the ninth century, Hawaii’s was a growth economy, sustained by a burgeoning population laboring to transform natural ecosystems into productive croplands at exponential rates. The revised Dye-Komori curve indicates that in the span of the single century preceding ad 1550 exponential growth doubled the popula- tion, that is, it added a population increment equal to that generated over the previous 650 years. During that century the rapid exponential increase of agricultural production that satisfied the growing popula- tion’s caloric requirements also presumably provided the chiefs with a most spectacular growth in wealth. As the benefits of such growth must have been obvious to chiefs at the time in the form of increasing revenues, so, it seems, chiefs of a few decades later probably began to perceive the effects of slowing growth and increasing uncertainty. Whether or not individual chiefs experienced the shift directly or in tales of a former prosperity told by their elders, traditional historical accounts show that for the chiefly 26 Robert J. Hommon establishment the prospect of revenue enhancement was a sufficient reason for going to war, as shown in the following four examples from Hawai‘i Island traditional literature. The first example is found in Kamakau’s account of a war between Kanaloa-kua‘ana and ‘Umi-o-ka-lani, two sons of Keawe-nui-a-‘Umi in the early ad 1600s.

Kanaloa-kua‘ana’s old priests and aged counselors passed urine like water, a sign of a chief without wealth . . . [When they urinate in the presence Kanaloa-kua‘ana, they say] “Your priests, counselors, and people have clear urines because they drink copiously of water. That is because you are a poor chief. The urine of chiefs, priests and counselors of a wealthy chief is yellow through drinking ‘awa and eating rich foods. Their lights never go out at nights.” [Their nights are spent in enjoyment.] The chief said “what must I do?” “Make war on ‘Umi-o-ka-lani and take the whole kingdom to yourself.” (Kamakau 1961, 46)

Following their advice, Kanaloa-kua‘ana waged war against his half-brother and eventually defeated him. Whether this victorious chief augmented his wealth thereby Kamakau does not say, but such a result appears to have been his intention. The second example spans the middle to late seventeenth and the early eighteenth centuries, when Hawai‘i Island was usually divided between two competing polities, one an alliance of Hilo and adjacent windward districts and the other consisting of Kona and adjacent leeward districts (Hommon 1976, 135, Fig. 13, 319–320). In the following passage, the intermittent war between the two factions is described in economic terms:

There was much fighting between the chiefs of Kona and Hilo, but neither was defeated. The chiefs of Kona de- sired Hilo, H¯am¯akua, and La‘a because of the mamo and ‘o‘o feathers, the war , and fine tapas such as the ‘o‘uholowai, ‘eleuli, pala‘a, and kalukalu of Waipi‘o. The chiefs of Hilo in turn desired warm food and drinking water, and tough and tender fish. Those were the wealth of Kona. (Kamakau 1961, 62) Watershed: Testing the Limited Land Hypothesis 27

The third example dates from the late eighteenth century when I-maka-koloa, chief of Puna District, rebelled against Kalaniopu‘u, ali‘i nui of Hawai‘i Island and

seized the valuable products of his district, which con- sisted of hogs, gray tapa cloth (‘eluli), tapas made of mamaki bark, fine mats made of young pandanus blos- soms (‘ahu hinalo), mats made of young pandanus leaves (‘ahuao), and feathers of the ‘o‘o and mamo birds of Puna. (Kamakau 1961, 106)

In the fourth example, following his re-conquest of Maui, Moloka‘i, Lana‘i and Kaho‘olawe in 1795, Kamehameha appointed as governor of Maui Ka‘ahumanu’s brother Kahekili Ke‘eaumoku, who

imposed a tax upon Maui, Molokai, Lanai, and Kahoolawe, and food piled up so the chiefs and people ate until they could eat no more. Still, there were quantities left over; so they gave the chief the name of “Great heap” (Pu‘u-nui). (Kamakau 1961, 198)

I suggest that if these examples of warfare during expansion phase 2 are considered to be expressions of economic competition then they tend to support the Limited Land Hypothesis. After about ad 1550 powerful chiefs who could no longer depend on constant, large-scale agricultural expansion and concomitant population growth for wealth enhancement began to turn to force to seize the productive lands of rival chiefs. Additional support for the economic focus of Hawaiian warfare can be found in the numerous examples of weakening of enemies’ productive capacity through destruction of their crops and agricul- tural infrastructure such as irrigation ditches (Arago 1823; Hommon 1972; Menzies 1920; Stokes 1937; Vancouver 1798). In war considered as a zero-sum game, the destruction of productive capacity of enemy territory, even if it is to be annexed, is a gain for the aggressor.

The Emergence of Hawaiian Archaic States

As I have discussed at length elsewhere, late pre-contact Hawaiian polities clearly exhibit the defining features of what are currently termed archaic states, societies ancestral to the highly successful and dangerous nation-states that now encompass the globe (Hommon 28 Robert J. Hommon

1972, 1973, 1976, 1986, 1999, 2000, 2001, 2005, 2007, 2008). The most all-pervasive revolution in the history of human society began with the emergence of primary or pristine archaic states in regions such as Egypt, Mesopotamia, and . A primary state is one that developed indigenously in the absence of contact with any pre- existing state. Hawai‘i was one of the handful of examples of primary state emergence, though, with rare exceptions (e.g. Allen 1991) , this development has seldom been noted by other investigators until recently (Kirch 2005). In some respects the Hawaiian kingdoms of the late eighteenth century are the most thoroughly documented of such states, having been observed and described in detailed written accounts by both native participants and foreign visitors. In a collection of seminar papers dealing specifically with archaic states, Marcus and Feinman present a recent version of the defining criteria as follows:

Although definitions varied among participants, many be- lieved that in contrast to modern nation-states, archaic states were societies with (minimally) two class-endoga- mous strata (a professional ruling class and a commoner class) and a government that was both highly centralized and internally specialized. Ancient states were regarded as having more power than the rank societies that preceded them, particularly in the areas of waging war, exacting tribute, controlling information, drafting soldiers, and regulating manpower and labor . . . For some well-known states where texts are available, one could add to this the stipulation that archaic states were ruled by kings rather than chiefs, had standardized temples implying a state religion, had full-time priests rather than shamans or part- time priests, and could hold on to conquered territory in ways no rank society could. For less well known states, where texts are absent, perhaps the best definition is the most general and simple, so as to encompass marked historical variability. (Marcus and Feinman 1998, 4–5)

This definition describes Hawaiian polities of the late eighteenth century so well that it seems almost to have been composed for that purpose. The following summary applies the defining features listed by Feinman and Marcus to Hawaiian data: Watershed: Testing the Limited Land Hypothesis 29

• The well-documented endogamy of the two major socioeco- nomic classes in part seems to have been a result of the protec- tion of chiefly prerogatives and the desire of chiefs for hyperga- mous unions to maintain the elite status of offspring (Kamakau 1961; Kirch and Sahlins 1992). The marked contrast between the lives of the locally-focused, economic producers relatively lacking in political power and the powerful elites is exemplified by the frequency of inter-island marriages among chiefs in con- trast with the tendency for commoners to marry within their own local community (Hommon 1975, 1986; Linnekin 1990).

• The government of a was a permanent in- stitution consisting of a stratified control hierarchy of offices including, at a minimum, ali‘i nui, ali‘i ‘ai moku, and ali‘i ‘ai ahupua‘a. Contrary to the proposition of Friedman (1981, 289) that large-scale Polynesian polities cycled between consolida- tion and disintegration, the traditional histories point to a clear though uneven trend toward increase in size and reduction in number of polities. The traditions also show that when measured in terms of frequency of usurpation the Hawai‘i Is- land polity from ad 1400 to 1800 was somewhat more stable than the kingdom of England during the same period (Hommon 2000, 145–146).

• Hawaiian kings held supreme authority and exercised the unique right to declare and end wars by consecrating luakini heiau (state temples) with human sacrifices, to redistribute administrative offices at the time of accession, and to receive taxes, including those collected during the Makahiki festival, and distribute them among the members of the royal court and bureaucracy (Malo 1951, 143, 159–160).

• The “ability to hold on to conquered territory” (Marcus and Feinman 1998, 5) was amply demonstrated by the tendency, beginning around ad 1600, for the larger Hawaiian polities to expand by annexing other polities in whole or in part (Hommon 1976, 133–136, Fig. 13; 315–331). Though Kamehameha was a brilliant and innovative strategist who eventually established a kingdom that united the whole archipelago, his career as a warrior king followed a political path established long before he was born. When Kamehameha was still battling rivals for control of his own island, Kahekili of Maui, pursuing a tradition 30 Robert J. Hommon

of conquest already nine generations long, managed to unite all the other islands by means of war and diplomacy (Fornander 1969; Hommon 1976).

• Hawaiian kings applied their stratified control bureaucracies to tasks including “waging war, exacting tribute, controlling information, drafting soldiers, and regulating manpower and labor” (Marcus and Feinman 1998, 4–5) as discussed in more detail below. Nineteenth century sources indicate that each ‘ili ‘aina¯ subdivision of an ahupua‘a community was rated on the basis of the number of able-bodied commoners it could provide for corvée labor4 or an appropriate amount of goods (Kirch and Sahlins 1992; Whitman 1979). Corvée laborers constructed civil works including large aquacultural ponds and temples. (Allen 2001c, 148) speculates, though it is “not empirically demonstra- ble,” that the long stone mounds (kuaiwi) of the 13,900 ha Kona Field System may have been built under the management of ‘Umi a Liloa, ruler of Hawai‘i Island around ad 1600. Hawaiian rulers also maintained intelligence and communication systems employing couriers who were selected for their swiftness of foot and trustworthiness and who doubled as spies (Hommon 1975, 103–112).

• Accounts of early visitors and Hawaiian scholars clearly describe ceremonies of the Hawai‘i Island state religion and other regional variants that were conducted in standardized temples by orders of professional priests (Kamakau 1961; Malo 1951; Valeri 1985).

The late pre-contact Hawaiian kingdoms, like all states, were inher- ently large-scale societies capable of encompassing extensive territo- ries and large populations. At the time of Captain Cook’s second visit to the islands, Kalaniopu‘u ruled a kingdom of some 10,000 km2 with a population of more than 170,000. In principle, Hawaiian kingdoms were expandable. They were capable of incorporating territory and population regardless of the -based allegiances that tended to constrain ancestral Polynesian chiefdoms. To survive, a state society must collect taxes, raise armies, fight wars, quell internal disputes, and construct civil works, all of which

4We may surmise that in time of war an ‘ili ‘aina¯ was also obliged to supply a similar number of recruits for the army. Watershed: Testing the Limited Land Hypothesis 31 require the coordinated work of large numbers of people. To accom- plish such tasks requires that accurate, detailed instructions reach the many participants and that reports of progress and problems reach the leader for expeditious response. Obviously, the ruler cannot communicate with all participants on a one-to-one basis. The problem is a common one for all large groups, and the solution too is virtually universal: the stratified control hierarchy. A stratified control hierarchy is a centrally controlled, task-oriented group of individuals5 organized in multiple tiers that generally in- crease in number of participants and diminish in task-responsibility directly with remoteness from the central task leader. Each major task is divided into sub-tasks, which are accomplished by individuals in the various strata, under the direction of supervisors at each level. This strategy often reduces the number of interactions required for a task to those among co-workers and direct supervisors. Often given short shrift in the literature on early states, in part because of dearth of evidence, are the strata of bureaucracy between, for example, the king who decrees the building of a temple or declares war and the people who carry the stones or fight on the front lines. Without the foremen and middle managers and without the sergeants and generals, there would be no monuments, no armies, and no states. Occupational specialization has been a hallmark of states since their beginning and I suggest that most professions evolved in the context of stratified control hierarchies. While archaeologists commonly refer to a state as a complex so- ciety, at its core is a structure—the stratified control hierarchy— that radically simplifies the management of each task (Hommon 1995, 1996, 1999, 2000, 2001, 2005). The invention of this structure has been fundamental to the emergence of every primary state. The broad spectrum of human behavior in social groups is more subject to bottom up or distributed control than to top down or central- ized control. Distributed control, which emerges from the constant interaction of the individual members in a social network, enables to maintain social cohesion, preserve culture, and, occasion- ally, facilitate the proliferation of innovations. Human social groups tend to be scale-free or small-world networks, in which relatively few participants maintain many links with others while most participants maintain relatively few links (Barabási and Bonabeau 2003).

5Such a system is in marked contrast to one composed of abstract institutions as, for example, in the Flannery (1972, 409–414) model. 32 Robert J. Hommon

Although human social networks successfully maintain traditional knowledge and behavioral rules, the decentralized, distributed con- trol upon which they are based is not well suited to making decisions about group tasks, even in a small group. Research has shown that as a leaderless group grows to six members its decision performance de- grades, a tendency that Johnson (1982, 392) ascribes to the increasing difficulty of dealing with the number of one-to-one interactions be- tween participants, a variable that grows non-linearly as participants are added. For example, the number of interactions increases from 6 to 15 to 28 in groups of 4, 6, and 8 participants, respectively, according to the formula (n2 − n)/2, where n = group size. The solution in a small group is to form a simple hierarchy by designating one member as the leader whose direction all participants agree to follow. This simple organization works efficiently for small, relatively simple tasks, but when a group grows to accomplish larger and more complex tasks a simple hierarchy’s efficiency is likely to degrade because of an over-extended span of control. The term span of control refers to “the number of individuals or organizational units directly subordinate to a given individual or organizational unit within a hierarchical structure” (Johnson 1982, 410). Johnson summarizes studies of more than 100 governmental agencies and business firms showing a mean span of control of about six, which seems to indicate that supervisors in organizations tend to adjust spans of control in accord with what can be called a Rule of Six (Johnson 1982, 410–412). Johnson, noting the range of span of control of one to as many as 24, suggests that the breadth of the span probably varies inversely with task scope and complexity (Johnson 1982, 412–413). It seems plausible that the simpler or more invariant the task the wider the span of control can be. Small-scale societies and communities routinely form work groups under the leadership of respected individuals who exhibit organiza- tional and practical skills. Some such groups are capable of expanding temporarily to form a hierarchical organization of two or more strata for the duration of an event such as a hunt requiring dozens or hun- dreds of participants. Such small, temporary hierarchical groups in some cases may have developed into the permanent, professional stratified control hierarchies at the core of nascent state societies. For the emerging ancient Hawaiian archaic states, a hierarchical template was readily available: the ancestral ranked kinship system that had probably arrived with the first settlers and was preserved by the chiefly class into the nineteenth century. Chiefly status in a Watershed: Testing the Limited Land Hypothesis 33 descending scale from n¯ı‘aupi‘o to kaukauali‘i depended on one’s parentage as well as genealogical distance from the reigning chief and his predecessors within 10 ascending generations (Malo 1951, 54– 56, 191–192). With the passage of generations the rank of a ruler’s descendants tended to be downgraded as their collateral distance from each successive ruler increased. For this reason, chiefs, includ- ing those as politically powerful as Kamehameha, frequently sought hypergamous marriages that would produce high ranked offspring (Hommon 1975, 2000; Kamakau 1964; Kirch and Sahlins 1992), In contrast, commoners kept no long genealogies and therefore could not aspire to high rank (Kamakau 1961, 1991; Kirch and Sahlins 1992). Though it was possible for the offspring of a chief and a commoner to be recognized as a chief of low rank (Malo 1951, 55), such rare ties of kinship had little effect on the sharp divide between the two generally endogamous classes. By the seventeenth century the high regard in which commoners might hold the chiefs apparently was not based on their being considered senior relatives in any but a general sense.

In historic practice, the ali‘i who are granted lands (moku or ahupua‘a) at the accession of a ruling chief, while usu- ally related in some way to him, would normally have no kinship with the people over whom they have been ‘seated’ (ho‘onoho). There is this important structural cleavage in the historic order . . . Explicitly likened by Hawaiians to the division of the spoils of victory, the customary redis- tribution of territories by the ruling chief at his accession establishes lords of the land (haku‘aina¯ ) who are effec- tively strangers to the people—kinsmen and henchmen of the chief above rather than of the cultivators below. If to the cultivators the ‘aina¯ , the ‘land,’ is the ‘plain of one’s bones’ (kula iwi), to the aristocracy it is a prize of war. (Kirch and Sahlins 1992, 1:26)

Virtually all governmental positions were occupied by members of the chiefly class. The political hierarchy was, however, functionally distinct from the system of chiefly ranks. The terms for ascribed chiefly ranks such as n¯ı‘aupi‘o, pi‘o, naha, wohi, papa, kaukau, la‘au, and maoli, pertained to parental ranks and degrees of sacredness (Kamakau 1964; Malo 1951). In contrast, the official titles of the political hierarchy reflected administrative functions of offices dis- 34 Robert J. Hommon tributed across a nested series of territorial units. Supreme chiefly rank, though highly respected, was not synonymous with rulership or degree of political success. In the environment of highly competitive states, accession to high office of chiefs like Kamehameha had less to do with ascribed status than with political and military skills. In order of size, the territorial units within a kingdom (an aupuni mo‘¯¯ı) included the mokupuni or moku (island), the moku or moku‘aina¯ (district), the ahupua‘a (local community or tax unit), and the ‘ili‘aina¯ (ahupua‘a subdivision) (Andrews 1865; Malo 1951), of which well over 1,000 examples can be identified by name and delineated boundaries. In addition, three terms refer to territorial subdivisions of a district containing multiple ahupua‘a: kalana, ‘okana, and poko (Lucas 1995; Malo 1951), about which extensive research in the literature so far yields relatively little information (D. Barrère, pers. comm.). Among the few names of kalana that have been preserved are Kona kai ‘opua¯ and Kekaha in the northern part of Hawai‘i Island’s Kona district (Maly 1998, 4). Kamoku in H¯am¯akua District appears to have been an ‘okana (Cordy 1994, 13, Table 3). Reasons for the paucity of references to kalana, ‘okana and poko in the nineteenth century records may be that by that time they were rendered unnecessary by depopulation and the abandonment of many communities as nucleated villages and towns sprang up around ports. The ancient Hawaiian stratified control hierarchy consisted of of- ficials who administered territorial units appropriate to their chiefly ranks (Malo 1951, 192). The offices frequently mentioned in the litera- ture include the ali‘i nui or mo‘¯ ¯ı (king); the ali‘i ‘ai moku (“chief that eats the district” or district chief), and the ali‘i ‘ai ahupua‘a (“chief that eats the ahupua‘a” or ahupua‘a chief) (Handy and Pukui 1958:5, 234, Sahlins 1992:18–19). Other than the mention by Andrews (1865) of “Alii okana” in the definition of ali‘i, there is little in the literature to indicate an administrative level between chiefs of moku and those of ahupua‘a, but the Rule of Six suggests that at least one such additional stratum existed in the bureaucracy of Hawai‘i Island if not the other kingdoms. Hawai‘i comprises about 1.7 times the land surface of all the other islands combined and each of its six districts was larger than whole islands. Ancient Kona district, for example, was larger than the entire island of Maui with its 12 districts. The average number of ahupua‘a per Hawai‘i Island district is about 70, three times the average of 23 on the other islands (Hommon 1976; Lyons 1875). I suggest that the breadth of a district chief’s span of control in a Hawai‘i Island district Watershed: Testing the Limited Land Hypothesis 35 probably was reduced to more manageable size by subdividing them into units like the kalana that have been identified in northern Kona district. To the extent that districts were subdivided, the stratified control hierarchy included at least four administrative levels: ruler, district chief, kalana chief, and ahupua‘a chief (Malo 1951, 142, 146). The governors that rulers sometimes installed over annexed territory added another tier between ruler and district chief (Hommon 1975, 96– 102). The variability of eighteenth century Hawaiian polities along a northwest-southeast geographical axis may shed light on the emer- gence of statehood. For example, according to traditional accounts, Kaua‘i, whose long unbroken line of rulers are not known to have an- nexed land by force (with the possible exception of Ni‘ihau), remained united and independent until the early ad 1700s. In contrast, Hawai‘i Island with its complex political history, punctuated with usurpations, rebellions, divisions, and unifications, was often ruled by chiefs whose success was more dependent on political and military prowess than ascribed status or succession by primogeniture. Douglas (1979) has proposed that leadership in indigenous Polynesian societies tended to vary along dimensions defined by rank and authority or power; ascription and achievement; sacred and secular; descent and local- ity; and ideology and practice. Though chiefly prerogatives along these continua shifted somewhat according to circumstance, those of Hawai‘i Island and Maui tended to be closer to the power-achievement- secular side and those of Kaua‘i to the rank-descent-ascriptive end (e.g. Kamakau 1961, 230). Kamakau’s story of Ma‘ilikukahi,¯ an O‘ahu chief of the early six- teenth century, tells of a time in O‘ahu’s past that honored rank, descent, and ascribed status as viewed from the nostalgic perspective of those living in a time when power, achievement and secularism are ascendant (Kamakau 1991, 56). Sahlins describes the contrast:

The striking difference between the traditional investi- ture of a legitimate king (a status Hawaiians reserved for Ma‘ilikukahi¯ and his kind, who were ali‘i of the ‘indigenous’ Nanaulu stock, born in the sacred temple of Kukaniloko¯ in upland Waialua) and the ritual metamorphosis of a wooden image into a true god (akua maoli) is that the latter alone is effected by human sacrifice. This is a point Hawaiians care to make: the kings of old did not practice human sacrifice. Warriors and usurpers, the more recent 36 Robert J. Hommon

kings characteristically achieve power by the sacrifice of rivals or rebels. (Kirch and Sahlins 1992, 1:21–22)

Sahlins remarks on the major transformation that led to what are here called the Hawaiian archaic states:

Everything looks as if Hawaiian society had been through a history in which the concepts of lineage—of a classic Polynesian sort, organizing the relations of persons and tenure of land by seniority of descent—had latterly been eroded by the development of chiefship . . . Intruding on the land and people from outside, like a foreign element, the chiefship usurps the collective rights of land control and in the process reduces the lineage order in scale, function, and coherence. Of course, no one knows when, how, or if such a thing ever happened. (Kirch and Sahlins 1992, 1:192)

The rest of this section attempts to shed light on Sahlins’ “when,” “how” and “if,” and to link the emergence of Hawaiian states to the Limited Land Hypothesis. I suggest that it is possible that the erosion of the classic lineage to which Sahlins refers took place in two major phases, the first of which can be traced to the Hawaii-Kahiki interac- tion of the Late Voyaging Period (ca. ad 1200–1400) as described in Appendix C. The second phase is discussed below. Professional stratified bureaucracies, the practice of conquest war- fare, and other features of the Hawaiian archaic states must have developed in situ considering the apparent absence of pre-existing exemplars in the Polynesian homeland. Archaeological and ethno- historical evidence suggests that the large, competitive kingdoms described here emerged late in ancient Hawaiian history. Given the literal and metaphorical significance of conquest and “spoils of war” in late pre-contact Hawai‘i (Kirch and Sahlins 1992, 1:26), one indication of the origin of states might be found in conditions that could be expected to lead to the intensity of competition and frequency of war known to have prevailed in the eighteenth century. Archaeological evidence pertaining to the five or six centuries immediately following the first successful Hawaiian settlement has been found in widely distributed locations on all islands. This pat- tern is consistent with the development of colonies in salubrious locations that eventually became the cores of 30 or so districts (Hom- mon 1986, 67). As population grew, settlement expanded inland and Watershed: Testing the Limited Land Hypothesis 37 laterally into previously unoccupied lands. It is plausible that, with continuing growth and expansion, each founding group began to fission into face-to-face neighborhoods that later developed into inde- pendent communities (Hommon 1986; Kirch and Green 2001; Lekson 1990). According to the revised Dye-Komori curve, population of the archipelago in ad 1200 was approximately 50,000 or 1/10th of the maximum (see Appendix A, pg. 53), and by ad 1450 it had grown to about 250,000. During much of the colonization process, the relatively low population size and density and the largely unhampered expansion process probably limited the incidence, scale, and intensity of competition or conflict. Though inter-community disputes no doubt arose from time to time, I suggest that few if any large scale wars of conquest between multi-district polities were waged before approximately ad 1450. Following this reasoning, the Hawaiian archaic states emerged sometime between about ad 1450 and 1778. If, as some archaeologists believe, traditional accounts of Hawaii’s past are not reliable, then we can say only that the archaic states described in eyewitness accounts at Western contact had been in existence for at least a generation or two, i.e., within the memory of people then living or, conservatively speaking, since about ad 1750. If, on the other hand, traditional histories can provide reasonably reliable coarse-grained outlines of events and cultural trends (though perhaps not such details as participants’ motives or moral character), then wars of conquest, and therefore the emergence of states can be traced back to the beginning of the seventeenth century. One reference to early inter-polity violence is provided by Fornan- der (1969, 2:67–69), who relates a traditional account of a military expedition led by Kalaunuiohua, a late fifteenth century ali‘i nui of Hawai‘i Island. After capturing the rulers of Maui and Moloka‘i and the chief of O‘ahu’s ‘Ewa and Wai‘anae districts, Kalaunuiohua landed on Kaua‘i, where he was defeated. The victorious Kaua‘i chiefs freed the captives. Then after some time had passed they allowed Kalaunuio- hua to return to Hawai‘i to continue his rule. This story does not depict the actions of a ruler of an aggressive Hawaiian state in that it lacks significant elements of later conquests, including sacrifice of defeated rulers, establishment of long-term control of captured land, and retaliation by the polities that had been attacked. The absence of these elements in this and other accounts indicates that as late as ad 1500 conquest states had not yet emerged in Hawai‘i. 38 Robert J. Hommon

Wars were waged by Hawaiian kingdoms of the late eighteenth century to annex and defend economically productive islands and districts. As discussed in the previous section, stories of such incor- poration by conquest, a definitive feature of archaic states, begin to appear in traditional histories dating to around ad 1600 (generation 41). Considering archaeological, ethnohistorical, and traditional sources then, we can say that the Hawaiian archaic states were emerging be- tween ad 1600 and 1750. This span corresponds closely with the middle part of expansion phase 2 during which, according to the Lim- ited Land Hypothesis, the rate of agricultural expansion had declined significantly. I have proposed that each of the variables discussed in the previ- ous sections supports the plausibility of the Limited Land Hypothesis by pertaining in some way to shortfalls of staple crop yields or sharply reduced access to reliably productive land during expansion phase 2. The boundaries of agricultural complexes that correspond to natural limits of aridity and fertility, the cultivation of high-risk marginal lands, the application of flood-water farming techniques, and the intensification of production in field systems all suggest that the avail- ability of new arable land had diminished significantly. The common knowledge of famine foods and of birth control methods and the ap- parent end of population growth evidenced by the Dye-Komori curve may have been consequences of staple food production shortfalls. These limits of land and food had to be faced most directly by the commoners, and it was they who had to devise their own solutions from below in each local community. In my view the emergence of Hawaiian archaic states between ad 1600 and 1750 also is consistent with the shortfalls of land and crops postulated by the Limited Land Hypothesis. In contrast to the bottom-up processes listed above however, state formation was primarily a top-down process that addressed problems faced by the elites. This process was based largely on the development of pro- fessional bureaucracies and conquest warfare, two innovations that served to alleviate the shortfall problems posed by markedly reduced rates of agricultural expansion. Though available evidence cannot unequivocally demonstrate a direct causal connection between slowed expansion and the rise of states, a link is strongly suggested by the contemporaneity of the two processes and the fact that political innovations of the latter process helped solve the problems posed by the former. In the previous Watershed: Testing the Limited Land Hypothesis 39 section we have seen that capturing and annexing new territories could serve to enhance government revenue. Whether or not conquest warfare was invented deliberately to solve the problem of a diminished tax base, for the victorious polities it appears to have served such a purpose admirably. At their inception the professional bureaucracies, in the form of stratified control hierarchies, also served revenue enhancement functions both internally and externally. In a Hawaiian state of the late eighteenth century, bureaucracies employing chiefs in at least three strata reached into each constituent ahupua‘a community, and even ‘ili, where they collected taxes, recruited armies, and organized corvée labor. The high standards of efficiency demanded of low level bureaucrats by chiefs at the district level served to maximize government revenues as indicated by the fact that a konohiki at the community level was “put off his land by the tax collector” if the taxes were not collected and placed at the ahupua‘a shrine in the quantity and according to the schedule demanded (Malo 1951, 145–146). The stress and uncertainty of agricultural shortfalls may or may not have initiated the development of conquest warfare and bureau- cracy. In either case, support provided by powerful chiefs who per- ceived benefits for themselves in these innovations probably triggered positive feedback processes that led to their subsequent elaboration. For example, if a fledgling bureaucracy successfully developed effi- cient ways to collect taxes and to command soldiers and laborers, the ruler would benefit from the reliable delivery of increasing quantities of goods and services. A part of such wealth could then be reinvested to improve the bureaucratic organization.

Part 3. Testing the Limited Land Hypothesis

The Limited Land Hypothesis proposes that the long-term rate of agricultural expansion in ancient Hawai‘i was reduced by at least half no later than ad 1550. As described in Part 1, testing the hypothesis requires analysis of archaeological data in a research area to arrive at reasonable estimates of five variables, namely: 1) the maximum area under cultivation; 2) the date of maximum expansion; 3) the area under cultivation at initial expansion; 4) the date of initial expansion; and 5) the area under cultivation in ad 1550. With these values, the LEX Index can be calculated by dividing the phase 2 (post-1550) expansion rate by the phase 1 (pre-1550) expansion rate. The threshold value 40 Robert J. Hommon of the LEX Index is 0.5. If the LEX Index of a given research area is 0.5 or less, the Limited Land Hypothesis is confirmed; if it is greater than 0.5, the hypothesis is disconfirmed. The strength of hypothesis confirmation or disconfirmation varies with the magnitude of the LEX Index, inversely for confirmation and directly for disconfirmation. The goal of a full test of the hypothesis is to arrive at a reasonable estimate of the LEX Index of a research area. However, if time and finances are insufficient for a full test, a preliminary or minimal test can be performed to determine whether the LEX Index is above or below the threshold value of 0.5. Once variables 1 through 4 listed above have been determined, the area under cultivation in 1550, variable 5, that would yield a LEX Index of 0.5 can be calculated. Variable 5 can be quite small; in the Kohala Field System model described in Part 1, minimal confirmation required that as little as 21.5 percent of the area was under cultivation before ad 1550. Whether conducting a minimal test or beginning a full test, it is important to identify zones of agricultural expansion in a research area. As described in Part 1, cultivation tended to begin in primary zones, where growing conditions were evidently perceived by Hawai- ian planters to be locally optimal, then to expand into secondary zones, and last, into final zones, which were considered marginal for various reasons. In some cases, final zones may be indistinguishable from secondary zones. As a rule, intensive testing of the hypothesis should begin with primary zones and, then, if necessary, move into secondary zones. Available evidence indicates that hypothesis-testing need not be as intensive in final zones since they tend to date to expansion phase 2. A typical research design for a research area would include the following procedures.

1. Field Work

a) Estimate the maximum area under cultivation, variable 1. In some cases, such as the Kohala Field System or some well-studied valley systems, fairly precise estimates of the maximum area are available. In areas that have not been as thoroughly researched, estimates may be based on recon- naissance surveys and the fact that agricultural complexes can be expected in the following environments: i. In alluvial deposits irrigable from permanent water sources with traditional Hawaiian technology; Watershed: Testing the Limited Land Hypothesis 41

ii. In nutrient-rich soils of geologically-young slopes of Hawai‘i and Maui that receive at least 20 in. of annual rainfall (Hartshorn et al. 2006; Kirch et al. 2004; Lade- foged and Graves 2000; Ladefoged et al. 1998, 2003; McCoy 2005; Mulrooney and Ladefoged 2005; Vitousek et al. 2004); and iii. In young soils on older islands where flood-water irri- gation was feasible. b) Estimate the date of maximum agricultural expansion, vari- able 2. Unless there is evidence of abandonment or with- drawal of land from cultivation, ad 1800 is usually a rea- sonable estimate. c) Estimate the area under cultivation at initial expansion, variable 3. Though evidence of the extent of cultivation near the beginning of a local sequence may be sparse, it is often possible to extrapolate from known data to estimate an area equal to one percent of the maximum cultivated area. d) Estimate the date of initial agricultural expansion, vari- able 4. If evidence in a research area is not yet available, comparisons with similar areas may be useful. As dis- cussed in Parts 1 and 2, research indicates that small-scale agricultural expansion in rain-fed areas was often under way by about ad 1300 and preliminary evidence suggests that valley irrigated systems were developing at least a century earlier. e) Calculate the area under cultivation in ad 1550, variable 5, that would confirm the Limited Land Hypothesis at the threshold level. This procedure is described in Part 1. f) Estimate the approximate location and extent of primary, secondary and final expansion zones based on analysis of environmental and cultural factors. This order of primary, secondary, and final reflects the relative importance of the zones and their priority during identification, testing and data analysis. For example, a preliminary confirmation of the hypothesis in the Kohala Field System might depend on little more than dating its small primary zones. Exam- ples discussed in Part 1 indicate that while sequences of expansion from primary to secondary to final zones can 42 Robert J. Hommon

be identified, expansion was neither necessarily linear nor unidirectional. i. Boundaries of complexes often coincide with natu- ral conditions such as sub-minimal rainfall or soils leached of nutrients that precluded indigenous cul- tivation. Examples of such a pattern have been doc- umented in the Kaho‘olawe inland agricultural zone (Hommon 1980, 1983), and in areas investigated by the Biocomplexity group (see 1a, pg. 40). Environmental factors useful in identifying primary zones include to- pography, rainfall patterns and edaphic variables, as discussed in Part 2 (pg. 13 ff.). A. Environmental factors useful in identifying pri- mary zones in irrigated systems include extensive land amenable to conversion to lo‘i and perma- nent, relatively-controllable sources of fresh water. Primary zones are likely to be near the mouths of irrigated valleys, where flood plains tend to be widest, but, as we have seen in the Wailau example (McElroy 2007, 238, Fig. 7.13), the earliest terraces may not necessarily be the closest to shore. As a general rule, final zones are expected to be located along environmentally determined boundaries, in- cluding margins of established field complexes, where the risk of flooding was greater, and in up- per valley reaches where sections of land suitable for the construction of lo‘i would be more limited. B. The ruling environmental factors in rain-fed areas are likely to be rainfall and soil fertility. For ex- ample, the apparent primary zones of the Kohala Field System are situated up-slope of the 20 in. isohyet on the west and down-slope of the eastern uplands where the soils are leached of nutrients (Hartshorn et al. 2006; Kirch et al. 2004; Ladefoged et al. 1998; Vitousek et al. 2004). C. Little is known about the expansion process of flood-water systems into young colluvial soils, but primary zones probably will be found along sec- tions of slope where surface runoff during rains was frequent yet subject to control with informal Watershed: Testing the Limited Land Hypothesis 43

terracing, diversion walls and other small-scale structures. ii. Cultural evidence that may be useful in identifying pri- mary zones include spatial relationships, forms, and stratigraphy of structural elements and the location of agricultural complexes within larger settlement pat- terns. A. Research has shown that in reticulated complexes such as the Kohala Field System and the Wailau Val- ley terraces, abutted and bonded corners and other horizontal and stratigraphic relationships can be used to assign field boundary walls, terraces, trails and other structural elements to relatively dated phases (Allen 2001c; Kirch 1985; Ladefoged and Graves 2000; Ladefoged et al. 2003; McElroy 2007; Mulrooney and Ladefoged 2005). B. In some cases, the form of a structure will indicate its relative age. For example, in one location in the Kona Field System, Allen (2001c, 137–142) has iden- tified a sequence of structural types, beginning with cross-slope terraces, followed by kuaiwi, and then stone mounds, that seems to mark phases of system development. A portion of a research area known to contain an early type of structural ele- ment would likely be a primary zone. Searching for early evidence beneath kuaiwi may be frustrated because they were sometimes built on ground that had been deliberately cleared of loose material (Major and Allen 2001, 97). C. Proximity to coastal resources and settlements was apparently a significant determinant in the selec- tion of primary zones. One of the reasons that the southern portion of the Kohala Field System was the last to be developed may have been that the cultivable section of the slope was farther from the shore than were sections to the north (Ladefoged and Graves 2000, 439). g) Conduct excavations to collect evidence specific to test- ing the hypothesis in primary zones, and, if necessary, in secondary and final zones. Testing the hypothesis in exten- 44 Robert J. Hommon

sive agricultural complexes by means of limited numbers of small excavation units requires that each sample col- lected reliably represent significant events in large-scale expansion processes. In rain-fed complexes these events likely included a sequence of initial clearing of natural vegetation by cutting and burning; clearing of stones; con- struction of initial boundary-delineating walls, retaining walls, mounds, and other structures; preparation of the ground with digging sticks and other ; and planting. In lo‘i complexes, the sequence also may have included cut-and-fill leveling and construction of canals and other water-control features. The most common and most effec- tive technique in a hypothesis-testing project is likely to be the 14C dating of vegetal material situated directly beneath stone structures such as boundary walls. Such samples would usually date a series of expansion events including vegetation clearance, wall construction and the initiation of cultivation. The cost of collection and analysis of chrono- metric samples will often limit the number of samples that can be processed, as compared with the spatial extent of the fields and other agricultural evidence that are to be dated. For this reason, samples should be collected from locations that are carefully selected to represent expansion on a large scale. For example, in a reticulated field system such as the Kohala Field System, a cluster of fields linked by bonded walls probably was built as a single unit. For the purposes of testing the hypothesis, as few as two 14C samples—perhaps one at each end of the cluster—might reliably date the whole cluster. Dating of stone structures is stressed here, but research should be designed to detect as well initial steps of expansion that may have involved swiddening with little or no stone construction.

i. Collect chronometric samples and record stratigraphic data including soil profiles and spatial relationships between stone retaining walls, field boundary walls, and other structures. A useful early step in any project testing the hypothesis is to clear, inspect and analyze available natural erosion faces. Because testing the hypothesis requires identifying and investigating the earliest cultivation evidence in a particular location, Watershed: Testing the Limited Land Hypothesis 45

the researcher should follow the recommendations of Dye (2001, 2004a, b) regarding surface structures. Care should be exercised in seeking out early structures and cultural deposits as distinguished from later, often su- perimposed, structures that are much more visible, plentiful, and intact. The stratigraphic evidence of an expansion event might be identified in soil stratigra- phy as for example, the interface between undisturbed soil indented by digging-stick marks and churned soil containing charcoal flecks. Even if the taxa of char- coal flecks cannot be determined, their appearance in an excavated column stratigraphically below other early agricultural evidence may indicate initial clear- ing. Excavations in the Kohala Field System exemplify the kind of field technique that would be useful in searching for evidence of an expansion event: [B]ackhoe Trench 50 was excavated to perpen- dicularly bisect a low, rocky field border wall near the seaward edge of the field system. Trenches throughout the field system often reveal intact stratification, with the wall’s rock and mounded earth fill overlying a clear soil se- quence of anthropogenically modified A-horizon with an underlying B-horizon. Often ash and charcoal lenses are visible at the interface be- tween the two horizons . . . Given the context, it is likely that this charcoal was burned at some point before, possibly, just before, the field wall was built. (Ladefoged et al. 2005, 365) Considering the importance of dating stone structures with vegetal samples found beneath their foundation stones, it may be necessary not simply to excavate cross-section trenches but to remove more extensive portions of certain structures to increase the probabil- ity of collecting high quality 14C samples. ii. The controlled collection of in situ botanical speci- mens, most of which will be carbonized, is essential to the research described here. The following kinds of identifiable botanical evidence can prove useful in testing the hypothesis. 46 Robert J. Hommon

A. Plant material of a wild vegetation community such as a climax forest can often be definitive evidence of an expansion event, especially if directly over- lain by evidence of agricultural activities. B. Samples of cultigens such as the sweet potato may evince expansion events if they are found in agri- cultural complexes (Clark and Kirch 1983; Coil and Kirch 2005; Ladefoged et al. 2005; Rosendahl and Yen 1971). C. Weedy taxa including Chenopodium, Amaranthus, Waltheria, Sicyos, and various grasses are known for their ability to colonize disturbed areas, such as fallow agricultural fields. Evidence of such coloniz- ing taxa can be crucial to identifying agricultural lands, especially where other types of evidence such as charcoal flecks or stone structures are lacking. An integral part of excavations designed for hypothesis testing will often be the collection of stratigraphic series of bulk soil samples for later water separation and sieving that may yield seeds and other plant parts, some of which are sub-millimeter in size. iii. Most chronometric analysis crucial to testing the hy- pothesis will be based on 14C dating of vegetal material. In some cases, such data will be supplemented by 14C dating of rat or other mammal bone or marine shell or 230Th coral dating (Kirch and Sharp 2005; Weisler et al. 2005). In the field, the aim is to collect samples that pertain specifically to the event of interest: the initial expansion of cultivation into a particular parcel. The success of any test of the hypothesis depends on con- trolling two factors during collection and preparation of chronometric samples (Dye 2000). A. The first factor is provenance. It is important that there be spatial, stratigraphic, botanical, or other evidence that the sample was deposited as part of an expansion event as noted in 1g (pg. 43). If material for the direct dating of an expansion event is not available, carefully selected samples from adjacent deposits may aid in bracketing the event. Watershed: Testing the Limited Land Hypothesis 47

B. The second factor is composition. Samples for 14C dating should be limited to fragments from short-lived taxa, or small-scale botanical structures such as twigs, seeds, or nutshells to avoid the risk of dating a sample with in-built age, such as the heartwood of long-lived trees, waterlogged wood, or driftwood (Dye 2000). The sample selected for dating should be a single piece, to avoid possible mixture of twigs, seeds, or nutshells of different ages.

iv. Though chronometric data used to test the hypoth- esis should usually be limited to direct evidence of expansion events, analysis of indirect evidence may also be of interest. Kaho‘olawe is a case in point. The inland zone at the summit of Kaho‘olawe contains a cluster of about 1,000 camp-sites and other features (Hommon 1980). Direct evidence of agricultural activ- ity such as stone mounds or retaining walls is lacking in this zone, probably because of the gentleness of the slope and the paucity of naturally-occurring stones. However, the fact that the limit of the inland zone is approximately coterminous with the 20 in. isohyet indicates that most of the sites served as camps for cultivators of dry-land staple sweet potato throughout much of the zone (Giambelluca and Schroeder 1998; Hommon 1980, 1983). Analysis of the 14C data from campsite and imu reported by Hommon (1983) and Rosendahl et al. (1987) might indicate whether the hypothesis would be confirmed on Kaho‘olawe. v. As with any subsurface archaeological research, test- ing the hypothesis will usually require simultaneous consideration of stratigraphic, botanical, and chrono- metric data, and all should be appropriate to the task. For example, consider the case of an irrigation ditch feeding a lo‘i system. Suppose a sample collected from the ditch were of a botanically-identified short-lived species ideally suited for chronometric analysis. Such a sample would not serve to test the hypothesis be- cause it would likely date a later maintenance event rather than original construction of the ditch. On the 48 Robert J. Hommon

other hand, 14C samples collected from the soil sur- face into which the ditch was excavated might supply a secure date for the establishment of the lo‘i.

2. Post-Field Work

a) Conduct archaeobotanical analyses of samples to address paleo-environmental and chronometric questions discussed in 1g (pg. 43). Detailed archaeobotanical analysis, rec- ognized as a vital element of archaeological research in Hawai‘i for more than a quarter century (Murakami 1983a, b), is of particular importance to testing the Limited Land Hy- pothesis.

i. Identifying the taxon and plant part of a sample to be submitted for 14C analysis has become a standard operating procedure in Hawai‘i in part because such analysis can help identify the natural and cultural con- text of the dated sample (see 1(g)iii, pg. 46). In a project designed to test the hypothesis, botanical analyses may provide a bonus as well. If the sample is of a taxon from the wild plant community that was removed for initial cultivation, it can be a strong indicator of an expansion event. ii. Paleo-environmental analysis should include separa- tion of botanical material from bulk soil samples by flotation. Among the taxa of particular interest to such a study are weedy taxa including the genera Chenopodium, Amaranthus, Waltheria, and Sicyos, all of which are known to colonize disturbed areas as noted in 1(g)iiC (pg. 46). Hawaiian species of Cheno- podium have been found in archaeobotanical assem- blages on Hawai‘i, Kaho‘olawe, Moloka‘i, and O‘ahu (Allen 1987b, A-13). In analyses of botanical samples collected during two phases of research on Kaho‘olawe, Allen (1983, 1987b) found that Chenopodium was by far the most frequent and most abundant taxon in analyzed soil samples from 24 of 26 localities, most of which were temporary habitation sites in the inland zone, believed to be the main agricultural area on the island (see 1(g)iv, pg. 47). Melinda Allen noted Watershed: Testing the Limited Land Hypothesis 49

a marked correlation between high Chenopodium concentrations and the proximity of the sam- ple localities to the inland plateau, particularly the portion of the plateau where archaeological sites are concentrated. (Allen 1987b, A-26) The fact that abundance of Chenopodium evidence, even tiny sub-millimeter seeds, correlated so well with the main agricultural zone of the island indicates the value of Chenopodium in testing the Limited Land Hypothesis. It is possible for disturbance caused by natural events such as lightning-ignited fires or land- slides to encourage Chenopodium growth. However, when found stratigraphically between a stratum with no evidence of agriculture and a stratum with later agricultural structures such as mounds or kuaiwi, a Chenopodium-rich layer likely evidences an early stage of agricultural clearing or fallowing. Chenopodium offers an added bonus, as well—it is known to have grown large enough to be collected as firewood and to produce charcoal in identifiable and datable chunks (Hommon 1983, 123, Table 19). b) Process, select, and submit 14C samples for analysis. i. Processing should include cataloging, cleaning, and taxonomic identification. ii. Selection of samples for analysis should be based on provenance, weight, and location. iii. Laboratory analysis should include 13C/12C fractiona- tion analysis, and accelerator mass-spectrometry (AMS) or other procedures to ensure high precision. iv. Calibrate the results using either the OxCal program (Bronk Ramsey 2001, 2007; Ramsey 1995) or other soft- ware capable of Bayesian analysis, if required. c) Report the results of research conducted to test the Limited Land Hypothesis. The following items pertaining specifi- cally to the testing of the Limited Land Hypothesis should be added to the standard elements that are expected in any publication of archaeological results in Hawai‘i. i. The introduction should include a summary of the results of the test. 50 Robert J. Hommon

ii. The project description should include an introduc- tion to the Limited Land Hypothesis and a discussion of testing terms and procedures including expansion phases, expansion zones, the five spatial and temporal variables, and the LEX Index. iii. The description of the project area should include the estimated values of variables 1 through 4, the expected value of variable 5 required for a threshold LEX Index of 0.5, and an account of the primary, secondary and final expansion zones and how they were identified. iv. The description of each excavation unit should include its expected contribution to testing the hypothesis and whether expectations were met. v. Each chronometric sample should be described in terms of provenance, composition, 14C age, fraction- ation data, and calibration results, along with infer- ences that are drawn from these results regarding the hypothesis. vi. The discussion of project results should include the LEX Index of the project area, and an assessment of its reliability. The level of reliability may vary, in the judgment of the researcher, according to factors such as how thoroughly the various zones and the complex as a whole has been surveyed and dated. vii. The conclusion of the report should summarize the results and estimate their local, regional, and island significance.

Summary and Conclusion

This paper recommends that archaeologists conduct research to test the Limited Land Hypothesis, which states:

By ad 1550 ancient Hawaiian agricultural expansion slowed significantly as it approached effective limits on staple pro- duction imposed by available technology, sociopolitical factors, and natural variables including rainfall and soil fertility.

Part 1 of the paper defines the spatial and temporal elements of the hypothesis and outlines procedures to test it. In any research Watershed: Testing the Limited Land Hypothesis 51 area where the hypothesis is to be tested expansion phase 1 begins with the initiation of agricultural expansion and ends no later than ad 1550. Expansion phase 2 begins no later than ad 1550 and ends with maximum expansion. If the rate of expansion during phase 2 is half that of phase 1 it is assigned the threshold LEX Index of 0.5 and the hypothesis is confirmed. The lower the LEX Index the stronger the confirmation of the hypothesis. A LEX Index of zero would signify that expansion ceased by ad 1550. If the LEX Index is greater than 0.5 the hypothesis is disconfirmed, and a LEX of 1.0 demonstrates that the expansion rate remained unchanged through both phases. In research areas such as the Kohala Field System and the Wailau Valley lo‘i system archaeologically identifiable primary, secondary and final expansion zones evidence a sequence of primary, secondary, and final stages of expansion. In these research areas initial development of primary zones apparently took place before ad 1550 and develop- ment in the final zones probably began after ad 1550. Determining the age of secondary zones will likely be crucial to establishing whether the hypothesis is confirmed in similar medium- to large-scale systems. Part 2 establishes the plausibility of the hypothesis by drawing on research in soils , demography, history, , and ar- chaeology indicating the following processes during expansion phase 2:

• Aridity, low soil fertility, and terrain routinely imposed bound- aries on agricultural complexes;

• Hawaiian planters cultivated extensive marginal lands using innovative techniques such as flood-water farming, indicating that more productive lands were no longer available for devel- opment;

• Hawaiians maintained traditional knowledge of an array of famine foods, a fact that suggests relatively frequent food short- ages;

• Activity in the Mauna Kea adze quarry apparently declined, possibly signaling reduced demand as the pace of clearing new lands for cultivation slowed;

• Reduced opportunity for expansion into new productive lands appears to have led to a decline in the archipelagic population 52 Robert J. Hommon

growth rate, probably effected largely by commoners’ applica- tion of traditionally-maintained knowledge of population control methods;

• Chiefs developed conquest warfare as a means by which, in an era of reduced expansion, they augmented their tax base by capturing and annexing communities complete with their resident producers; and

• Archaic states emerged as ruling chiefs, whose authority de- pended increasingly on success in war and politics and less on ascribed status, governed by means of stratified control hi- erarchies of professional bureaucrats that maintained order, collected taxes, oversaw public works and raised armies.

Part 3 offers a model research design the application of which can facilitate archaeological tests of the hypothesis. Whether the test is the sole purpose of the project or an element of a more broad-based effort, the research should, at a minimum, delineate primary zones and establish whether a sufficient portion of the research area was under cultivation before ad 1550 to yield a LEX Index at the threshold value of 0.5. The Limited land Hypothesis describes a process with ramifica- tions affecting virtually every aspect of ancient Hawaiian culture. Though formal tests of the hypothesis will sharpen our understand- ing of this watershed process, likely effects can be modeled as follows:

• Agricultural expansion, which had proceeded exponentially for centuries, encountered various natural and technological limits abruptly, within the span of an average lifetime, in regions such as the west coast of Hawai‘i Island;

• Commoners experienced the slowed pace of expansion in a time of continued population growth in the form of shortages of food and sharply reduced opportunities to expand into reliably productive lands;

• Commoners responded by accelerating intensification in existing fields, by applying new methods such as flood-water farming to bring marginal lands under production, by limiting family size with traditional methods of population control, and by seeking the blessings of the ‘aumakua; Watershed: Testing the Limited Land Hypothesis 53

• Chiefs, who experienced shortages and near zero population growth as a diminished flow of taxes and gifts, responded by seeking the help of the gods in major temples, by coercion, and by organizing large-scale efforts to intensify production with terraces, wind walls and other public works;

• With the growing importance of administrative and military skills relative to ascribed status, chiefs retooled the ancient arts of war to capture and annex large tracts of land together with resident producers so as to expand the tax base;

• Tasks including efficient administration of captured territories, tax collection, and recruiting armies and corvée laborers were facilitated by retooling the ancient ranked structure of the Poly- nesian lineage to create professional bureaucracies, the organiza- tional framework around which the emergent Hawaiian archaic states were constructed.

Appendix A. New Estimates of the Ancient Hawaiian Maximum Population

Archaeologists in Hawai‘i, following Schmitt (1968, 1971, 1973), have often adopted a maximum ancient Hawaiian population estimate of between 200,000 and 300,000 (e.g. Dye 1994; Hommon 1976; Kirch 2000), though they often mention in passing Stannard’s (1989) argu- ment for a population of 800,000 to 1,000,000. I propose here that early Hawaiian census data and comparisons with census figures for central East Polynesia support a maximum estimate for Hawai‘i in ad 1550 of approximately 500,000. To arrive at this estimate I have assigned a population value to the late end of a revised version of the Dye-Komori 14C-based population curve (see Part 1; Dye and Komori 1992b; Dye 1994) by extrapolating back to ad 1778 the rate of depopulation evident in early nineteenth century Hawaiian censuses. The maximum population in ad 1550 is then calculated in accordance with the Dye-Komori curve, which indicates a total decline in Hawaii’s population of about ten percent between ad 1550 and 1778. The following is a discussion of the method I used to arrive at four alternative estimates of maximum ancient Hawaiian population at ad 1550, including the one selected for use in this paper. The 54 Robert J. Hommon term depopulation rate refers to a continuously compounded rate of negative population growth, as defined in equation (1.1). It is generally recognized that the decline in Native Hawaiian pop- ulation following Western contact can be attributed largely to the direct and indirect effects of introduced diseases to which the Hawai- ian population, whose ancestors had been isolated from continental populations for millennia, had little natural resistance (Bushnell 1993; Schmitt 1968, 1970b). Table 1.2 summarizes the archipelago-wide data from Hawaii’s first eight censuses as well as the calculated average annual depopulation rates for each inter-censal period.

Table 1.2 Hawaiian Islands depopulation rates

Inter-Censal Growth Census Year Population Years Rate (%) 1831–1832 128,713∗ 1835–1836 108,568† 4 -4.26 1850 84,165† 14 -1.82 1853 71,019‡ 3 -5.66 1860 67,084‡ 7 -0.81 1866 58,765‡ 6 -2.21 1872 51,531‡ 6 -2.19 1878 47,508‡ 6 -1.35

∗Adjusted figure according to Schmitt (1973, 7–8). †Schmitt (1973, 39). ‡Hawaiians and Part-Hawaiians (Schmitt 1968, 74, Table 16).

Three observations support the suggestion that the rate of Hawai- ian depopulation during the pre-censal period was at least as high as that documented in the early nineteenth century census data. First, the initial decades following the first contact with the non- Polynesian world offered many opportunities for the introduction of continental pathogens to which the Hawaiians were vulnerable. At least 192 ship-visits (including multiple visits by some ships) are documented during the 54-year span (1778–1832) from first contact to the first Hawai‘i census (Judd 1974, 1–24). During such sojourns, which were often several weeks or months in duration, virtually every member of most ship’s companies interacted with Hawaiian residents. A conservative estimate of 100 close encounters (sexual and other- wise) between individual crew members and Hawaiians during the average ship’s visit yields more than 19,000 opportunities for disease contagion. This estimate includes over 6,000 such contacts in the Watershed: Testing the Limited Land Hypothesis 55

first 20 post-contact years, when the Hawaiian population was most vulnerable to newly-introduced diseases. Second, though the early post-contact history of disease in Hawai‘i is poorly documented, the few brief accounts of epidemics, hinting at high frequency, wide distribution, and marked lethality, lend support to the notion of relatively high pre-censal depopulation rates. Schmitt (1970b) presents what little is known about four major epidemics that “produced a great mortality” or “killed thousands” in the scant 22 years between 1804 and 1826. The greatest of these was the oku‘u of 1804 which killed between 5,000 and 15,000 people (some 2 percent to 6 percent of the 1804 Hawaiian population as estimated in this paper). The third indication of high pre-censal rates of depopulation is to be found in census data and early estimates from the Society, Marquesas and Hawaiian Islands. Though depopulation rates based on data sets tend to vary greatly in the short term, as shown in column 4 of Table 1.2, decadal-scale comparisons show a marked tendency to diminish with time (see column 4 in Tables 1.3 and 1.4).

Table 1.3 Tahiti: Estimated depopulation rates, 1769–1863

Estimate Estimated Years Between Growth Year Population Estimates Rate (%) 1769 35,000 1797 16,050 28 -2.78 1829 8,674 32 -1.92 1863 7,642 34 -0.37

Sources: McArthur (1968, 240, 260, 261, Table 53), Oliver (1974, 1:26–34).

Table 1.4 Marquesas: Estimated depopulation rates, 1840–1926

Estimate Estimated Years Between Growth Year Population Estimates Rate (%) 1840 19,300 1867 7,411 27 -3.55 1889 4,472 22 -2.30 1926 2,260 37 -1.85

Source: McArthur (1968, 286, Table 54, 315, Table 55).

The documented tendency for depopulation rates to decline with time is consistent with a rate in Hawaii’s pre-censal decades that was 56 Robert J. Hommon

Table 1.5 Hawaiian Islands: Ancient population estimates

Population Estimate Alternative Interval Years Growth Rate (%) ad 1778 ad 1550 High 1832–1853 21 -2.83 593,350 659,278 1853–1872 19 -1.69 1872–1896 24 -1.11 Medium∗ 1832–1860 28 -2.33 452,952 503,280 1860–1872 12 -2.20 1872–1896 24 -1.11 Low 1832–1872 40 -2.29 443,273 492,526 1872–1896 24 -1.11 Lowest 1850–1872 22 -2.23 419,213 465,792 1872–1896 24 -1.11

∗Preferred alternative.

greater than that documented by the early censuses. On the other hand, such factors as the absence of additional contacts for six years immediately following the departure of the Cook expedition might argue for a somewhat lower long term rate than that derived from the census data. Taking the middle ground, I propose that the size of the ad 1778 population can be approximated by applying an early census-based depopulation rate to the pre-censal period. The question then is: which early census-based rate? Calculating an average depopulation rate depends on which census figures are selected and the length of the interval between them. Table 1.5 presents four alternatives (column 1). Pairs of census figures from the designated time-spans (column 2) of varying lengths (column 3) have been used to calculate sets of inter-censal depopulation rates (column 4), the earliest of which, projected back to ad 1778, yields in each set a population estimate for that year (column 5). Each ad 1778 estimate is then used to calculate the ad 1550 maximum population figure based on the ten percent decline evident in the Dye-Komori curve between ad 1550 and 1778. Alternatives 1 through 3 in Table 1.5 begin with the first “mission- ary” census in 1831–1832. In his detailed treatment of the missionary censuses, Schmitt (1973, 6–7) notes apparent net over-counting in the 1831–1832 census and net under-counting in 1835–1836. For the reported 1831–1832 total population of 130,313, Schmitt (1973, 7) sug- gests a more accurate figure of 128,713 and this is the figure used here. Watershed: Testing the Limited Land Hypothesis 57

The relatively high early depopulation rate in Alternative 1 may be attributable to a time span of 21 years that is somewhat shorter than those of other alternatives; however it should be noted that this rate is well within the range of the early rates estimated for Tahiti and the Marquesas. Alternative 4 may avoid problems of the 1830’s under- and over-counts by starting with the 1850 census, but it increases the risk of underestimating the pre-censal figures by applying to the pre-censal period a depopulation rate that is unrealistically low given the tendency for depopulation rate to decrease over the ensuing 18 years.

The four alternatives yield population estimates for 1778 ranging from 419,213 to 593,350. Considering all factors discussed here, the al- ternative 2 figure of 452,952, rounded to 450,000 serves the purposes of the present paper.

The impact of decades of depopulation on the Hawaiian people at the -2.33 percent annual rate estimated here undoubtedly was devastating. In comparison, the U.S. death toll attributed to the Spanish Flu of 1918–1919, recognized as the most lethal pandemic in American history, represented 0.53 percent to 0.65 percent of the population, and it lasted less than a single year (American Red Cross 2007; Miami University 2007).

One final comment is appropriate. As noted above, population estimates in the range of 200,000 to 300,000 have often been accepted by archaeologists and others as reasonable for the Hawaiian Islands at the time of Western contact. Applying the analysis documented in this appendix shows that, depending on whether the 1832 or 1850 census figure is used, the rate of pre-censal depopulation ranged from -0.82 percent to -1.20 percent for a 1778 population of 200,000 and -1.57 percent to -1.77 percent for 300,000. All these figures are lower than any Hawaiian depopulation rate between 1832 and 1872, with the exception of the -0.81 rate during the brief period between 1853 and 1860. I would argue that it is exceedingly unlikely that the depopulation rate of a population so vulnerable to the introduction of continental diseases would increase after five to seven decades of contact. For this reason a total Hawaiian population as low as 200,000 to 300,000 cannot be supported by available evidence. 58 Robert J. Hommon

Appendix B. A Population Growth Rule of Thumb

The following details how the application of the population growth rule of thumb noted in Part 2 provides provisional support to the Limited Land Hypothesis in five research areas. As in the body of the paper, ad 1550 is the terminus ante quem separating expansion phase 1 from expansion phase 2. The population growth rule of thumb can be stated as follows:

To arrive at a preliminary assessment of whether the Lim- ited Land Hypothesis will be confirmed in a research area, calculate the ratio of phase 1 to phase 2 person-years based on the ratio of phase 1 to phase 2 14C ranges and estimates of initial and maximum population of the area. If the person-year ratio indicates near-zero population growth during phase 2, a formal test of agricultural expan- sion in the area will probably confirm the hypothesis.

As discussed below, application of the rule of thumb indicates that phase 2 population growth in the five research areas on four islands probably approximated zero, suggesting that formal tests would confirm the Limited Land Hypothesis in those areas. The rule of thumb provides a cautionary note in that it shows that an increase of evidence of human activity after ad 1550 by a factor of up to 4.6 does not necessarily demonstrate phase 2 population growth and may in fact be consistent with zero or negative growth. The procedural steps for applying the population growth rule of thumb are as follows: 1. Calculate the ratio of phase 1 calibrated 14C date ranges to those of phase 2. For example, if 15 of the date ranges are pre-1550 and 45 are post-1550, then the ratio is 1:3.

2. Estimate the maximum population of the research area in ad 1550 and assume zero growth until ad 1800.

3. Estimate the size of the initial population of the research area in a specific year. In rain-fed areas an initial population in ad 1300 equal to one to five percent of the maximum population of the area may be appropriate.

4. Applying the initial and maximum population estimates, cal- culate the numbers of person-years in phase 1 and in phase 2, using equation (1.2) on page 60. Watershed: Testing the Limited Land Hypothesis 59

5. Summarize the figures from step 4 as a ratio of phase 1 person- years to phase 2 person-years.

6. Compare the ratio of 14C calibrated date ranges to the person- year ratio. A second term of the 14C-derived ratio that is less than or equal to the second term of the person-year ratio indi- cates a near-zero phase 2 population growth rate, a result that is consistent with confirmation of the Limited Land Hypothesis.

Table 1.6 summarizes 254 14C date ranges from five research areas on four islands. Of these date ranges, 134 are fully within either phase 1 or phase 2. The remaining 120 ranges overlap portions of both phases. Half of these ranges have been assigned arbitrarily to phase 1 and half to phase 2. The resulting ratios of phase 1 to phase 2 date ranges are found to vary from 1:1.3 for the ‘Ewa Plain, O‘ahu to 1:3.4 for Kahikinui, Maui. The mean ratio for all 254 ranges is 1:1.8. The figures in column 4 of Table 1.6 represent the maximum number of phase 1 14C ranges that might represent human activity in ad 1300, the beginning of phase 1.

Table 1.6 Ratio of Phase 1 and 2 14C dates∗

Calibrated 14C Range

Island Location P1† 1300‡ P2† P1:P2§ Zero Growth

Hawai‘i¶ Kealakekua 12 3 24 1:2.0 1:1.9 to 1:4.6 Hawai‘ik Waimea 5 0 9 1:1.8 1:3.2 to 1:4.6 Maui∗∗ Kahikinui 10.5 1 35.5 1:3.4 1:2.6 to 1:4.6 Moloka‘i†† Kalaupapa 4 0 12 1:3.0 1:3.2 to 1:4.6 O‘ahu‡‡ ‘Ewa Plain 61 21 81 1:1.3 1:1.6 to 1.4.6 Totals 92.5 25 161.5 1:1.8 1:1.8 to 1:4.6

∗ Compared to hypothetical person-year ratios with Phase 2 zero growth. † P1 = phase 1,P2 = phase 2. ‡ Phase 1 14C ranges that might represent human activity in ad 1300. § Person-year ratio range. ¶ Source: Tomonari-Tuggle (2006, 12–14, Table 2, 124, Table 52). k Source: Burtchard and Tomonari-Tuggle (2004, 60, Fig. 5). ∗∗ Source: Coil and Kirch (2005, 79, Fig. 6), Dixon et al. (2002, 16). †† Source: McCoy (2005, 345, Fig. 3). ‡‡ Source: Tuggle (1997, 24–29), Appendix B. Note: Date ranges at 2σ except Tuggle (1997, 29).

Person-years, defined as the sum of living persons in every year during a specified time period, provide an approximate measure of 60 Robert J. Hommon human activity during that period. Given population figures in the first and last years of a span of time, the number of person-years, p, in that span can be calculated with the following equation:

(n − n )(t − t ) p = 2 1 2 1 (1.2) ln n2 − ln n1 where n1 = population at time t1, n2 = population at time t2 (with t2 > t1), and ln is the natural logarithm (Keyfitz 1966, 581). Let us suppose, consistent with the Dye-Komori assumption (Dye and Komori 1992b), that the 14C frequency in a given research area was isomorphic with population size. The greater the number of person-years during a specified period, the greater the number of datable carbon samples we are likely to recover. In the absence of population estimates, equation (1.2) cannot be applied directly to the research areas listed in Table 2. However, it is possible to interpret the 14C-derived phase 1-to-phase 2 ratios by comparing them with population growth scenarios based on the application of equation (1.2) to hypothetical population figures. Each of the scenarios in column 7 of Table 1.6 represents phase 1 growth from a small population in ad 1300 to a hypothetical maxi- mum population of 1,000 in ad 1550, followed by zero growth between 1550 and 1800. Consistent with evidence of agricultural expansion discussed in Part 1, major population growth in these research areas, which were dependent on rain-fed cultivation, appears to have been mainly a fifteenth century phenomenon. Estimating the initial popu- lation size in ad 1300 in these hypothetical scenarios is challenging for several reasons.6 Nevertheless, it seems reasonable to propose that for those research areas with no ad 1300 date ranges the initial population can be estimated at somewhere between one percent and five percent of the maximum population. Solving equation (1.2) with these figures we find that the ratio of phase 1 person-years to phase 2 person-years is between 1:3.2 (with an initial population of 50) and 1:4.6 (with an initial population of 10).7 For the Kealakekua, Kahikinui, and ‘Ewa Plain research areas, whose 14C ranges include ad 1300, the maximum initial population

6In addition to the possibility of sampling error, most of the early calibrated ranges are long enough to represent events well after ad 1300, and the ranges of as many as half the samples in Table 1.6 may have in-built ages of unknown magnitude. 7 The figures entered in equation (1.2) are as follows: Model 1: early period, n1 = 10, n2 = 1000, t1 = 1300, t2 = 1550; late period, 1,000 persons × 250 years = 250,000 person-years. Model 2: early period, n1 = 50, n2 = 1000, t1 = 1300, t2 = 1550; late period, 1,000 persons × 250 years = 250,000 person years. Watershed: Testing the Limited Land Hypothesis 61 estimates are, respectively, 250, 95, and 340, values based on the assumption that every 14C range that includes ad 1300 evinces initial human activity at that time. In conclusion, the 14C ratios of all five research areas appear to exhibit near-zero population growth rates during phase 2. The 14C phase 1-to-phase 2 ratios for Waimea and the ‘Ewa Plain are lower than the hypothetical zero-growth figures, which may suggest negative population growth during phase 2. The ratios from the other research areas may be indicative of zero population growth after ad 1550, though assuming maximum founding populations in ad 1300 Kahikinui and Kealakekua may have continued to grow at low rates. Assuming that population growth was proportional to agricultural expansion, these results provide indirect support for the Limited Land Hypothesis.

Appendix C. “An Era of National Unrest and of Tribal Commotion”: New Light on the Kahiki Connection, AD 1200–1400

About the commencement of the eleventh century, after a period of comparative quiet and obscurity, the Polynesian folklore in all the principal groups becomes replete with the legends and songs of a number of remarkable men, of bold expeditions, stirring adventures, and voyages undertaken to far-off lands. An era of national unrest and of tribal commotion seems to have set in, from causes not now known, nor mentioned in the legends . . . Be the cause what it may, a migratory wave swept the island world of the Pacific, embracing in its vortex all the principal groups, and probably all the smaller. Chiefs from the southern groups visited the Hawaiian group, and chiefs from the latter visited the former, accompanied by their relatives, priests, and retainers, and left indelible traces of their séjour and permanent settlement on the genealogies of succeeding chiefs, in the disuse of old and substitution of new names for places and landmarks, in the displacement of old and setting up of new tutelar gods, with enlarged rites of worship and stricter tabus.

Fornander (1969, 6–7)

Hawaiian traditions tell of long-distance voyaging by both Hawaiian and foreign chiefs between Hawai‘i and East Polynesian archipelagos collectively known as Kahiki. Depending on context the Hawaiian term “Kahiki” might range from simply any foreign land (Cachola- Abad 1993; Ellis 1963) to “invisible and celestial realms beyond the horizon, the legendary source of the great gods, ancient kings and cultural good things” (Sahlins 1985, 74). 62 Robert J. Hommon

In Sahlins’ encapsulation of the evident transformation of ancient Hawaiian society, he refers to the erosion of the classic Polynesian lineage by the development of chiefship.

Intruding on the land and people from the outside, like a foreign element, the chiefship usurps the collective rights of land control and in the process reduces the lineage order in scale, function, and coherence. (Kirch and Sahlins 1992, 1:192)

In light of recent research in a variety of fields archaeologists should again consider the possibility that at least some of the traditions of Kahiki document long-distance voyaging that actually did intrude on Hawai‘i and that the resulting interactions with foreign chiefs may have played a significant role in the transformation of Hawaiian soci- ety. The fact that stories of exploits of and genealogical connections with Kahiki chiefs and their descendants were preserved (and probably elaborated) as chiefly lore for some 20 generations after the apparent end of long-distance voyaging testifies to the high value placed on these connections. Plausibly, it was the tendency of Hawaiian chiefs to identify more strongly with the high status voyagers from Kahiki than with Hawaiian commoners that so weakened the traditional lineage bonds that native chiefs came to seem a “foreign element” in their own land. The following is an outline of converging archaeological, archaeo- botanical, linguistic, traditional, genealogical, and climatological evi- dence, much of it recently published, that lends support to the idea that interaction between Hawai‘i and Kahiki increased in intensity during what can be called the Late Voyaging Period, spanning ap- proximately the thirteenth and fourteenth centuries, midway through ancient Hawaiian history.8

Traditional Accounts of Kahiki Chiefs, ca. AD 1200–1400

Fornander (1969, 2:6–7, 34) describes the Migratory Period as begin- ning several centuries after Hawaii‘s original settlement and spanning roughly 10 chiefly generations. Traditions of this period recount a revival or increase of long-distance voyaging between Hawai‘i and East

8The genealogical chronology of ancient Hawai‘i that parallels the ca. ad 800– 1800 chronology is based on traditional chiefly generations averaging 20 years and numbered 1 through 50, as noted in Part 1 (Hommon 1976; Stokes 1933). Watershed: Testing the Limited Land Hypothesis 63

Polynesian archipelagos. According to a compilation by Cachola-Abad (1993, 24–25, Table 3.3) the Hawaiian traditions refer to 48 named Hawaiian and foreign voyagers participating in 24 voyages, provided we exclude gods, demigods, nineteenth century fictional characters (cf. Barrère 1969, 37–38) and voyagers unidentified as to generation. Archaeologists have been debating the existence and significance of Fornander’s Migratory Period for decades (e.g. Cachola-Abad 1993; Cordy 1974, 2000; Emory 1968; Green 1974; Hommon 1976, 1989; Sinoto 1968). Today most researchers, largely in the apparent absence of convincing artifactual evidence—with the possible exception of feather girdles of investiture (Rose 1978)—tend to ignore or reject the idea of a late period of intensified Hawaii-Kahiki interaction (Cachola- Abad 1993; Cordy 2000; Goto 1986). Cachola-Abad (1993) identifies two models of East Polynesian set- tlement. According to what she terms the orthodox model of dual settlement, Hawai‘i was settled by two waves of voyagers, the first from the Marquesas Islands and the second, corresponding to For- nander’s Migratory Period, from Tahiti. She advocates replacing this model with an interaction model that

proposes that the Hawaiian populace was comprised of individuals from the wide geographic region of central East Polynesia. Recurrent arrivals are seen as diminishing through time. (Cachola-Abad 1993, 14)

Analyses of mtDNA showing that Polynesian rats (Rattus exulans) were introduced to Hawai‘i from the Marquesas, Society, and South- ern Cook groups (Matisoo-Smith et al. 1998, 15148–15149) support Cachola-Abad’s first point. Voyages between Hawai‘i and other East Polynesian archipelagos probably did, as Cachola-Abad suggests, take place throughout some 600 years of Hawaii’s early history, ca. ad 800– 1400. However, Fornander’s analysis, dated according to a revised genealogical chronology based on an average generation length of 20 years, indicates that the voyages described in the traditional sources were limited almost entirely to generations spanning approximately the thirteenth and fourteenth centuries. Cachola-Abad rejects Fornander’s Migratory Period concept in part because she calculates that these traditionally-documented inter- archipelagic voyages were much more widely distributed in time than Fornander indicated. Cachola-Abad’s analysis of the voyagers named in the traditions span at least 23 generations and 460 years with 64 Robert J. Hommon a single earlier isolated outlier, that of Nana‘ulu, that would add nine generations and 180 additional years to the total. In contrast, Fornander’s approach yields about 10 generations spanning 200 years, preceded by the Nana‘ulu outlier some 200 years earlier. The difference between the two versions can be ascribed to the fact that Cachola-Abad based her calculations on the ‘Ulu genealogical line and Fornander on what he terms the Nanaulu genealogy. Fornander explains his choice this way:

[W]hile the Nanaulu genealogy for the period between Kii and Maweke has been assailed by no doubts and by no diversity of opinion among subsequent generations of Hawaiians, the Ulu genealogy has been subject to nu- merous varying constructions, no two agreeing together throughout, and as a Hawaiian genealogy on Hawaiian soil is disproven in several places by its own discrepancy with the Nanaulu line, as well as by the direct testimony of the legends and genealogies of the South Pacific groups. (Fornander 1969, 1:204)

Fornander (1969, 1:188–207) discusses in some detail why he be- lieves the ‘Ulu genealogy includes numerous erroneously inserted additional generations between what are reckoned here as genera- tions 20 and 31. The present discussion is based on Fornander’s genealogical analysis. Excluding the early outlier, 23 voyages from Cachola-Abad’s compilation began with chiefs of generation 21 in the early thirteenth century, and ended abruptly with those of generation 32 in the early fifteenth century. In light of the considerable labor and material investment en- tailed in any long-distance voyage, the relatively small number of voyagers from the southern archipelagos who settled in Hawai‘i proba- bly accounted for an insignificant portion of population growth in the thirteenth and fourteenth century. The phrase Late Voyaging Period used here avoids the connotation of mass population movement of Fornander’s term Migratory Period.

Identical Hawaiian and Maori Genealogical Series

As shown in Table 1.7, the genealogies of Hawaiian and Maori chiefs include series of five successive chiefly names that, adjusted for consonant shifts, are identical (Emory 1977; Fornander 1969). The genealogies containing this series were collected during the middle Watershed: Testing the Limited Land Hypothesis 65 of the nineteenth century and there is no evidence that they were transferred between the two island groups after Western contact. A plausible explanation for the occurrence of the series in two distant locations is that it was part of a genealogy maintained in a region from which one group traveled to Hawai‘i and another to New Zealand. As discussed in the next section, this region of origin was probably that in which Proto Tahitic was spoken, i.e., a broad band including the Southern Cook, Austral, Society, and Tuamotu island groups (Marck 2000, 139).

Table 1.7 Cognate names of chiefs, Hawai‘i and New Zealand

Generation Chief’s Name Birth Year∗ Hommon† Fornander‡ Hawai‘i New Zealand 1170 21 23 ‘Aikanaka Taitangata 1190 22 24 Hema Hema 1210 23 25 Kaha‘i Tawhaki 1230 24 26 Wahieloa Wahieroa 1250 25 27 Laka Rata

∗ Approximate year ad. † This paper. ‡ Fornander (1969, 204).

Assuming a reasonably accurate generation count, the transfer of the five name series to Hawai‘i happened no earlier than the late ad 1200s, the genealogically-reckoned time of Laka/Rata, and no later than the early ad 1400s, by which time long-distance voyaging apparently had ended.

Tahitic Loans in Hawaiian

Though Hawaiian is classified as a Marquesic language, it contains borrowings from Tahitic.

No Hawaiian borrowing from Tahitian stands out quite so boldly as the Hawaiian word kahuna ‘priest’. This word began its journey to Hawai‘i in the form of PPN *tufunga which meant not “priest” but “expert, craftsman,” appar- ently in the context of carpentry and other artisan activi- ties. Only Tahitic languages and Hawaiian are known to have the “priest” sense and there was an odd (sporadic) change in the first vowel of the word in Proto Tahitic 66 Robert J. Hommon

such that the pronunciation had come to be *tahunga . . . Tahitic loans in Hawaiian show innovations of Proto Tahitic and none which seem particular to any of the Tahitic languages. Those Hawaiian borrowings cannot be ordered relative to the breakup of Tahitic and apparently occurred at about the time NZ Maori diverged and before the remaining Tahitic languages developed their separate innovations . . . The date of circa ad 1200 is here adduced by reference to the approximate time of New Zealand’s settlement from the Tahitic heartland, as I have found no ordering arguments to place the divergence of NZ Maori or the Tahitic influences on Hawaiian speech either before or after one another. (Marck 2000, 117, 117 fn. 2, 135)

Marck’s linguistic analysis lends credence to the Late Voyaging Period in that it indicates approximate contemporaneity of the coloniz- ing of New Zealand and the Tahitic borrowing by Hawaiian. Aside from his estimate of ad 1200, presumably derived from archaeological evi- dence, Marck’s analysis of the linguistic data provides corroboration for the Late Voyaging Period that is independent of the archaeological, genealogical, and other information sources noted here.

Late Polynesian Colonization

Recent archaeological research indicates that the widely-distributed islands in the following list were initially colonized during the Late Voyaging Period, between about ad 1200 and 1400.

New Zealand (Aotearoa): thirteenth or fourteenth century

Citing previous estimates (Anderson 1991; Higham and Hogg 1997; McFadgen et al. 1994) of a twelfth or thirteenth century colonization of New Zealand, Higham et al. report the results of new 14C analyses from the site of Wairau Bar:

Taken together, these data suggest that the New Zealand region was settled as part of a mobile, expansive phase of exploration from tropical East Polynesia, no earlier than about 750 years ago. (Higham et al. 1999, 426)

In another recent review, (Anderson et al. 2001, 41) conclude that “there is no evidence of human habitation of New Zealand before 800–600 bp.” Watershed: Testing the Limited Land Hypothesis 67

Norfolk Island: ca. thirteenth century—Enderby Island: ca. thirteenth or fourteenth century—Raoul Island: ca. fourteenth century

Norfolk Island (35 km2) lies about 750 km northwest of New Zealand. Enderby is a small island in the sub-Antarctic Auckland Islands group situated 460 km south of New Zealand. Raoul Island, 1,100 km north- east of New Zealand, is the largest of the Kermadec Islands. The Chatham Islands, situated about 950 km east of New Zealand’s South Island, are not included in this list because the earliest available evidence dates to the fifteenth century, however, Anderson (2005) notes that observations of surface artifacts indicate occupation began several hundred years earlier. Though Anderson (2005) indicates that all the listed islands (as well as the Chathams) were settled from New Zealand, their histories differ. Obsidian sourcing studies indicate that a permanent colony was established on Norfolk Island early in the thirteenth century “from New Zealand via the Kermadecs” (Anderson and White 2001, 41). Matisoo-Smith et al. (1999, 195) conclude from MtDNA analysis that the number of variant lineages of Polynesian rats from Raoul Island, first settled around the fourteenth century, suggest multiple introductions. Some of these lineages may have been introduced directly from central East Polynesia. The evidence on Enderby Island seems to suggest a short-lived occupation by a small group of hunters and foragers (Anderson 2005). These three examples of colonization (especially Enderby) seemingly did not produce the thriving societies we see on larger islands. Nevertheless, I suggest that they represent the same tradition of long-distance voyaging into the unknown that populated the rest of Polynesia.

Rapa: ca. AD 1150–1250

Islands in the Gambier Islands in southeastern East Polynesia appear to have been settled by the beginning of the twelfth century (Anderson et al. 2003).

It is worth noting, however, that Rapa is the southernmost island in East Polynesia and that all of South Polynesia, which lies to the south-west of it, was colonised later again, about ad 1200 . . . Evidence for early settlement (ca. ad 1150 to 1250) is confined to Tangarutu, the most desir- able coastal rockshelter on the island and then expands 68 Robert J. Hommon

to less attractive locations (Akatanui, Angairao, and Taga) starting ca. ad 1350–1450. (Kennett et al. 2006, 341, 351)

Easter Island (Rapa Nui): ca. AD 1200

Hunt and Lipo obtained eight radiocarbon dates from Easter Island’s Anakena Dune Site that support a colonization date of about ad 1200, at least 300 years later than previous long chronology estimates.

Forty-one radiocarbon dates directly associated with de- forestation from multiple sites around the island . . . all fall after 1200 ad, and most cluster after 1300 ad. The long chronology . . . requires that Polynesian settlers had virtually no impact on the island’s and maintained an exceptionally low population growth rate for several centuries until a point of abrupt, dramatic human impacts. (Hunt and Lipo 2006, 1605)

Assuming that the above-listed inferences concerning colonizing events are accurate, long-distance voyaging from central East Poly- nesia to the east, west, and south appears to have been relatively frequent during the Late Voyaging Period. Each successful coloniza- tion event confirmed by archaeological evidence undoubtedly repre- sents only a small sample of multiple expeditions including successful one- and two-way voyages following colonization and unsuccessful attempts for which no evidence survives. Judging from the success- ful establishment of at least the seven listed colonies, many of the voyages of this period evidently were major expeditions led by chiefs capable of commanding the necessary planning, construction, and outfitting, as well as the journey itself. Considering that such voyages were bound for east, west, and south, it seems reasonable that some as well made their way northward to Hawai‘i. Some of the voyages may have been undertaken with knowledge preserved from the time of Hawaii’s colonization.

Collapse of the Central East Polynesian Long-Distance Exchange Network

One of the most interesting aspects of the Late Voyaging Period is that traditional Hawaiian accounts of voyages to and from Kahiki cease abruptly with chiefs of generation 32, whose voyages can be Watershed: Testing the Limited Land Hypothesis 69 genealogically-dated to the first half of the fifteenth century. The tim- ing of this event is virtually identical with the collapse of the central East Polynesian long-distance exchange network about ad 1450 (Allen and Johnson 1997; Collerson and Weisler 2007; Rolett et al. 1997; Shep- pard et al. 1997; Stokes 1933; Weisler 1996, 1997a, 1998, 2002). The maintenance of the long-distance exchange network, which required inter-archipelagic voyaging, and the mid-fifteenth century collapse of that network are documented and dated by archaeological data. The fact that these data are wholly independent of Hawaii’s traditions and their genealogically-based chronology lends credence to the tradi- tional accounts of both the multi-generational period of Hawaii-Kahiki voyaging and its end by ad 1450.

Adze C7727

Adze C7727 from Napuka atoll in the Tuamotus was made of hawaiite from Kaho‘olawe according to analysis of its trace element and isotope chemistry (Collerson and Weisler 2007). The hawaiite is believed to have been transported to Napuka sometime between ad 1200, when the atoll first became habitable, and ad 1450, the archaeologically- dated collapse of the central East Polynesian long-distance exchange network (Collerson and Weisler 2007; Dickinson 2004). Adze C7727 apparently is not only the first physical evidence of south-bound voyaging taking place between Hawai‘i and Kahiki, but situates it within the genealogically-dated Late Voyaging Period.

Transfer of the Sweet Potato, Bottle Gourd, and Breadfruit

According to Green (2005, 48), the sweet potato, having been trans- ferred from South America, was present on Mangaia of the Cook group and probably other islands in East Polynesia’s “central ellipse” by ad 1100 at the latest and “probably a century or two earlier.” Given the diverse evidence for long-distance voyaging throughout Polynesia dur- ing the thirteenth and fourteenth centuries and allowing a century or two for the spread of the crop within the central ellipse, it is plausible that sweet potatoes were part of the cargo carried by colonizers of the islands settled or visited during the Late Voyaging Period. While we currently lack direct evidence that the colonizers of New Zealand and Easter Island brought the sweet potato with them it seems likely that they did so given the evidence in another destination of the late period voyagers. The presence of sweet potato in Hawai‘i 70 Robert J. Hommon by the fourteenth century has recently been demonstrated by radio- carbon analysis of a tuber fragment from the Kohala Field System that produced a calibrated date range of ad 1290–1430 (Ladefoged et al. 2005, 368). It is possible, given Green’s estimate of ad 900–1100 for the sweet potato’s arrival in the central ellipse, that the crop was introduced to Hawai‘i centuries earlier than ad 1300. Small-scale sweet potato cultivation may have begun in the ad 1300s, as noted in Section 1, a few decades before major development of the large rain-fed field systems in the fifteenth century (Allen 2001c; Carson 2005; Coil and Kirch 2005; Hartshorn et al. 2006; Hommon 1976; Kirch et al. 2004; Ladefoged and Graves 2000; Ladefoged et al. 1998, 2003; McCoy 2005; Mulrooney and Ladefoged 2005). However, if we were to posit that the sweet potato arrived in Hawai‘i around ad 1000 or 1100 it would be difficult to explain why the Hawaiians waited 300–400 years to fully exploit a new crop that was so well suited to the extensive leeward regions of their islands. Based on this circumstantial evidence I suggest that the sweet potato may have been introduced to Hawai‘i as late as the thirteenth century. Another cultivar of South American origin that probably arrived in central East Polynesia about the same time as the sweet potato was the bottle gourd (Lagenaria siceraria) (Green 2005; Storey et al. 2007). Since bottle gourds were commonly made into containers, they represent both a cultivar and an type (in the narrow sense) that may have been a relatively late introduction to Hawai‘i. Estab- lishing whether the bottle gourd arrived during the Late Voyaging Period should be easier than dating the arrival of the sweet potato, since fragments of gourd are a great deal more common in the ar- chaeological record than sweet potatoes. Direct 14C dating of curated samples of gourd collected from selected sites would undoubtedly yield interesting results. Allen (2001c, 147) has argued that breadfruit (Artocarpus altilis) was a late introduction in Hawai‘i based on the following evidence. First, Hawaiian breadfruit are all of a single variety, in contrast to the numerous varieties found in central East Polynesian archipelagos. Second, in a “limited review” of the wood identification literature, Allen and Murakami (1999) could find no breadfruit charcoal samples dating earlier than ad 1500. Third, according to tradition, bread- fruit was introduced by a mid-fourteenth century chief named Kaha‘i (Fornander 1917; Handy and Handy 1972). Watershed: Testing the Limited Land Hypothesis 71

On the Road of the ENSO

The El Niño-Southern Oscillation or ENSO is

the quasi-periodic cycle of alternating warming (El Niño) and cooling (La Niña) phases which stems from ocean- atmosphere coupling in the equatorial Pacific. (Allen 2006, 253)

Of interest here are recent papers linking proxy evidence of ENSO sequences and Pacific voyaging around the thirteenth century (Allen 2006; Anderson et al. 2006). Analysis of sediment cores from Laguna Pallcacocha, an alpine lake in southern Ecuador, has yielded a contin- uous proxy record of moderate to strong ENSO frequency spanning the last 7,000 years (Moy et al. 2002b). This record includes a high fre- quency of such ENSO events from the twelfth through the fourteenth centuries (Allen 2006; Anderson et al. 2006; Moy et al. 2002a, b). Anderson et al. (2006, 2) have proposed that “initial colonization across previously uninhabited regions of Remote Oceania was strongly episodic.” Based on the Laguna Pallcacocha record, they hypothesize that the episodes of eastern movement occurred during periods of high El Niño frequency when the weakening of easterly trade winds facilitated sailing eastward on westerly wind reversals. The authors include in one of these episodes, dating to the latter half of the first millennium ad, the settlement of East Polynesia, Hawai‘i, and Easter Island. Anderson et al. (2006, 4, Fig. 3) do not label the period of inter- est here as one of their colonization episodes, however, the Laguna Pallcacocha record upon which they base their hypothesis (Moy et al. 2002a, b) indicates that the frequency of moderate to strong El Niño events in the thirteenth century was the highest of the millennium and the second highest in the 5,000-year ENSO record. Further, the twelfth through the fourteenth centuries together exhibit a greater frequency than that of any other three consecutive centuries in the record. If Anderson et al. are correct in proposing that during previous episodes the frequent El Niño conditions had increased the success of sailing eastward, then more voyaging activity eastward out of central East Polynesia would be expected as well during the Late Voyaging Period. Evidence of such a revival of voyaging, I suggest, may include the colonization of Easter Island and the expeditions 72 Robert J. Hommon between Hawai‘i and Kahiki depicted in the oral traditions of the thirteenth and fourteenth centuries. Anderson et al. note that the westward sailing required to colonize the south Polynesian islands

occurred outside the tropical zone and therefore beyond the El Niño wind changes described here. Experimental voyaging has shown that subtropical easterlies in summer provide the best passage from East to South Polynesia . . . and it is possible that the strengthening of subtropical easterlies which occurs during El Niño episodes was a factor in the timing. (Anderson et al. 2006, 3–4)

ENSO Warming, Drought, and Flooding

In addition to shifts in wind patterns, the high frequency of moderate to strong ENSO episodes during the twelfth through the fourteenth centuries evident in the Laguna Pallcacocha proxy data probably was accompanied by a heightened likelihood of extreme weather anomalies including warming, sustained droughts, and heavy rainfalls. For example, Allen (2006, 530) points out that a warming trend similar to that indicated around ad 1300 by proxy data from Palmyra Atoll coral, can result in a reduction in marine resources due to increased bleaching and mortality of corals such as resulted from the 1997–98 El Niño. ENSO-driven weather conditions can vary greatly from region to region and can affect livelihood in different ways in different eco- zones within a single valley as Field (2004, 84–85) observes in her intensive study of Sigatoka Valley, Viti Levu, Fiji. In Fiji, La Niña episodes following El Niño droughts have been observed to bring continued drought in some cases and high rainfall in others. In her discussion of the rise of social complexity in Fiji, Field focuses on the transition period between the Little Climatic Optimum and the Little Ice Age, dating to about the fourteenth and fifteenth centuries, a span that overlaps the last half of the Late Voyaging Period. Field (2004, 95) hypothesizes that the apparent emergence of social complexity following the fifteenth century may have been related to the array of competitive strategies that Fijians developed to deal with stresses arising from the uncertainties posed by the high ENSO frequency of the transition period. Perhaps the increased frequency of long-distance voyaging by East Polynesians during the thirteenth Watershed: Testing the Limited Land Hypothesis 73 and fourteenth centuries in part evinces a different strategy, that of escaping the extreme weather conditions that brought on famine, flooding, and other disasters by seeking out more salubrious lands to colonize.

Conclusion

The disparate evidence summarized above delineates a Late Voyag- ing Period spanning the thirteenth and fourteenth centuries during which increasingly frequent long-distance voyages from central East Polynesia colonized previously uninhabited islands and renewed or intensified contact with Hawai‘i. The droughts and other weather anomalies that accompanied the unusually frequent moderate to strong ENSO episodes of these centuries may have encouraged people to seek more salubrious environments and at the same time offered otherwise rare conditions for sailing to previously undiscovered is- lands. Well-provisioned colonial expeditions apparently established viable communities on Easter Island, Rapa, and New Zealand and, perhaps underlining the intensity of the colonizing effort of the time, on small isolated islands hundreds of kilometers to the north, south and east of New Zealand. Though the current estimated date for the transfer of the sweet potato from South America to central East Polynesia precedes the Late Voyaging Period, the cultivar probably was carried to the newly-established colonies during this period. The traditional accounts of as many as 24 voyages between Hawai‘i and Kahiki led by Hawaiian and foreign chiefs can be dated by a count of 20-year generations to a period closely approximating the thirteenth and fourteenth century span of the Late Voyaging Period. Additional evidence for HawaiiâA˘Zs´ inter-archipelagic contacts during the Late Voyaging Period includes the borrowing of Hawaiian words from Tahitic around the time of the Maori linguistic divergence and the introduction of a Hawaiian genealogical sequence of five names that is identical with a Maori sequence, both of which plausibly date from the late thirteenth or early fourteenth century, together with the possible transfer of the sweet potato, bottle gourd, and breadfruit. I suggest that a major source of the decoupling of the Hawaiian classes and the “erosion of the lineage order” to which Sahlins (1992) refers was the tendency of the Hawaiian chiefs to deemphasize con- nections with indigenous lineages as they sought links with respected Kahiki chiefs who arrived during the Late Voyaging Period. The ef- fects of such a process were apparently widespread, considering that 74 Robert J. Hommon succeeding generations of prominent chiefs on all the islands could trace ascent to Kahiki chiefly lines. The Kahiki chiefs traditionally identified as commanders of voyag- ing expeditions apparently were revered as chiefs of high rank. Those who settled in Hawai‘i, or at least those who stayed long enough to ap- pear in the genealogies of later chiefs, initially had no connection with indigenous Hawaiian lineages. Although they undoubtedly were ac- cepted into their adoptive societies, in some sense the voyaging chiefs were considered to have remained foreigners whose home islands and genealogies were elsewhere, as demonstrated by the fact that tradi- tions preserved the link to Kahiki for 400 years after long-distance voyaging had ceased. I suggest that in the competition for support of and marriage alliances with the Kahiki chiefs and their descendants, Hawaiian chiefs came to identify with the stranger chiefs and tended to emulate their disconnection from the indigenous lineage tradition. In this disconnection of a chiefship “intruding on the land and people like a foreign element” (Kirch and Sahlins 1992, 1:192) may lie the roots of the ancient Hawaiian transformation to which Sahlins refers.

References

Allen, J., editor (1987a). Five Upland ‘Ili: Archaeological and Historical Investigations in the Kane‘ohe Interchange, Interstate Highway H- 3, Island of O‘ahu. Number 87–1 in Departmental Report Series. Department, B. P. Bishop Museum, Honolulu.

Allen, J. (1991). The role of agriculture in the of the pre- contact Hawaiian state. Asian Perspectives, 30(1):117–132.

Allen, J. (1992). Farming in Hawai‘i from colonisation to contact: Radiocarbon chronology and implications for cultural change. New Zealand Journal of Archaeology, 14:45–66.

Allen, M. S. (1983). Report on the palaeoethnobotanical remains recov- ered from Kaho‘olawe Island. In Hommon (1983), appendix A, pages 153–167.

Allen, M. S. (1987b). Kahoolawe archaeobotanical materials. In Rosendahl et al. (1987), appendix A, pages A–1–A–35.

Allen, M. S. (2001a). Cultural setting and recent garden history. In Allen (2001b), chapter 1, pages 1–13. Watershed: Testing the Limited Land Hypothesis 75

Allen, M. S., editor (2001b). Gardens of Lono: Archaeological Investiga- tions at the Amy B. H. Greenwell Ethnobotanical Garden, Kealakekua, Hawai‘i. Bishop Museum Press, Honolulu.

Allen, M. S. (2001c). The Kona field system in spatial and temporal perspective. In Allen (2001b), chapter 11, pages 137–155.

Allen, M. S. (2006). New ideas about late climate variability in the Central Pacific. Current Anthropology, 47(3):521–535.

Allen, M. S. and Johnson, K. T. M. (1997). Tracking ancient patterns of interaction: Recent geochemical studies in the Southern Cook Islands. In Weisler (1997b), chapter 7, pages 111–133.

Allen, M. S. and Murakami, G. M. (1999). Lana‘i island’s arid lowland vegetation in late . Pacific Science, 53:88–112.

American Red Cross (2007). The influenza pandemic of 1918 and the Red Cross response. http://www.redcross.org/museum/ history/influenza.asp#more. Accessed March 8, 2007.

Anderson, A. (1991). The chronology of colonization in New Zealand. Antiquity, 65:767–795.

Anderson, A. (2005). Subpolar settlement in South Polynesia. Antiquity, 79:791–800.

Anderson, A., Chappell, J., Gagan, M., and Grove, R. (2006). Prehistoric maritime migration in the Pacific Islands: An hypothesis of ENSO forcing. The Holocene, 16(1):1–6.

Anderson, A., Conte, E., Kirch, P. V., and Weisler, M. (2003). Cultural chronology in Mangareva (Gambier Islands), French Polynesia: Ev- idence from recent radiocarbon dating. Journal of the Polynesian Society, 112:119–140.

Anderson, A., Higham, T., and Wallace, R. (2001). The radiocarbon chronology of the Norfolk Island archaeological sites. In Anderson and White (2001), pages 33–42.

Anderson, A. and White, P., editors (2001). The of Norfolk Island, Southwest Pacific. Number 27 in Records of the Australian Museum, Supplement. Australian Museum, Sydney. 76 Robert J. Hommon

Andrews, L. (2003 [1865]). A Dictionary of the . Island Heritage, Waipahu, HI.

Arago, J. (1823). Narrative of a Voyage Round the World, in the Uranie and Physicienne Corvettes. Treuttel & Wurtz, Treuttel, jun. & Richter, London.

Athens, J. S., Tuggle, H. D., Ward, J. V., and Welch, D. J. (2002). Avi- faunal extinctions, vegetation change, and Polynesian impacts in prehistoric Hawai‘i. Archaeology in Oceania, 37:57–78.

Ballard, C., Brown, P., Bourke, R. M., and Harwood, T., editors (2005). The Sweet Potato in Oceania: A Reappraisal, volume 56 of Oceania Monographs. University of Sydney, Sydney.

Barabási, A.-L. and Bonabeau, E. (2003). Scale-free networks. Scientific American, 288(5):60–69.

Barrère, D. B. (1969). The Kumuhonua Legends: A Study of Late 19th Century Hawaiian Stories of Creation and Origins. Number 3 in Pacific Anthropological Records. Anthropology Department, B. P. Bishop Museum, Honolulu.

Bennett, W. C. (1931). Archaeology of Kauai. Number 80 in B. P. Bishop Museum Bulletin. Bishop Museum Press, Honolulu.

Blaikie, P. and Brookfield, H. (1987). Land Degradation and Society. Routledge, London.

Bronk Ramsey, C. (2001). Development of the radiocarbon calibration program OxCal. Radiocarbon, 43(2A):355–363.

Bronk Ramsey, C. (2007). OxCal. http://c14.arch.ox.ac.uk/ embed.php?File=oxcal.html. Accessed February 22, 2007.

Burtchard, G. C. and Tomonari-Tuggle, M. J. (2004). Agriculture on lee- ward Hawai‘i Island: the Waimea agricultural system reconsidered. Hawaiian Archaeology, 9:50–73.

Bushnell, O. A. (1993). The Gifts of Civilization: Germs and Genocide in Hawai‘i. University of Hawaii Press, Honolulu.

Cachola-Abad, C. K. (1993). Evaluating the orthodox dual settlement model for the main Hawaiian Islands: An analysis of artefact dis- tribution and Hawaiian oral traditions. In Graves, M. W. and Green, Watershed: Testing the Limited Land Hypothesis 77

R. C., editors, The Evolution and Organisation of Prehistoric Society in Polynesia, number 19 in Monographs of the New Zealand Ar- chaeological Association, pages 13–32. New Zealand Archaeological Association, Auckland.

Carson, M. T. (2005). A radiocarbon dating synthesis for Kaua‘i. In Carson and Graves (2005), pages 11–32.

Carson, M. T. and Graves, M. W., editors (2005). Na Mea Kahiko o Kaua‘i: Archaeological Studies in Kaua‘i. Number 2 in Special Publication. Society for Hawaiian Archaeology, Honolulu.

Clague, D. A. (1998). Geology. In Juvik and Juvik (1998), pages 37–46.

Clark, J. T. and Kirch, P. V., editors (1983). Archaeological Investigations of the Mudlane-Waimea-Kawaihae Road Corridor, Island of Hawai‘i: An Interdisciplinary Study of an Environmental Transect. Number 83–1 in Departmental Report Series. Anthropology Department, B. P. Bishop Museum, Honolulu.

Coil, J. and Kirch, P. V. (2005). An ipomoean landscape: Archaeology and the sweet potato in Kahikinui, Maui, Hawaiian Islands. In Ballard et al. (2005), pages 71–84.

Collerson, K. D. and Weisler, M. I. (2007). Stone adze compositions and the extent of ancient Polynesian voyaging and trade. Science, 317:1907–1911.

Cordy, R. (1974). The Tahitian migration to Hawaii ca. 1100–1300 ad: An argument against its occurrence. New Zealand Archaeological Association Newsletter, 17:65–76.

Cordy, R. and Kaschko, M. W. (1980). Prehistoric archaeology in the Hawaiian Islands: Land units associated with social groups. Journal of Field Archaeology, 7:403–416.

Cordy, R. H. (1994). A Regional Synthesis of Hamakua District, Island of Hawai‘i. Historic Preservation Division, Department of Land and Natural Resources, State of Hawaii, Honolulu.

Cordy, R. H. (2000). Exalted Sits the Chief: The Ancient History of Hawai‘i Island. Mutual, Honolulu. 78 Robert J. Hommon

Darwin, F. and Seward, A. C., editors (1903). More Let- ters of Charles Darwin: A Record of His Work in a Se- ries of Hitherto Unpublished Letters, volume 1. John Murray, London. http://darwin-online.org.uk/content/frameset? itemID=F1548.1&viewtype=side&pageseq=1.

Dega, M. F. and Kirch, P. V. (2002). A modified culture history of Anahulu Valley, O‘ahu, Hawai‘i and its significance for Hawaiian prehistory. Journal of the Polynesian Society, 111(2):107–126.

Dickinson, W. R. (2004). Impacts of eustacy and hydro-isostacy on the evolution and landforms of Pacific atolls. Palaeogeography, Palaeoclimatology, Palaeoecology, 213:251–269.

Dixon, B., Conte, P. J., Nagahara, V., and Hodgins, W. K. (2002). Settle- ment patterns and subsistence strategies in Kahikinui, Maui. Hawai- ian Archaeology, 8:13–32.

Dye, T. S. (1994). Population trends in Hawai‘i before 1778. Hawaiian Journal of History, 28:1–20.

Dye, T. S. (2000). Effects of 14C sample selection in archaeology: An example from Hawai‘i. Radiocarbon, 42(2):203–217.

Dye, T. S. (2001). Hawaii’s dynamic built environment or what is a site? Paper presented at the Society for Hawaiian Archaeology Speaker Series, B.P. Bishop Museum, Honolulu, Hawaii.

Dye, T. S. (2004a). How to fix the inventory rule. Hawaiian Archaeology, 9:123–132.

Dye, T. S. (2004b). The tyranny of the traditional Hawaiian built envi- ronment. Paper presented at the University of Arizona Anthropology Colloquium Series, Tucson, AZ.

Dye, T. S. and Komori, E. K. (1992a). Computer programs for creating cumulative probability curves and annual frequency distribution dia- grams with radiocarbon dates. New Zealand Journal of Archaeology, 14:35–43.

Dye, T. S. and Komori, E. K. (1992b). A pre-censal population history of Hawai‘i. New Zealand Journal of Archaeology, 14:113–128. Watershed: Testing the Limited Land Hypothesis 79

Dye, T. S., Major, M., Desilets, M. E., and Maigret, M. B. (2002). Ar- chaeological Inventory Survey of Portions of Kekaha Kai State Park. Prepared for Group 70 International. T. S. Dye & Colleagues, Archae- ologists, Honolulu.

Earle, T. (1978). Economic and Social Organization of a Complex Chief- dom: The Halelea District, Kaua‘i, Hawaii. Number 63 in Anthropo- logical Papers. Museum of Anthropology, University of Michigan, Ann Arbor, MI.

Earle, T. K. (1980). Prehistoric irrigation in the Hawaiian Islands: An evaluation of evolutionary significance. Archaeology and Physical Anthropology in Oceania, 15:1–28.

Ellis, W. (1963). Journal of William Ellis. Advertiser Publishing Com- pany, Honolulu.

Emory, K. P. (1928). Archaeology of Nihoa and Necker Islands. Num- ber 53 in B. P. Bishop Museum Bulletin. B. P. Bishop Museum, Hon- olulu.

Emory, K. P. (1968). East Polynesian relationships as revealed through adzes. In Yawata and Sinoto (1968), pages 151–169.

Emory, K. P. (1977). Comparison of Polynesian genealogies in the Bishop Museum. Pacific Studies, 1(1):1–14.

Field, J. S. (2004). Environmental and climatic considerations: A hypothesis for conflict and the emergence of social complexity in Fijian prehistory. Journal of Anthropological Archaeology, 23:79–99.

Flannery, K. V. (1972). The cultural evolution of civilizations. Annual Review of Ecology and Systematics, 3:399–426.

Fornander, A. (1917). Fornander Collection of Hawaiian Antiquities and Folk-lore, volume 4, Part 2 of Memoirs of the B. P. Bishop Mu- seum. Bishop Museum Press, Honolulu. Translations revised and illustrated with notes by Thomas G. Thrum.

Fornander, A. (1969). An Account of the Polynesian Race, its Origins and Migrations, and the Ancient History of the Hawaiian People to the Times of Kamehameha. Charles E. Tuttle, Rutland, VT. 3 vols., originally published in 1878–1885. 80 Robert J. Hommon

Friedman, J. (1981). Notes on structure and history in Oceania. Folk, 23:275–295.

Giambelluca, T. W. and Schroeder, T. A. (1998). Climate. In Juvik and Juvik (1998), pages 49–59.

Goto, A. (1986). Prehistoric Ecology and Economy of Fishing in Hawaii: An Ethnoarchaeological Approach. PhD thesis, University of Hawaii.

Green, R. C. (1974). Tahiti-Hawaii ad 1100–1300: Further comments. New Zealand Archaeological Association Newsletter, 17:206–210.

Green, R. C. (2005). Sweet potato transfers in Polynesian prehistory. In Ballard et al. (2005), pages 43–62.

Hack, J. T. (1942). The Changing Physical Environment of the Hopi Indians of Arizona, volume 35 of Papers of the Peabody Museum of American Archaeology and . Harvard University, Cam- bridge, MA.

Handy, E. S. C. (1940). The Hawaiian Planter, volume 1. Bishop Museum Press, Honolulu.

Handy, E. S. C. and Handy, E. G. (1972). Native Planters in Old Hawaii: Their Life, Lore, and Environment. Number 233 in B. P. Bishop Museum Bulletin. Bishop Museum Press, Honolulu. With the collab- oration of .

Hartshorn, A. S., Chadwick, O. A., Vitousek, P. M., and Kirch, P. V. (2006). Prehistoric agricultural depletion of soil nutrients in Hawai‘i. Proceedings of the National Academy of , 103(29):11092– 11097.

Higham, T., Anderson, A., and Jacomb, C. (1999). Dating the first New Zealanders: The chronology of Wairau Bar. Antiquity, 73:420–427.

Higham, T. and Hogg, A. G. (1997). Evidence for late Polynesian colonisation of New Zealand: University of Waikato radiocarbon measurements. Radiocarbon, 39:149–192.

Hommon, R. J. (1969). Report 4: An interim report on archaeological zone 1. In Green, R. C., editor, Makaha Valley Historical Project: Interim Report No. 1, number 4 in Pacific Anthropological Records, pages 41–53. Anthropology Department, B. P. Bishop Museum, Hon- olulu. Watershed: Testing the Limited Land Hypothesis 81

Hommon, R. J. (1970). Final report on the upper valley survey. In Green, R. C., editor, Makaha Valley Historical Project: Interim Report No. 2, number 10 in Pacific Anthropological Records, pages 105–121. Anthropology Department, B. P. Bishop Museum, Honolulu.

Hommon, R. J. (1972). Hawaiian cultural systems and archaeological site patterns. Master’s thesis, University of Arizona, Tucson.

Hommon, R. J. (1973). Archaeology and the primitive state: The Hawaiian example. Paper presented at the 38th Annual Meeting, Society for American Archaeology, San Francisco.

Hommon, R. J. (1975). Use and Control of Hawaiian Inter-Island Chan- nels, Volume I: Polynesian Hawaii ad 1400–1784. Prepared for the Office of the Governor of Hawaii. National Park Service Pacific West Regional Office, Honolulu.

Hommon, R. J. (1976). The Formation of Primitive States in Pre-Contact Hawaii. PhD thesis, University of Arizona, Tucson, AZ.

Hommon, R. J. (1980). Multiple resources nomination form for Kaho‘olawe archaeological sites. National Register of Historic Places.

Hommon, R. J. (1983). Kaho‘olawe Archaeological Excavations, 1981. Prepared for U.S. Navy, Pacific Division, Naval Facilities Engineering Command. Science Management, Honolulu.

Hommon, R. J. (1986). Social evolution in ancient Hawaii. In Kirch, P. V., editor, Island Societies: Archaeological Approaches to Evolu- tion and Transformation, pages 55–68. Cambridge University Press, Cambridge.

Hommon, R. J. (1989). The Kahiki connection: Extra-archipelagic contact and the development of complex polities in ancient Hawaii. Paper presented at the Circum-Pacific Prehistory Conference, Seattle, WA.

Hommon, R. J. (1992). The view from out on a limb. New Zealand Journal of Archaeology, 14:151–158.

Hommon, R. J. (1995). Cultural complex adaptive systems. Paper presented at the Seminar on Complexity, Santa Fe Institute, Santa Fe New Mexico. 82 Robert J. Hommon

Hommon, R. J. (1996). Social complex adaptive systems: Some Hawai- ian examples. In Davidson, J. M., Irwin, G., Leach, B. F., Pawley, A. K., and Brown, D., editors, Oceanic Culture History: Essays in Honour of Roger Green, New Zealand Journal of Archaeology Special Publication, pages 579–590. New Zealand Journal of Archaeology, Dunedin North, NZ.

Hommon, R. J. (1999). C-cast, polyarchy, ecaptation, stratified control hierarchies, and the simplification necessary for the emergence of large-scale society. Paper presented at the 64th Annual Meeting, Society for American Archaeology, Chicago.

Hommon, R. J. (2000). Can a small fish swallow a big one? or Orwell among the Hawaiians. In Diehl, M. W., editor, Hierarchies in Action: Cui Bono?, number 27 in Occasional Paper, pages 134–151. Center for Archaeological Investigations, Southern Illinois University, Car- bondale, IL.

Hommon, R. J. (2001). The emergence of large-scale society in Hawai‘i. In Stevenson, C. M., Lee, G., and Morin, F. J., editors, Pacific 2000: Proceedings of the Fifth International Conference on Easter Island and the Pacific, pages 141–149. Bearsville Press, Los Osos, CA.

Hommon, R. J. (2005). As water for fish: Human groups as complex ecaptive systems. In Beekman, C. S. and Baden, W. W., editors, Nonlinear Models for Archaeology and Anthropology: Continuing the Revolution, pages 13–33. Ashgate Publishing, Burlington, VT.

Hommon, R. J. (2007). Sweet potatoes, population dynamics, and political evolution: New light on Hawaii‘s ancient history. Paper presented at the 20th Annual Hawaiian Archaeology Conference, Keauhou, HI.

Hommon, R. J. (2008). Boom, bust, and the rise of archaic states in ancient Hawaii. Paper presented at the 73rd Annual Meeting, Society for American Archaeology, Vancouver, BC.

Hommon, R. J. and Barrera, Jr., W. M. (1971). Archaeological Survey of Kahana Valley, Koolau District, Island of Oahu. Number 71–3 in Departmental Report. Anthropology Department, B. P. Bishop Museum, Honolulu.

Hunt, T. L. and Lipo, C. P. (2006). Late colonization of Easter Island. Science, 311:1603–1606. Watershed: Testing the Limited Land Hypothesis 83

Johnson, G. (1982). Organizational stress and scalar stress. In Ren- frew, C., Rowlands, M. J., and Seagraves, B. A., editors, Theory and Explanation in Archaeology: The Southampton Conference, pages 389–421. Academic Press, New York.

Judd, B. (1974). Voyages to Hawaii Before 1860: A Record Based on Historical Narratives in the Libraries of the Hawaiian Mission Children’s Society and the Hawaiian Historical Society, extended to March 1860. University Press of Hawaii for Hawaiian Mission Children’s Society, Honolulu. Enlarged and edited by Helen Yonge Lind.

Juvik, S. P. and Juvik, J. O., editors (1998). Atlas of Hawai‘i. University of Hawaii Press, Honolulu, third edition.

Kamakau, S. M. (1961). Ruling Chiefs of Hawaii. Press, Honolulu.

Kamakau, S. M. (1964). Ka Po‘e Kahiko: The People of Old. Number 51 in B. P. Bishop Museum Special Publication. Bishop Museum Press, Honolulu.

Kamakau, S. M. (1991). Tales and Traditions of the People of Old: Na Mo‘olelo a ka Po‘e Kahiko. Bishop Museum Press, Honolulu. Translated from the Newspapers Ka Nupepa Kuokoa and Ke Au Okoa by Mary Kawena Pukui. Arranged and edited by Dorothy B. Barrère.

Kaschko, M. W. (1973). Functional analysis of the trail system of the Lapakahi area, North Kohala. In Tuggle and Griffin (1973), pages 127–144.

Kay, D. E. (1973). Root Crops. Number 2 in Tropical Products Institute Crop and Product Digest. Tropical Products Institute, London.

Kennett, D., Anderson, A., Preble, M., Conte, E., and Southon, J. (2006). Prehistoric human impacts on Rapa, French Polynesia. Antiquity, 80:340–354.

Keyfitz, N. (1966). How many people have lived on the earth? Demog- raphy, 3(2):581–582.

Kirch, P. V. (1984). The Evolution of the Polynesian Chiefdoms. New Studies in Archaeology. Cambridge University Press, Cambridge. 84 Robert J. Hommon

Kirch, P. V. (1985). Feathered Gods and Fishhooks: An Introduction to Hawaiian Archaeology and Prehistory. University of Hawaii Press, Honolulu.

Kirch, P. V. (1994). The Wet and the Dry: Irrigation and Agricultural Intensification in Polynesia. University of Chicago Press, Chicago.

Kirch, P. V. (2000). On the Road of the Winds: An Archaeological History of the Pacific Islands Before European Contact. University of California Press, Berkeley, CA.

Kirch, P. V. (2003). Hawaiian biocomplexity project. http: //sscl.berkeley.edu/~oal/research/biocomplexity/ biocomplexity.htm. Accessed August 30, 2006.

Kirch, P. V. (2005). From Chiefdom to Archaic State: Social Evolution in Hawaii. Grace Elizabeth Shallit Lecture Series. Department of Anthropology, Brigham Young University, Provo, UT.

Kirch, P. V. (2007). “Like shoals of fish”: Archaeology and population in pre-contact Hawai‘i. In Kirch and Rallu (2007), pages 52–69.

Kirch, P. V. and Green, R. C. (2001). Hawaiki, Ancestral Polynesia: An Essay in Historical Anthropology. Cambridge University Press, Cambridge.

Kirch, P. V., Hartshorn, A. S., Chadwick, O. A., Vitousek, P. M., Sherrod, D. R., Coil, J., Holm, L., and Sharp, W. D. (2004). Environment, agri- culture, and settlement patterns in a marginal Polynesian landscape. Proceedings of the National Academy of Sciences, 101:9936–9941.

Kirch, P. V. and Rallu, J.-L., editors (2007). The Growth and Collapse of Pacific Island Societies: Archaeological and Demographic Perspec- tives. University of Hawai‘i Press, Honolulu.

Kirch, P. V. and Sahlins, M. (1992). Anahulu: The Anthropology of History in the Kingdom of Hawaii. University of Chicago Press, Chicago. 2 vols.

Kirch, P. V. and Sharp, W. D. (2005). Coral 230Th dating of the impo- sition of a ritual contral hierarchy in precontact Hawaii. Science, 307:102–103. Watershed: Testing the Limited Land Hypothesis 85

Ladefoged, T. N. and Graves, M. W. (2000). Evolutionary theory and the historical development of dry-land agriculture in North Kohala, Hawai‘i. American Antiquity, 65(3):423–448.

Ladefoged, T. N. and Graves, M. W. (2007). Modeling agricultural development and demography in Kohala, Hawaii. In Kirch and Rallu (2007), pages 70–89.

Ladefoged, T. N., Graves, M. W., and Coil, J. H. (2005). The introduction of sweet potato in Polynesia: Early remains in Hawai‘i. Journal of the Polynesian Society, 114(4):359–373.

Ladefoged, T. N., Graves, M. W., and Jennings, R. (1998). Dryland agricultural expansion and intensification in Kohala, Hawai‘i Island. Antiquity, 70:861–880.

Ladefoged, T. N., Graves, M. W., and McCoy, M. (2003). Archaeological evidence for agricultural development in Kohala, Island of Hawaii. Journal of , 30:923–940.

Lekson, S. H. (1990). Cross-cultural perspectives on the community. In Turan, M., editor, Vernacular Architecture: Paradigms of Environ- mental Response, pages 122–145. , Brookfield, VT.

Linnekin, J. (1990). Sacred Queens and Women of Consequence. Uni- versity of Michigan Press, Ann Arbor, MI.

Lucas, P. F. N. (1995). A Dictionary of Hawaiian Legal Land-Terms. Native Hawaiian Legal Corporation and University of Hawai‘i Com- mitteee for the Preservation and Study of Hawaiian Language, Art and Culture, Honolulu.

Lyons, C. J. (1875). Land matters in Hawaii. The Islander, 1(18):103–104, 1(19):111, 1(20):118–119.

Major, M. and Allen, M. S. (2001). Archaeological excavations, 1996. In Allen (2001b), pages 75–98.

Malo, D. (1951). Hawaiian Antiquities (Moolelo Hawaii). Number 2 in B. P. Bishop Museum Special Publication. Bishop Museum Press, Honolulu, second edition. Translated by Nathaniel B. Emerson.

Maly, K. (1998). “Kekaha Wai ‘Ole o na¯ Kona,” A Report on Archival and Historical Documentary Research, and Oral History Inter- views for Kekaha Kai State Park, Ahupua‘a of Kaulana, Mahai‘ula, 86 Robert J. Hommon

Makalawena, Awake‘e, Manini‘owali,¯ and Kuki‘o,¯ District of North Kona, Island of Hawai‘i (TMK Overview:7–2). Prepared for Group 70 International. Kumu Pono Associates, Hilo, HI.

Marck, J. (2000). Topics in Polynesian Language and Culture History. Number 504 in Pacific Linguistics. Pacific Linguistics, Canberra, AU.

Marcus, J. and Feinman, G. M. (1998). Introduction. In Feinman, G. M. and Marcus, J., editors, Archaic States, pages 3–13. School of American Research, Santa Fe, NM.

Matisoo-Smith, E., Roberts, R. M., Irwin, G. J., Allen, J. S., Penny, D., and Lambert, D. M. (1998). Patterns of prehistoric human mobility in Polynesia indicated by mtDNA from the Pacific rat. Proceedings of the National Acadamy of Sciences, USA, 95:15145–15150.

Matisoo-Smith, E., Sutton, D. G., Ladefoged, T. N., Lambert, D. M., and Allen, J. S. (1999). Prehistoric mobility in Polynesia: MtDNA variation in Rattus exulans from the Chatham and Kermadec Islands. Asian Perspectives, 38(2):186–199.

McArthur, N. (1968). Island Populations of the Pacific. University of Hawaii Press, Honolulu.

McCoy, M. (2005). The development of the Kalaupapa field system, Moloka‘i Island, Hawai‘i. Journal of the Polynesian Society, 114(4):339– 358.

McCoy, P. C. (1990). Subsistence in a ‘non-subsistence’ environment: Factors of production in a Hawaiian alpine desert quarry. In Yen, D. E. and Mummery, J. M. J., editors, Pacific Production Systems: Ap- proaches to Economic Prehistory, number 18 in Occasional Papers in Prehistory, pages 85–119. Department of Prehistory, Research School of Pacific Studies, The Australian National University, Canberra, AU.

McCoy, P. C. (1999). Neither here nor there: A rites of passage site on the eastern fringes of the Mauna Kea adze quarry, Hawai‘i. Hawaiian Archaeology, 7:11–34.

McElroy, W. K. (2007). The Development of Irrigated Agriculture in Wailau Valley, Moloka‘i Island, Hawai‘i. PhD thesis, University of Hawai‘i at M¯anoa, Honolulu. Watershed: Testing the Limited Land Hypothesis 87

McFadgen, B. G., Knox, F. B., and Cole, T. R. L. (1994). Radiocarbon cali- bration curve variations and their implications for the interpretation of New Zealand prehistory. Radiocarbon, 36(2):221–236.

Menzies, A. (1920). Hawaii Nei 128 Years Ago. W. F. Wilson, Honolulu.

Miami University (2007). The population of the United States, 1790– 2005. http://www.eh.net/databases/uspop/uspop.txt. Ac- cessed March 8, 2007.

Mills, P. R. (2005). Connecting with Kaua‘i’s canyonlands. In Carson and Graves (2005), pages 47–61.

Moy, C. M., Seltzer, G. O., Rodbell, D. T., and Anderson, D. M. (2002a). Laguna Pallcacocha Sediment Color Intensity Data. Number 2002–76 in IGBP PAGES/World Data Center for Paleoclimatology Data Con- tribution Series. NOAA/NCDC Paleoclimatology Program, Boulder, CO.

Moy, C. M., Seltzer, G. O., Rodbell, D. T., and Anderson, D. M. (2002b). Variability of El Niño/Southern oscillation activity at millenial timescales during the Holocene epoch. , 420:162–165.

Mulrooney, M. A. and Ladefoged, T. N. (2005). Hawaiian Heiau and agricultural production in the Kohala dryland field system. Journal of the Polynesian Society, 114(1):45–67.

Murabayashi, E. T. (1970). An analysis of the soils and their early agricultural implications in the Lapakahi area of Kohala, Hawaii. In Newman (1970), appendix I, pages 251–279.

Murakami, G. M. (1983a). Analysis of charcoal from archaeological contexts. In Clark and Kirch (1983), report 20, pages 514–524.

Murakami, G. M. (1983b). Identification of charcoal from Kaho‘olawe archaeological sites. In Hommon (1983), appendix B, pages 168–188.

Nakamura, W. T. (1933). A study of the variation in annual rainfall of Oahu Island (Hawaiian Islands) based on the law of probabilities. Monthly Weather Review, 61:354–360.

Newman, T. S. (1970). Hawaiian Fishing and Farming on the Island of Hawaii in ad 1778. Department of Land and Natural Resources, Division of State Parks, Honolulu. 88 Robert J. Hommon

Norman, M. J. T., Pearson, C. J., and Searle, P. G. E. (1984). The Ecology of Tropical Food Crops. Cambridge University Press, Cambridge.

Oliver, D. L. (1974). Ancient Tahitian Society. The University Press of Hawaii, Honolulu. 3 vols.

Portlock, N. (1968 [1789]). A Voyage Round the World . . . Performed in 1785, 1786, 1787 and 1788. Da Capo Press, New York.

Pukui, M. K. (1983). Olelo Noeau: Hawaiian Proverbs and Poetical Sayings. Number 71 in B. P. Bishop Museum Special Publication. Bishop Museum Press, Honolulu.

Pukui, M. K. and Elbert, S. H. (1973). Hawaiian Dictionary. University of Hawaii Press, Honolulu, first edition.

Pukui, M. K., Hartig, E. W., and Lee, C. A. (1972). Nan¯ a¯ i ke Kumu (Look to the Source). Hui H¯anai (Queen Lili‘uokalani Children’s Center), Honolulu. 2 vols.

Purseglove, J. W. (1968). Tropical Crops. Dicotyledons. Longmans, Harlow.

Ramsey, C. B. (1995). Radiocarbon calibration and analysis of stratig- raphy: The OxCal program. In Cook, G. T., Harkness, D. D., Miller, B. F., and Scott, E. M., editors, Proceedings of the 15th International 14C Conference, volume 37 of Radiocarbon, pages 425–430.

Rick, J. W. (1987). Dates as data: An examination of the Peruvian preceramic radiocarbon record. American Antiquity, 52:55–73.

Riley, T. J. (1975). Survey and excavations of the aboriginal agricultural system. In Kirch, P. V. and Kelly, M., editors, Prehistory and Ecology in a Windward Hawaiian Valley: Halawa Valley, Molokai, number 24 in Pacific Anthropological Records, pages 79–115. Anthropology Department, B. P. Bishop Museum, Honolulu.

Ripperton, J. C. and Hosaka, E. Y. (1942). Vegetation Zones of Hawaii. Number 89 in Hawaii Agricultural Experiment Station Bulletin. Uni- versity of Hawaii, Honolulu.

Rolett, B. V., Conte, E., Pearthree, E., and Sinton, J. M. (1997). Marque- san voyaging: Archaeometric evidence for inter-island contact. In Weisler (1997b), chapter 8, pages 134–148. Watershed: Testing the Limited Land Hypothesis 89

Rose, R. G. (1978). Symbols of Sovereignty: Feather Girdles of Tahiti and Hawai‘i. Number 28 in Pacific Anthropological Records. Anthro- pology Department, B. P. Bishop Museum, Honolulu.

Rosendahl, P. H. (1972a). Aboriginal Agriculture and Residence Patterns in Upland Lapakahi, Island of Hawaii. PhD thesis, University of Hawaii, Honolulu.

Rosendahl, P. H. (1972b). Archaeological Salvage of the Hapuna- Anaehoomalu Section of the Kailua-Kawaihae Road, Queen Kaahu- manu Highway, Island of Hawaii. Prepared for Department of Transportation. Anthropology Department, B. P. Bishop Museum, Honolulu.

Rosendahl, P. H., Haun, A. E., Halbig, J. B., Kaschko, M. W., and Allen, M. S. (1987). Kahoolawe Excavations, 1982–83 Data Recov- ery Project, Island of Kahoolawe, Hawaii. Prepared for Department of the Navy, Pacific Division, Naval Facilities Engineering Command. Paul H. Rosendahl, Ph.D., Hilo, HI.

Rosendahl, P. H. and Yen, D. E. (1971). Fossil sweet potato remains from Hawaii. Journal of the Polynesian Society, 80(3):379–385.

Sahlins, M. (1985). Islands of History. University of Chicago Press, Chicago.

Schmitt, R. C. (1968). Demographic Statistics of Hawaii: 1778–1965. University of Hawaii Press, Honolulu.

Schmitt, R. C. (1970a). Famine mortality in Hawaii. Journal of Pacific History, 4:109–115.

Schmitt, R. C. (1970b). The Oku‘u—Hawaii’s greatest epidemic. Hawaii Medical Journal, 29(5):359–364.

Schmitt, R. C. (1971). New estimates of the pre-censal population of Hawaii. Journal of the Polynesian Society, 80:237–243.

Schmitt, R. C. (1973). The Missionary Censuses of Hawaii. Number 20 in Pacific Anthropological Records. Anthropology Department, B. P. Bishop Museum, Honolulu.

Sheppard, P. J., Walter, R., and Parker, R. J. (1997). Basalt sourcing and the development of Cook Island exchange systems. In Weisler (1997b), chapter 6, pages 85–110. 90 Robert J. Hommon

Sinoto, Y. H. (1968). Position of the Marquesas Islands in East Polyne- sian prehistory. In Yawata and Sinoto (1968), pages 111–118.

Smith, K. and Schilt, A. (1973). North Kohala: Agricultural field systems and geographic variables. In Tuggle and Griffin (1973), pages 309– 320.

Spear, R. L. (1992). Settlement and expansion in an Hawai‘ian valley: The archaeological record from North H¯alawa, O‘ahu. New Zealand Journal of Archaeology, 14:79–88.

Stannard, D. E. (1989). Before the Horror: The Population of Hawai‘i on the Eve of Western Contact. Social Science Research Institute, University of Hawaii, Honolulu.

Stokes, J. F. G. (1933). New bases for Hawaiian chronology. Annual Report of the Hawaiian Historical Society, 41:23–65.

Stokes, J. F. G. (1937). Dune sepulture, battle mortality and Kame- hameha’s alleged defeat on Kauai. In Forty-Fifth Annual Report of the Hawaiian Historical Society for the Year 1936, pages 30–46. Hawaiian Historical Society, Honolulu.

Storey, A. A., Ramirez, J. M., Quiroz, D., Burney, D. V., Addison, D., Walter, R., Anderson, A. J., Hunt, T. L., Athens, J. S., Huynen, L., and Matisoo-Smith, E. (2007). Radiocarbon and DNA evidence for a pre- Columbian introduction of Polynesian chickens to Chile. Proceedings of the National Academy of Sciences, 104:10335–10339.

Streck, Jr., C. F. (1992). Prehistoric settlement in the upland portions of the Island of Hawai‘i. New Zealand Journal of Archaeology, 14:99– 111.

Sutton, D. G. and Molloy, M. A. (1989). Deconstructing Pacific palaeode- mography: A critique of density dependent causality. Archaeology in Oceania, 24:31–36.

Tomonari-Tuggle, M. J. (2006). Archaeological Data Recovery Investi- gations for the Northern Portion of Captain Cook Ranch. Prepared for Captain Cook Ranch, LLC. International Archaeological Research Institute, Honolulu.

Trigger, B. G. (2003). Understanding Early Civilizations: A Comparative Study. Cambridge University Press, New York. Watershed: Testing the Limited Land Hypothesis 91

Tuggle, H. D. (1997). The ‘Ewa plain. Hawaiian Archaeology, 6:8–36.

Tuggle, H. D. and Griffin, P. B., editors (1973). Lapakahi, Hawaii: Archaeological Studies. Number 5 in Asian and Pacific Archaeol- ogy Series. Social Science Research Institute, University of Hawaii, Honolulu.

Valeri, V. (1985). Kingship and Sacrifice: Ritual and Society in Ancient Hawaii. University of Chicago Press, Chicago.

Vancouver, G. (1967 [1798]). Voyage of Discovery to the North Pacific Ocean and Round the World, volume III. Da Capo Press, New York.

Vitousek, P. M., Ladefoged, T. N., Kirch, P. V., Hartshorn, A. S., Graves, M. W., Hotchkiss, S. C., Tuljapurkar, S., and Chadwick, O. A. (2004). Soils, agriculture, and society in precontact Hawai‘i. Science, 304:1665–1669.

Weisler, M. I. (1996). Interisland and interarchipelago transfer of stone tools in prehistoric Polynesia. Proceedings of the National Academy of Sciences, 93:1381–1385.

Weisler, M. I. (1997a). Prehistoric long-distance interaction at the margins of Oceania. In Weisler (1997b), chapter 9, pages 149–172.

Weisler, M. I., editor (1997b). Prehistoric Long-Distance Interaction in Oceania: An Interdisciplinary Approach. Number 21 in New Zealand Archaeological Association Monograph. New Zealand Archaeological Association, Auckland, NZ.

Weisler, M. I. (1998). Hard evidence for prehistoric interaction in Polynesia. Current Anthropology, 39:521–532.

Weisler, M. I. (2002). Centrality and the collapse of long-distance voyaging in East Polynesia. In Glasscock, M. D., editor, Geochemical Evidence for Long-Distance Exchange, pages 257–273. Bergin and Garvey, Westport, CT.

Weisler, M. I., Collerson, K. D., Feng, Y.-X., Zhao, J.-X., and Yu, K.-F. (2005). Thorium-230 coral chronology of a late prehistoric Hawaiian chiefdom. Journal of Archaeological Science, 33(2):273–282.

Weisler, M. I. and Kirch, P. V. (1985). The structure of settlement space in a Polynesian chiefdom: Kawela, Molokai, Hawaiian Islands. New Zealand Journal of Archaeology, 7:129–158. 92 Robert J. Hommon

Whitman, J. B. (1979). An Account of the Sandwich Islands: The Hawai- ian Journal of John B. Whitman. Topgallant Publishing, Honolulu. Edited by John Dominis Holt.

Williams, S. S. (1992). Early inland settlement expansion and the effect of geomorphological change on the archaeological record in K¯ane‘ohe, O‘ahu. New Zealand Journal of Archaeology, 14:67–78.

Yawata, I. and Sinoto, Y. H., editors (1968). Prehistoric Culture in Oceania: A Symposium. Bishop Museum Press, Honolulu.

Yen, D. E. (1973). The origins of Oceanic agriculture. Archaeology and Physical Anthropology in Oceania, 8:68–85.

Abstract

This paper recommends that archaeological projects in Hawaii incorporate procedures to test the Limited Land Hypothesis, which states: By ad 1550 ancient Hawaiian agricultural expansion slowed significantly as it approached effective limits on staple production imposed by available technology, sociopolitical factors, and natural variables including rainfall and soil fertility. The hypothesized reduction in the rate of agricultural expansion by at least 50 percent no later than ad 1550 as described in Part 1 of the paper was a watershed event in ancient Hawaiian history. In Part 2 the robustness of the hypothesis is supported by a wide array of archaeological and ethnohistorical evidence indicating environmentally- determined agricultural boundaries, extensive cultivation of marginal lands using flood-water irrigation and other techniques, traditional knowledge of a variety of famine foods, a possible reduction in demand for adzes that may signal a reduced pace of land-clearing, traditional methods of population control that may have contributed to an end to rapid population growth, the development of conquest warfare, and the emergence of competing archaic states. Tests of the Limited Land Hypothesis, aided by the procedural outline presented here in Part 3, are expected to shed light on the major indigenous transformation of Hawaiian culture of the last few centuries before Western contact in 1778. 2

Traditional Hawaiian Surface Architecture

Absolute and Relative Dating

Thomas S. Dye T. S. Dye & Colleagues, Archaeologists

Introduction

Traditional Hawaiians were accomplished dry masons and inveter- ate builders. One result of the masons’ labors is an archaeological landscape that is, in many places, dominated today by surface ar- chitecture. Abandoned Hawaiian settlements are often immediately visible, where they are not overgrown, and typically consist of a vari- ety of structures—walls, terraces, platforms, enclosures, pavements, alignments, mounds, paths, pits, ditches, and/or clearings. Identifying and describing the structures in abandoned settlements is a preoccu- pation of Hawaiian archaeologists that began more than a century ago when the first professional archaeologist in Hawai‘i, Bishop Museum’s John F. G. Stokes, drew maps, wrote descriptions, and recorded oral historical information about the prominent structures (Stokes 1991). This same type of work continues today with a broadened focus that now includes smaller, more informally constructed structures in an attempt to record the entire pattern of settlement in a region. This idea of settlement patterns is one of the inspirations for Hawai‘i’s

93 94 Thomas S. Dye historic preservation program and its emphasis on archaeological in- ventory surveys, one goal of which is a complete record of the surface structures in the archaeological landscape. Adjunct to this descriptive effort have been investigations of the relative or absolute ages of architectural features, but these are less frequent and, as will be argued below, often less successful. Conse- quently, progress in understanding settlement pattern change lags descriptions of surface features and their distributions. In fact, the situation in Hawai‘i today is only slightly improved from its state a quarter century ago when Roger Green, in a review of settlement pattern studies in Oceania, noted that “diachronic relationships and changes in settlement pattern over time are as yet not well estab- lished” (Green 1984, 65). Statements such as the conclusion, quoted below, from archaeological investigations for the H-3 freeway in North H¯alawa Valley are all too rare.

Stone architectural construction at sites appears to be late in the sequence of use of many locales . . . habitation and other activities often occurred at sites prior to the con- struction of terraces, enclosures, and mounds. In general, enclosures and mounds tend to be some of the latest stone architectural features at any given locale. (Hartzell et al. 2004, 300)

One reason for the paucity of similar statements is that absolute dating methods are often applied uncritically by archaeologists work- ing in Hawai‘i. This is somewhat surprising, given that Hawai‘i was one of the last places on the earth to be settled by humans, and its short archaeological sequence makes rigor in dating analysis important. It is sobering to consider the fact that five decades of archaeological 14C dating, from one of the early 14C dates processed by Willard Libby on charcoal from a shelter on O‘ahu Island (Emory et al. 1968) through to the end of the twentieth century, left archaeologists debating whether initial settlement of the islands took place in the first or fifth centuries ad (Hunt and Holsen 1991; Kirch 1985). The dates being argued were at least 400 years earlier than the late eighth or early ninth century date for initial settlement now convincingly established by paleoenvironmental investigations (Athens et al. 2002), an error that overestimated the length of the archaeological sequence by half. But the paucity of statements about diachronic settlement in the Hawaiian archaeological literature is also due to insufficient Traditional Hawaiian Surface Architecture 95 development of dating methods and techniques suited to the investi- gation of traditional Hawaiian structures. Clearly, there is a need for a programmatic statement on dating Hawaiian archaeological sites.

This chapter develops the conceptual and methodological frame- works needed to investigate the relative and absolute dating of tradi- tional Hawaiian surface architectural features. Its goal is to provide a general research design for the investigation of the diachronic set- tlement pattern in all its aspects: the structural plans devised by traditional architects, kahuna kuhikuhipu‘uone; the handiwork of masons; the development of individual settlements; and the larger pattern of settlement in its land-use context. Its main line of argument is that this is something Hawaiian archaeologists can achieve with only a modest field effort, well within the range of effort typically expended during a routine inventory survey. The two key changes in practice that it prescribes—a holistic, systematic approach to stratig- raphy and a model-based approach to chronological analysis—are both situated in precisely formulated, well established, and actively developing lines of archaeological research. In the realm of stratig- raphy, the Harris Matrix (Harris 1989; Harris et al. 1993) provides a clear and easily-applied framework for integrating the typical study of stratigraphy in Hawaiian archaeology, which focuses primarily on the sediments at a site, with the construction sequences of architectural features. In the realm of chronological analysis, modern software packages, such as BCal (Buck et al. 1999) and OxCal (Bronk Ramsey 2001, 2007; Ramsey 1995), provide user-friendly access to sophisti- cated and powerful analytic routines that demonstrably increase the yield of chronological information from archaeological investigations. A series of worked examples taken from archaeological investigations on Maui and Hawai‘i Islands (fig. 2.1) illustrate the application of these stratigraphic and chronologic analyses, and provide the data for a hypothesis on the diachronic settlement pattern of traditional Hawai‘i. The hypothesis is framed in such a way that it can be routinely tested under a wide variety of field conditions. Finally, an argument is made that testing and refinement of the hypothesis will have immediate implications for broader questions of traditional Hawaiian history, such as the history of population. 96 Thomas S. Dye

Figure 2.1 Map of the Hawaiian Islands, showing places mentioned in the text. The UTM Zone 4N coordinates use the NAD 1983 datum.

Relative Dating of Construction Sequences

Relative dating techniques have a long history in Hawaiian archaeol- ogy. Introduced to Hawaiian archaeology during the Pacific Science Congress at Honolulu in 1921, relative dating by “stratigraphic super- position” was proposed by anthropologists as a scientific approach to studying change over time in architectural forms. This proposal was a response to the work of Stokes and Bishop Museum’s first director, William T. Brigham, who believed, on the basis of traditional historical accounts, but without stratigraphic evidence, that the architectural style of religious temples, or heiau, had changed over time. In their reading of the traditions, platform heiau were remnants of an original Hawaiian religion that welcomed participation in religious ceremony. The platform heiau were supplanted by walled heiau after a migration to Hawai‘i of priests and chiefs from Kahiki that “hardened” the re- ligion and instituted the practice of carrying out religious ceremony out of the site of common people (Dye 1991). One index of how seri- Traditional Hawaiian Surface Architecture 97 ous archaeologists in the 1920s and 1930s were about investigating relative chronology is their development and use of the oblique site schematic that, in their work, augmented or replaced detailed site plans and profiles of the type drawn by Stokes (1991). In contrast with Stokes’s detailed site plans and profiles, which are early examples of a type of site record that is popular among archaeologists today,1 the oblique schematics developed by archaeologists in the 1920s and 1930s potentially isolate architectural components of structures and often display their relative stratigraphic positions as well, bringing this analytic information to the fore (fig. 2.2). Unlike the plans drawn by Stokes, which attempt to illustrate the rock sizes used as con- struction material and which draw everything in a uniform scale, the oblique schematics typically make no attempt to illustrate the rocks used in construction and sacrifice a uniform scale for an illustrative perspective; the focus is fixed on the architectural components, their forms, and positions relative to one another. For various reasons, primary among them that the archaeologists of the 1920s and 1930s didn’t find examples of the stratigraphic superposition for which they were looking, little came of this work and it wasn’t until the late 1960s that relative dating was taken up by Hawaiian archaeologists again, this time in the context of investigating architectural modifications to a structure or group of structures over time. This renewed focus perhaps shows most clearly in the work of Edmund J. Ladd, a Na- tional Park Service archaeologist from the American Southwest, who documented a construction sequence at Kaneaki Heiau in M¯akaha Valley, O‘ahu (Ladd 1973). Excavations, focused on the heiau because this is a type of structure known traditionally to have been modified and reconstructed over its life history (Malo 1996, 241). At Kaneaki, Ladd drew oblique schematics, not to scale, to illustrate the changing configurations of the heiau through six stages of construction. These oblique schematics make full use of the relative chronological infor- mation exposed during excavations, which yielded many examples of stratigraphic superposition. Again, no attempt was made to illus- trate the rocks or to render the structure in a realistic fashion; the sole focus was on the forms of components and their stratigraphic relations. Soon afterward, Patrick Kirch, then at Bishop Museum, inves- tigated the construction sequence of a men’s house in the residential

1Examples of Stokes-style plans drawn recently are presented in figure 2.9 on page 121, figure 2.11 on page 124, figure 2.12 on page 125, figure 2.13 on page 127, figure 2.14 on page 128, and figure 2.16 on page 134. 98 Thomas S. Dye compound of a famous mid-seventeenth century Hawai‘i Island chief (Cordy 2000), Lonoikamakahiki, at Kahalu‘u, Hawai‘i (Kirch 1973). This work proved that structures other than heiau were rebuilt over time. In addition, Kirch illustrated three of the six building phases using isometric projection, which was an advance over the earlier use of oblique schematic drawings because isometric projection equally fore- shortens scale along each of the three , thus preserving metric data lost in an oblique projection; Kirch’s figure shows a scale for the x, y, and z dimensions instead of the human figure which was often placed on the oblique schematic maps to give some sense of a structure’s size. By the early 1970s, then, field methods and report- ing standards for routine investigation of the relative chronology of architectural features were fully worked out.

Figure 2.2 Oblique schematic of Kaneaki Heiau, M¯akaha, O‘ahu. Note that the walls on the upper platform were built on the platform’s surface. The scale is approximate; scale on the original drawing was indicated by a human figure standing on one of the platforms. After McAllister (1933, fig. 39).

Alongside this development of displays that focus on stratigraphic information is the elaboration of the Stokes-style plan and profile map, with considerable effort devoted to illustrating the rocks used in construction. One goal of this type of map is to record existing conditions in as much detail and with as much precision as possible, using a single scale that is constant among all the map elements. This type of map is especially useful in preservation situations, where changes over time in the condition of a structure might indicate the need for repairs or for changes in accessibility or other details of how the structure is being preserved. These maps are also useful during reconstruction projects where masons require building plans that Traditional Hawaiian Surface Architecture 99 clearly show the lines of a structure and the heights to which walls and platforms are to be taken. This type of map is less useful for stratigraphic study because the stratigraphic position of architectural components is often difficult to illustrate in plan view, and profiles typically illustrate the heights of different parts of a structure without isolating individual components. In the extreme case, an attempt is made not to analyze a structure and call out its architectural com- ponents in a map, but instead to give equal illustrative weight to all rocks, regardless of whether they are in their original positions within the structure or have tumbled out. It might be argued that these maps are in some way scientifically objective, because they shun interpretation in favor of recording facts on the ground, but the fact is that maps drawn in this way are often of little use in reconstructing either the size and form of the original structure or for illustrating the construction sequence. Map-making is, by its nature, an interpretive activity that isn’t well-served by misplaced claims of scientific objectiv- ity. Clearly, a balance must be struck between accurately recording the ruins of a long-abandoned structure and documenting the information needed to reconstruct the architectural components of the structure as they stood when the structure was in use. This is especially so in a situation common in Hawai‘i today where the archaeologist is tasked with recording structures that will be destroyed to make room for development. In these situations, where there is no consideration of preservation or reconstruction, one might expect to find that archae- ologists typically map structures in a way that most fully illustrates their configurations while in use and that calls attention to details of construction history. The logical choice here, if only one kind of map is produced, would be an oblique view, either schematic or isometric, where individual architectural components and their stratigraphic relations are emphasized. Such maps are relatively rare, however, and the default map style—virtually the only type used in cultural re- sources management reports—is instead very similar to the plans and sections drawn by Stokes in the early 1900s, with little or no attempt to isolate architectural components or illustrate their stratigraphic relations. This choice of a Stokes-style plan in situations where an oblique view would better serve is an indication of the subsidiary role played by relative chronological investigations in much Hawaiian archaeology as practiced today. One possible reason that relative chronological investigations are not better represented in the mainstream of Hawaiian archaeology today is the advent of 14C dating in the early 1950s and its increasing 100 Thomas S. Dye application to Hawaiian archaeology (Dye 1989). The seduction of the 14C method is its promise to establish absolute dates of archaeolog- ical events, a clear advance over relative chronological information. Archaeologists in Hawai‘i have expended considerable effort in the pursuit of absolute chronological information, most of it through the 14C method but also including a short-lived and ill-fated attempt to develop a reliable method to date volcanic glass (Olson 1983), and more recently the very precise dating of coral pieces by measuring the relative concentrations of the radioactive elements 238U and 230Th (Kirch and Sharp 2005; Weisler et al. 2005). However, Dye (2004) has argued that archaeologists in Hawai‘i are poorly prepared to carry out the analytic work needed to obtain absolute construction dates for structures because they often conflate the definition of surface architecture with the related, but separate, definition of site. This con- fusion has led them to downplay or ignore the importance of context in dating analyses. A site is typically defined in archaeology as a loca- tion at which evidence of past human activity is preserved; one type of evidence of past human activity, among others, is surface architecture. In Hawai‘i, however, the ubiquity of surface architecture and the great effort spent identifying and describing abandoned settlements has led archaeologists to define sites not as locations, but as surface ar- chitectural features. This semantic shift lends an undeserved primacy to surface architecture and changes the way archaeologists view the relationship between architectural features and other evidence of past human activity at a site. Because archaeologists mistakenly consider a surface architectural feature to be a site, it is no longer viewed, as it should be, as an attribute of a site. Instead, this relation is turned on its head. Other evidence of past human activity at a site is viewed as an attribute of the surface architecture; investigation of that other evidence becomes, in the mind of the archaeologist, an extension of the main project—the identification and description of surface architecture. Because of this, archaeologists in Hawaii, particularly in the realm of cultural resources management, often fail to investigate the stratigraphic position of surface architecture in the cultural de- posits of a site, a necessary step in estimating the construction date of a structure. Thus, the promise of absolute 14C dates for surface architectural features, which helped lead archaeologists away from investigating relative chronological relations, has not been fulfilled. Traditional Hawaiian Surface Architecture 101

Architectural Components

The relative dating of construction sequences requires an analytic framework that breaks surface architectural remains into components that can be designated with a limited set of component descriptors. There are many ways to do this; the approach advocated here is similar to one worked out by Weisler and Kirch (1985) at Kawela and Makakupa‘ia Iki Ahupua‘a, Moloka‘i in the 1980s and, in its current form,2 was inspired by the Harris Matrix and its application to analyses of construction sequences of historic buildings (Harris 1989; Harris et al. 1993). A surface architectural feature consists of one or more components, either contiguous or, if spatially discrete, reasonably hypothesized to have functioned as an integral whole. Given the level of technological development in traditional Hawai‘i and the constraints imposed by raw materials available in the environ- ment, the number and variety of architectural components is limited. Lacking mortar or other means save gravity for holding together struc- tures constructed of terrigenous and marine-derived clay, silt, sand, gravel, pebbles, cobbles, and boulders, traditional Hawaiian builders were limited to piling materials in various forms, or removing mate- rials, to create the architectural features that archaeologists record today. It should be noted that natural features of the environment such as overhangs, , lava blisters and tubes, or massive boulders often attracted development and that architectural features were sometimes built in such a way as to augment or otherwise incorporate them. Strictly speaking, these natural features are not components of an architectural feature, but are more properly considered part of the environment in which the feature was constructed. The component descriptors of architectural features can be classi- fied as a paradigm in two dimensions: • The first dimension distinguishes between additive components, built by piling materials together, and subtractive components, built by removing materials. In practice, construction of a com- ponent involves taking materials from one place and putting them somewhere else, and so involves both subtraction and addition. In almost all cases, one or the other activity is clearly primary and the other secondary, but it is useful to recognize

2Earlier efforts along these lines at Bishop Museum in the 1980s benefited greatly from discussions with Alan Haun and Patrick McCoy, both of whom were construc- tively critical of the approach. 102 Thomas S. Dye

that there is scope for some judgment on the part of the archae- ologist using the classification and to be alive to the possibility that an archaeologist might mis-attribute primacy. • The second dimension specifies the ideal geometric form that, on a human scale, the feature most closely resembles: point, line, plane, cylinder, or polyhedron. The use of one- and two- dimensional ideal forms, in addition to the three-dimensional ideal forms that one might expect to use when describing three- dimensional objects, is intended to emphasize the primary di- mensions of components, those that were most directly relevant to their function, and to indicate that the other dimensions are best considered secondary attributes of the components. Here, it might be noted that the use of geometric forms is an analytic device that aims to provide some clear points of formal reference in an archaeological landscape in which ideal forms are rarely, if ever, realized. The paradigm thus defined includes twelve states, none of which appear to be logical impossibilities. These states are labeled with terms in common use among archaeologists in Hawai‘i (table 2.1), although no claim is made that the classification encompasses their range of application in the literature. The terms themselves are trivial; it is the classification that brings structure and utility to the descriptions.

Table 2.1 Architectural component descriptors

Geometric Representation Descriptor Primary Dimension(s)

One-dimensional Additive vertical line mound height Subtractive vertical line pit depth Additive horizontal line length Subtractive horizontal line path length Two-dimensional Additive horizontal plane pavement diameter or length and width Subtractive horizontal plane clearing diameter or length and width Additive vertical plane wall length and height Subtractive vertical plane trench length and depth Three-dimensional Additive cylinder berm length and cross-section diameter Subtractive cylinder ditch length and cross-section diameter Additive polyhedron foundation length, width, and height Subtractive polyhedron open hole length, width, and depth Traditional Hawaiian Surface Architecture 103

An Example from Hawai‘i Island

Traditional Hawaiian architectural features in the makai portion of Kaiholena Ahupua‘a have been studied in detail for many years. Fol- lowing limited archaeological survey in the 1960s (Bonk 1968; Soehren 1964, 1969), archaeological inventory survey of the Kaiholena project area was accomplished in three steps. In June 1979, Sinoto (1979) com- pleted an archaeological reconnaissance survey of five parcels of land in North Kohala, one of which was the makai portion of Kaiholena Ahupua‘a. Between August 13 and 28 of that year, Sinoto returned with co-director Rose Schilt and four assistants to complete “a lim- ited Phase I type archaeological survey” (Schilt and Sinoto 1980, 1) of four parcels, once again including the makai portion of Kaiholena Ahupua‘a. The inventory survey was completed twenty-three years later by Dye (2003). An archaeological data recovery project, includ- ing plane-table mapping, was subsequently completed at 29 historic properties (Major and Dye 2005). This work, carried out primarily by Maurice Major, interpreted the remains of surface architectural fea- tures in a dynamic framework, using a set of criteria concerned with feature integrity. The attributes used in the interpretation included: • presence of cultural deposits buried by surface architecture; • variable integrity of surface architectural features within 100 m of each other; • presence of lag deposits indicating paved floors and/or sta- ble occupations where surface architecture is either absent or unexpectedly minimal; • presence of well-set foundations with little or no above-ground stacking; and • presence of loose piles of stone between remnant and relatively intact architectural features. A preservation plan developed the next year called for detailed plane table mapping of architectural features in 15 historic preserves; the work reported here was completed during the preservation planning. First recorded as site 14–1–86 (Schilt and Sinoto 1980, 57), site 50–10–04–22268 is located in a coastal village just north of the large gulch near the center of the Kaiholena Ahupua‘a. It is a compound structure dominated by an open U-shape enclosure with tall, thick walls (fig. 2.3). Within this enclosure are interior walls that Schilt and Sinoto (1980) recognized as later additions to the structure. One of these walls abuts the interior face of the inland wall of the U- shape enclosure (fig. 2.4). The interior walls include numerous flat, 104 Thomas S. Dye water-worn boulders that resemble remnants of a paving within the U -shape enclosure; it is likely that much of the material used to build these recent interior walls was taken from components of the U -shape enclosure and was not brought to the site from a distance.

Figure 2.3 Plan of site 50–10–04–22268. The coast is a short distance west of the site. Original map drawn with a plane table and alidade at a scale of 1:100.

Adjoining the U-shape enclosure on its seaward side are a series of low wall sections and alignments that mark out three, possibly four, Traditional Hawaiian Surface Architecture 105

Figure 2.4 Interior of the U-shape enclosure at site 50–10–04–22268, looking north. The wall in the foreground abuts, and is later than, the inland wall of the U-shape enclosure at the right hand side of the photograph. Note also the large number of flat water-worn stones used in construction of the more recent wall. The scale is marked in decimeters. rectangular enclosures (fig. 2.5). These enclosures are oriented about 15 degrees off the main axis of the U-shape enclosure. One section of wall at the north end is faced on the interior and exterior; it is the only wall section in this part of the site that is faced on both sides. Two other sections of wall have an exterior face, but lack one on the interior. At the southwest corner of the site, six upright boulders are found in the alignment that marks out the large, nearly square enclosure. Upright boulders are commonly found set as part of the basal course of walls in this portion of coastal Kohala; one upright in the U-shape enclosure has collapsed from the interior of the inland wall near its middle. This evidence indicates at least three periods of construction and use at the site. Initially, a compound of three or possibly four walled enclosures was established at the makai end of the site. The walls of the southern enclosure in the compound had a basal course marked by several upright boulders. Judging from the width of the last remnant of bi-faced wall, the north, south, and makai walls of these enclosures were not particularly massive. This is a common pattern along this portion of the Kohala coastline, where the inland walls are typically high and thick to break the strong winds that famously 106 Thomas S. Dye

Figure 2.5 Panoramic view of site 50–10–04–22268, looking south. Note the vestigial walls, in many cases reduced to alignments, in the foreground. They contrast with the substantial walls of the long enclosure behind them. The scale is marked in decimeters. sweep down from the mountains here. Nothing appears to remain of these inland walls and it is likely that their stones were taken, along with stones from the upper courses of the other walls, to build the U - shape enclosure. The shape of this structure, its open end, and length sufficient to house a canoe suggest that it was used as a halau, or canoe shed, despite the fact that it is located some distance from the coast and is somewhat wider than is typical for this type of structure. During the final phase of construction and use, the structure could no longer be used to store a canoe. Its interior was partially dismantled to construct walls that divide the interior in an awkward way, yielding two small rooms and a narrow hallway leading to the back end of the original enclosure. Several questions could not be answered from an inspection of the surface remains alone. The exterior face of the U-shape enclosure’s seaward wall is now collapsed and this has obscured stratigraphic evidence for the sequence of construction here. Based on the rela- tive integrity of the architectural components, the U -shape enclosure seems younger than the adjoining enclosure, but this could not be con- firmed by observing the contact between the two. Again, the contact between the south wall of the enclosure and the middle of the west wall of the U -shape enclosure did not yield any stratigraphic evidence at the surface because the one wall is just a small remnant here and Traditional Hawaiian Surface Architecture 107 the other is largely collapsed along its face. Removing rocks fallen from the U -shape enclosure, perhaps augmented by some excavation, might reveal the relevant architectural stratigraphy. Excavation might also reveal the extent to which the boulder paving of the U-shape enclosure was torn up to construct the interior walls. Finally, absolute dates from contexts beneath the walls of the various enclosures might indicate their relative ages, an approach that is set out in detail below.

Absolute Dating of Construction Events

The Stratigraphic Position of Surface Architecture

Hawaiian archaeologists’ failure to distinguish sites from surface ar- chitecture has contributed to a general neglect of issues surrounding the analysis and recording of stratigraphy. The state’s archaeology rules only require that sediments be described according to conven- tions developed by soil scientists and promoted today by the U.S. Department of Agriculture. In practice, this usually means that soil colors are described with terms from the Munsell chart, that soil de- scriptions include reference to the code that designates hue, chroma, and value, and that soil texture be determined, generally by feel (Thien 1979). At times, the topography and distinctness of the boundary be- tween two deposits is also described systematically, as are stickiness and plasticity. Very rarely, other soil properties, such as the size and shape of peds or the abundance of large clasts, are also described according to the agricultural standards. There is little recognition among practicing archaeologists that most of this description misses what should be the primary goal of stratigraphic excavation of archae- ological sites—the identification and temporal ordering of cultural events and processes.3 The archaeologists left with the unenviable task of writing up the extensive excavations carried out during data recovery for the H-3 Freeway in upper North H¯alawa Valley found that reconstructing stratigraphic sequences from field notes and lab- oratory analyses based on the agricultural standards was “unduly

3The stratigraphic practice of many Hawaiian archaeologists stakes out an ex- treme position on the geologist’s side of a once acrimonious debate about the rela- tionship between archaeological and geological stratigraphy (e.g. Harris 1989; Stein and Farrand 1985). Although the approach outlined here takes the opposite position, there is no real sense in getting caught up in the debate; what is important is that the approach developed for the Harris Matrix provides the conceptual framework needed to formulate the prior stratigraphic information used in the Bayesian calibration of dated samples. 108 Thomas S. Dye time-consuming” (Hartzell et al. 2004, 302). They were able to work through sequences at three important sites, but were “unable to as- sess all 75 sites using this technique” (Hartzell et al. 2004, 302). This experience led them to recommend that future work develop “a mas- ter correlation matrix for the site, such as the Harris matrix . . . while work is ongoing in the field” (Hartzell et al. 2004, 302). Without this correlation matrix and the framework of events and processes that it yields, there can be little hope of accurately estimat- ing the construction date of an architectural feature. This is because construction of a traditional Hawaiian architectural feature typically involves moving volcanic rocks of various sizes from one place to another, a procedure that creates nothing that can be dated directly. Instead, a construction event must be dated archaeologically by ref- erence to events and processes of the stratigraphic record that can be shown to have taken place either before the construction event or after it. One set of techniques for accomplishing this level of strati- graphic analysis, known today as the Harris Matrix (Harris 1989), was worked out in the 1970s by archaeologists involved in the excavation of complex urban sites in Great Britain. The Harris Matrix is based on four axioms of archaeological stratig- raphy, termed laws:

Law of Superposition In a series of layers and interfacial features, as originally created, the upper units of stratification are younger and the older are lower, for each must have been deposited on, or created by the removal of, a pre-existing mass of archaeologi- cal stratification (Harris 1989, 30);

Law of Original Horizontality Any archaeological layer deposited in an unconsolidated form will tend towards a horizontal posi- tion. Strata which are found with tilted surfaces were originally deposited that way, or lie in conformity with contours of a pre-existing basin of deposition (Harris 1989, 31);

Law of Original Continuity Any archaeological deposit, as originally laid down, or any interfacial feature, as originally created, will be bounded by a basin of deposition,4 or may thin down to a feather edge. Therefore, if any edge of a deposit or interfacial feature is exposed in a vertical view, a part of its original extent must

4A basin of deposition is the “area which defines the pattern of deposition of layers, e.g. the shape of a cave, room or pit” (Harris 1989, 155). Traditional Hawaiian Surface Architecture 109

have been removed by excavation or erosion, and its continuity must be sought, or its absence explained (Harris 1989, 32); and

Stratigraphic Succession A unit of archaeological stratification takes its place in the stratigraphic sequence of a site from its position between the undermost (or earliest) of the units which lie above it and the uppermost (or latest) of all the units which lie below it and with which the unit has a physical contact, all other superpositional relationships being redundant (Harris 1989, 34).

Units of stratification recorded by the Harris Matrix include the natural and man-made layers routinely identified by Hawaiian ar- chaeologists (fig. 2.6 a, b) and five others that are typically ignored or considered non-stratigraphic: upstanding strata, such as walls or other similar man-made deposits (fig. 2.6 c); horizontal layer interfaces, which represent the surface of a layer upon which other material was deposited (fig. 2.6 d); vertical (or upstanding) layer interfaces, most typically the faces of walls (fig. 2.6 e); horizontal feature interfaces, most typically the tops of partially destroyed walls (fig. 2.6 f ); and vertical feature interfaces, the edges of holes that have been dug at a site (fig. 2.6 g). This scheme has obvious appeal for the surface architecture dat- ing project because its comprehensive catalog of stratification units encompasses not only the sediments that are the usual subject of stratigraphic description in Hawaiian archaeology but surface archi- tecture, as well. In the hypothetical example shown in figure 2.6, the walls of an enclosure, like the walls of many traditional Hawaiian structures, were built on a pre-existing cultural deposit. The broken line connecting the bases of the two walls indicates the ground surface upon which the walls were built. It is typically the case in Hawai‘i that this ground surface cannot be readily identified by examining the sediments themselves with standard archaeological techniques. Instead, the sediments deposited within the enclosure can be distin- guished stratigraphically from similar sediments laid down before the structure was built by their interface with an upstanding wall stratum; sediments with this interface were laid down after the enclosure was built. Investigating the stratigraphic position of surface architecture by exposing the full vertical interface of an upstanding wall stratum not only makes it possible to distinguish cultural materials and sed- iments deposited after a structure was built from those deposited before it was built, it potentially gives the archaeologist access to 110 Thomas S. Dye

Figure 2.6 Hypothetical stratigraphic section through an enclosure illustrating the units of stratification recorded in a Harris Matrix: a, natural layer; b, man-made layer or cultural deposit; c, upstanding stratum; d, horizontal layer interface; e, vertical layer interface; f, horizontal feature interface; and g, vertical feature interface. The horizontal broken line indicates the ground surface when the wall was constructed. the ideal context for determining the age of the surface architecture. Organic matter collected from the cultural layer below the broken line, preferably immediately beneath the portions of the basal stones that would have rested on the ground surface when the structure was built, is older than the wall construction event. When this material is dated, it yields information that can be used in a routine Bayesian analysis to estimate the date of wall construction.

A Bayesian Approach

The 14C dating method provides archaeologists with a potentially powerful means to determine the ages of events that took place in the distant past. Based on the well-established rate of decay for the radioactive isotope of carbon, 14C, the method can be applied to preserved or charred remains of animals or plants. Typically in archaeology, a piece of charcoal or marine shell recovered from a context that can be associated with an event of interest is sent to a laboratory that measures the ratio of 14C to the stable carbon isotope, 12C. This ratio is high in young samples and low in older Traditional Hawaiian Surface Architecture 111 samples where the 14C has had more time to decay. The laboratory converts this ratio to a conventional 14C age (Stuiver and Polach 1977) using an internationally accepted formula and returns this result to the archaeologist along with an estimate of the uncertainty of the measurement. The archaeologist must then calibrate the conventional 14C age to gain an estimate of the age of the sample in years. This estimate is best expressed as a range of years, rather than a single year, to take into account the uncertainties of the laboratory measurement and the calibration procedure. There are several options for calibration. A calibration that relies solely on the information returned by the laboratory yields an age estimate for when the dated material was alive and growing within an animal or plant. This is often useful information, but in many cases it doesn’t relate directly to the age of the archaeological event of interest. One of the most common reasons for differences in the ages of a dated material and the archaeological event with which it is associated is that the material was already old when it was deposited in an archaeological site (Bowman 1990; Taylor 1987). This is frequently the case with firewood when, for instance, the trunk of a several hundred year old tree is burned. A 14C date on a piece of old tree wood charcoal from a fire pit might yield an age estimate several hundred years older than the burning event. Archaeologists concerned with establishing the true age of an archaeological event control for the effects of old wood by dating samples from short- lived plants or animals, or by choosing a plant part, such as a twig or nutshell, that lived for a relatively short period. Alternatively in Hawai‘i, certain marine shells that live for a short period and take up their carbon from the general ocean , typically filter-feeding bivalves, can be used (Dye 1994). Another reason that a 14C date might not relate directly to the age of an archaeological event is that it comes from a different, but strati- graphically related, context. This situation commonly arises when the archaeological event of interest did not leave behind pieces of plants or animals suitable for dating with the 14C method. A common example of this type of event in traditional Hawai‘i is the construction of a stone structure where volcanic rocks are skillfully laid up to make a sturdy wall or platform, but plant and animal parts are not used. An archaeologist hoping to estimate the age of the structure might recover material older than the structure from the sediment beneath it, or, more rarely, material younger than the structure from sediment that buried it, but there is no material suitable for 14C dating that is 112 Thomas S. Dye directly associated with the construction event.5 In situations such as this, the archaeologist uses a Bayesian calibration procedure that integrates information about the relative ages of the 14C date and the event of interest, in addition to the conventional 14C age returned by the laboratory. The ability of Bayesian calibration to integrate chronological infor- mation of different types turns out to be a powerful feature (Buck et al. 1996). Perhaps the easiest way to make this point is with a thought experiment whose basic outline will be familiar to most practicing archaeologists. Consider a simple, two layer deposit that has been carefully excavated and determined to be free of intrusive features and other characteristics that would indicate the possibility that ma- terials from the two layers had been mixed. Suitable dating materials were collected from each of the layers and submitted to a reliable dating laboratory, which reported that the samples were processed normally and that nothing was amiss. The material from the upper layer, Layer I, returned a date of 800 ± 50 bp and that from the lower layer, Layer II, 750 ± 50 bp. When the two dates are calibrated using only the information returned by the dating laboratory, the sample from the Layer I dates to ad 1160–1284 at 2σ (fig. 2.7a) and that from the Layer II ad 1180–1304, also at 2σ (fig. 2.7b). How to interpret this stratigraphic inversion? One might note that the two calibrated ages overlap by more than a century and assign the Layer II sample to the first half of that century and the Layer I sample to the second half. Such a procedure seems sensible, given the stratigraphic situation, but there are other seemingly sensible, but equally ad hoc, procedures. It is not, in principle, possible to choose which among them is most likely to be true. In the end, the choice of one ad hoc procedure or another will be based on some predilection of the interpreting archaeologist, which might or might not be explicitly addressed in the report of the results. Another archaeologist coming to the same data with a different set of predilections will likely choose some other ad hoc procedure and arrive at a different result. There is nothing in the approach that will help decide whose answer is most likely correct. In contrast, the Bayesian approach starts with what is known about the relative ages of the two samples and then modifies this knowledge in the light of the 14C dating information. Calibrated with the BCal

5The advent of 238U/230Th dating of branch corals opens up the possibility of dating materials that are directly associated with the construction event. This possibility is discussed below on page 146. Traditional Hawaiian Surface Architecture 113

Figure 2.7 Bayesian calibration yields interpretable results: a, Layer I sample without stratigraphic information; b, Layer II sample without stratigraphic information; c, Layer I sample with stratigraphic information; d, Layer II sample with stratigraphic information; e, the Layer I/II boundary. software package (Buck et al. 1999), the two samples yield calibrated ages that agree with their stratigraphic positions. The sample from Layer I dates to ad 1205–1294 (fig. 2.7c), using the 95% highest pos- terior density region6 and that from Layer II ad 1160–1279 (fig. 2.7d), also using the 95% highest posterior density region. Comparing the two sets of date ranges, it can be seen that the largest differences are in the older range for the sample from the Layer I and the younger range for the sample for the Layer II. In other words, the Layer I

6The 95% highest posterior density region is the Bayesian equivalent of 2σ . 114 Thomas S. Dye sample yields a younger age estimate with Bayesian calibration and the Layer II sample an older age estimate. With Bayesian calibration, there is an older calendar year in the calibrated range of the Layer II sample for every calendar year in the calibrated range of the Layer I sample, and there is a younger calendar year in the calibrated range of the Layer I sample for every calendar year in the calibrated range of the Layer II sample. There is no longer a need to resort to ad hoc procedures to interpret the results in an archaeologically meaningful way. By taking into account the hard-won stratigraphic information collected in the field, the Bayesian calibration yields results that are immediately interpretable. Until fairly recently, the Bayesian approach was considered some- what outside the mainstream of statistical thought. Introductory statistics courses, of the sort likely to be studied by archaeologists, typically focus on classical statistics with only a brief mention of the Bayesian approach. This might have given archaeologists the impression that the approach is not based on a solid footing, or raised suspicions that it involves assumptions that can’t be justified in an ar- chaeological situation. For these reasons, the history of the approach and its logical basis are briefly reviewed here.7 The reader interested in the details of a Bayesian approach to archaeological problems is strongly encouraged to read Buck et al. (1996), which is a thorough and precise introduction to the subject. The Bayes in “Bayesian calibration” refers to the Reverend Thomas Bayes, whose posthumously published “Essay Toward Solving a Prob- lem in the Doctrine of Chances” contains the germ of the reasoning that now bears his name. Published in 1763, it appeared about a century after the word probability took on its modern meanings asso- ciated with i) the likelihood that an event will occur or ii) the degree to which we believe a statement to be true (Hacking 1975). Prior to this time, knowledge was divided into true knowledge, gained by demon- stration via deduction, and opinion; probability at that early time was an attribute of opinion, indicating something worthy of belief, and it often referred to an opinion approved as worthy by some authority. In his essay, Bayes refers to the problem he was attempting to solve as one of “inverse probability,” by which he meant the inverse of deduction, or what we know today as induction. Induction is a more complex form of reasoning than deduction (fig. 2.8). In it, one argues from effects or observations back to probable (in the modern sense)

7Much of the discussion that follows is based on Hacking (2001). Traditional Hawaiian Surface Architecture 115 causes, rather than from causes to necessary outcomes as in deduc- tion. In Bayes’ day, there was considerable skepticism over the use of induction. The Scottish philosopher David Hume argued in 1739 and 1748 (Hume 1956; Hume et al. 1980) that all induction ultimately rested on the uniformitarian assumption that the future would be like the past. Bayes’ essay didn’t prove the future would be like the past; Bayesians deny this is really necessary. Instead, they champion a form of reasoning that claims to be a reasonable way to learn from experience, one that scientists use routinely in their work (Howson and Urbach 1989).

Figure 2.8 Deduction, top, and induction, bottom. Modified after Sivia (1996, Fig. 1.1).

One way to express this learning process, or Bayes’ theorem, is that the product of one’s prior belief in a hypothesis and the likelihood of one’s observations yields an updated view of one’s belief in the hypothesis. This updated view is known as the posterior belief. In other words, Bayes’ Theorem can be used to update one’s belief in a hypothesis by assessing the likelihood of observing the data if the hypothesis were true. If one observes data that meet expectations, given the hypothesis, then the posterior belief will resemble the prior belief. If, however, the data are unexpected, then the posterior belief will be different from the prior belief to a corresponding degree. This result can be derived fairly easily from the quantitative rules necessary for logical and consistent reasoning (Cox 1946). These include a transitivity condition, a sum rule, and a product rule.

Transitivity condition Given 3 propositions, A, B and C, if we believe A more than B, and B more than C, then we must necessarily 116 Thomas S. Dye

believe A more than C. A convenient way to express this is to assign a real positive number to each proposition such that the larger the value, the greater the degree of belief. In modern probability theory these real numbers are in the range [0,1], where 0 indicates something certainly false and 1 indicates something certainly true. Typically, in problems using induction, values fall somewhere between 0 and 1. For instance, the belief that something has a 50/50 chance of being true would be assigned a probability of 0.5.

Sum rule An expression of belief that something is true, pr ob(X), is an implicit expression of the belief that it might be false, pr ob(X)¯ : pr ob(X) + pr ob(X)¯ = 1 (2.1)

Another way to state this is that we can be certain that a belief is either true or false, so the probabilities assigned them must sum to 1.

Product rule An expression of a belief that (proposition) Y is true, pr ob(Y ), and that X is true given that Y is true, pr obX|Y , is an implicit expression of our belief that both X and Y are true, pr ob(X, Y ):

pr ob(X, Y ) = pr ob(X|Y) × pr ob(Y ) (2.2)

For example, if we let proposition Y be the statement that “it is overcast” and state our belief that it is overcast half the time so that pr ob(Y ) = 0.5 and let X be the statement that “it is raining” and note that it rains half the time it is overcast, so that pr ob(X|Y) = 0.5, then the product rule tells us that the probability that it is raining and overcast is one in four, 0.5 × 0.5 = 0.25.

Bayes’ Theorem is a corollary of the sum and product rules and the transitivity condition. First, transpose X and Y in equation 2.2:

pr ob(Y , X) = pr ob(Y |X) × pr ob(X) (2.3)

Next, note that the pr ob(Y , X) = pr ob(X, Y ) so the right hand sides of equations 2.2 and 2.3 are equal:

pr ob(X|Y) × pr ob(Y ) = pr ob(Y |X) × pr ob(X) (2.4) Traditional Hawaiian Surface Architecture 117

Then, divide through by pr ob(Y ) to yield Bayes’ Theorem: pr ob(Y |X) × pr ob(X) pr ob(X|Y) = (2.5) pr ob(Y ) Equation 2.5 can be simplified by eliminating the denominator on the right hand side and replacing the equality sign with a sign of proportionality. The denominator is a normalization constant that is useful in some advanced situations, such as model selection, but need not concern us here. This yields:

pr ob(X|Y) ∝ pr ob(Y |X) × pr ob(X) (2.6)

The components of equation 2.6 are frequently referred to by Bayesians as, from left to right, the posterior, likelihood, and prior. Hence, the formulation of Bayes’ Theorem as the posterior is proportional to the product of the the prior and the likelihood. One drawback of such a general theorem is that the devil is in the implementation details. Exactly what is meant by each of the terms of the theorem varies from case to case. For many years, this meant that applying Bayes’ theorem required more or less sophisticated modeling that typically required a well-trained and apt mathematician to derive a set of complex formulae. Consequently, outside of some general cases and special topics of intense interest, Bayesian reasoning was seldom used. This situation changed radically with the spread of cheap computing power and development of a computer algorithm, the Gibbs sampler (Geman and Geman 1984), that through a proba- bilistic process applied many thousands of times, can yield results that closely approximate the exact results yielded by sophisticated mathematical models. This is the algorithm at the core of modern Bayesian calibration software packages, such as BCal (Buck et al. 1999) or OxCal (Bronk Ramsey 2001, 2007; Ramsey 1995), which can be used to model a very wide range of stratigraphic situations and prior beliefs. The difference this makes can be seen by comparing the two ap- proaches to solving the simple example of inverted dates discussed on page 112. The analytical approach requires two pages of mathematical formulae (Vincent 1988, 158–159), including this typically complex 14 equation for deriving the expected value of t1, one of the C dates:

 0 − 0 2  √σ1σ0 exp − (t2 t1) 2(σ 2+σ 2) ¯ 0 2πσ2 1 2 t1 = t1 − ( ) 0 0 t2−t1 I q 2 2 σ1 +σ2 118 Thomas S. Dye

Other situations with different priors would each require their own set of similarly complex equations tailored specifically to the task. In contrast, statistical software incorporating the Gibbs sampler can provide a uniform, relatively simple, interface for specifying a wide range of priors. The corresponding model for the example of inverted dates is easily specified with the BCal software package. First, a period is designated for each of the two layers, Layer 1 and Layer 2. A period represents the time interval over which the layer was deposited. The early and late bounds of each period, corresponding to the lower and upper boundaries of the layer, are conventionally identified as α and β, respectively, and associated with a layer by means of a subscript. Thus, the lower and upper bounds of the period associated with Layer 1 are identified as α1 and β1 and those for Layer 2 as α2 and β2. Conventionally, the unknown calendar date in which a dating sample was introduced to the deposit is designated as θ, and individual dates are distinguished by subscripts. In this simple example there is a single 14C date for each of the periods identified 8 as θ1 from Layer 1 and θ2 from Layer 2. The prior for this simple example can then be described as α2 > θ2 > β2 = α1 > θ1 > β1, where the > symbol means “is older than.” In simple words, this model indicates that the sample whose calendar date is labeled θ2 was introduced into the Layer 2 deposit during the time that Layer 2 was accumulating, and that the same is true for the sample whose calendar date is labeled θ1 and Layer 1. Further, the model specifies that Layer 2 is older than Layer 1 and that there is no hiatus between the two layers, as indicated in the model by β2 = α1. This prior information is collected by the BCal software in a natural way, through a series of dialogues that ask for the number of periods, the number of age determinations for each period, and the temporal relationship between the two periods. In the case of the temporal relationship between the two periods, the software presents a list of all the logical possibilities and the user selects one of these by clicking on a radio button. Another set of dialogues collects the information returned by the laboratory for each of the 14C dates. The entire data entry process for this simple example took less than five minutes. A typical run of the software requires 50,000–100,000 iterations of the Gibbs sampler,

8Note that in most archaeological situations there will be different numbers of 14C dates and stratigraphic layers so that there will be no correspondence between the subscripts for the θ and the layer designations. In any case, the labels associated with the bounds of the time periods and the 14C dates merely serve to identify them and do not affect the outcome of the calibration. Traditional Hawaiian Surface Architecture 119 followed by a statistical check that the Gibbs sampler has converged on a result; non-convergence indicates that the sampler might have failed to approximate the true value. In this case, the sampler did converge, yielding the archaeologically interpretable results for the inverted 14C dates given on page 113. The Bayesian calibration does more than this, though. It also yields an age estimate for the other elements of the model, in this case for the layer boundaries, α2, β2, α1, and β1. The most interesting of these are α2 and β1, which date the horizontal layer interface between Layers 1 and 2, a unit of stratification in the Harris Matrix. The age estimate for this interfacial feature is ad 1165–1289 (see fig. 2.7e, pg. 113), which fits neatly between the Bayesian age estimates for the two 14C dates and, like the age estimates for the inverted 14C dates, is immediately interpretable archaeologically. This ability to estimate the age of all model elements, whether or not they are associated directly with dating information, turns out to be a key feature in dating the construction of traditional Hawaiian surface architecture.

Examples of a Simple Bayesian Model

It is frequently the case in Hawaiian archaeology that organic materials are found beneath the basal stones of a surface architectural feature (Dixon et al. 2000). Where this is the case, a 14C date on short-lived material gathered from a stratigraphically secure position immediately beneath a basal stone can be used in a Bayesian calibration to yield an estimate of when the feature was constructed. This is something that can be accomplished routinely during inventory survey with a minimum of effort. Construction dates for three surface features at sites 50–10–04– 22119,–22201, and –22248 on the coast of Kaiholena Ahupua‘a in North Kohala district, Hawai‘i Island and two surface features at sites 50– 50–17–1088 and –1089 on the coast of ‘Alae Iki Ahupua‘a in K¯ıpahulu district, Maui Island have been estimated with material collected by a simple field procedure. The three Hawai‘i Island features and one of the Maui Island features were identified as habitations based on their modest sizes and enclosure forms. The other Maui Island feature is a larger, architecturally complex structure believed to be a fisherman’s shrine, or heiau. In each case, a small test pit was excavated against the face of an architectural component, the vertical layer interface was fully exposed so that the base of the component could be con- fidently identified, and suitable dating material was dug out of the 120 Thomas S. Dye face of the pit beneath the base of the component. In four of the five cases, the material was recovered directly beneath a basal stone of the structure. In the fifth case, the material came from a charcoal concentration beneath a pavement that was stratigraphically inferior to the surface feature. In each case, the excavator took whatever steps were necessary to be confident that the material recovered for dating was either in or on the ground when the architectural component was built and that it was not introduced to its stratigraphic position at some later time. The dated structure at site 50–10–04–22119 is located in the north half of the site (fig. 2.9). The structure shares its high eastern wall with an ill-defined enclosure to the south; another high wall to the north is now almost completely broken down and the living areas in its lee to the west can’t be identified on the surface. South of the high eastern wall are the partial remains of two enclosures presumably abandoned during traditional Hawaiian times when their rocks were taken to build new structures, such as the nearly square enclosure at the south end of the site that contains an internal platform that probably marks a human burial. The dated structure today is an open-ended enclosure much broken down on its western, seaward end, probably by the action of storm waves that have deposited calcareous marine sand throughout the interior (fig. 2.10). The south wall is broken down, especially in its mid-section. A long cupboard, or chamber interior to a wide wall or other suitable architectural component and believed to have been used for storage, was found partially collapsed in the north wall. Similar cupboards are a common feature of habitation structures in coastal Kaiholena. A small test pit was excavated with a folding shovel against the interior face of the eastern wall, next to a large boulder that appeared to be set deeply in the ground. The wall here has protected the sediment in its lee from erosion; most of the soil from the gentle slope inland of the structure is now gone. Excavation here revealed four layers. From the surface to a depth of 10 cm, Layer I is a dark brown loam. Layer II, from 10–15 cm below surface, is a dark brown very coarse, poorly sorted, sub-angular sand that represents a mixture of terrigenous Layer I sediment with calcareous beach sand. Both Layers I and II share an interface with the upstanding stratum of the wall, and were deposited after construction of the wall. Layer III, from 15 to 40 cm below surface, is a dark brown loam containing traditional Hawaiian artifacts, water-worn paving pebbles, and other cultural material. The wall was constructed on Layer III sediment. The large, Traditional Hawaiian Surface Architecture 121

Figure 2.9 Plan of site 50–10–04–22119 at coastal Kaiholena Ahupua‘a, North Kohala, Hawai‘i Island. The excavation is labeled “TU” on the map. The features labeled “cupboard” are constructed hollow spaces in the interior of the structure, some more than 1 m long, that were used for storage. Note that the coast is a short distance west of the site. Original plane table map drawn at a scale of 1:100 by E. Komori.

deeply-set boulder visible at the surface was found to be the second course of the original wall; it was resting at 28 cm below surface on a water-worn volcanic boulder laid flat. A trowel was used to excavate layer III sediment from beneath the flat boulder; a piece of cowry shell collected from this sediment was submitted for 14C dating (table 2.2). Layer IV, a dark reddish brown loam, is culturally-sterile subsoil. Site 50–10–04–22201 was ably described by Maurice Major during data recovery at the site. 122 Thomas S. Dye

Figure 2.10 Dated enclosure at site 50–10–04–22119, looking east. The scale against the enclosure wall, marked in decimeters, is placed at the location of the test pit. Photograph by E. Komori.

Table 2.2 14C dates from beneath surface architecture

Site Laboratory # Material δ13C CRA∗ 50–10–04–22119 Beta-142513 cowry shell +3.5h 500 ± 40 50–10–04–22248 Beta-142514 cowry shell +3.1 540 ± 30 50–10–04–22201 Beta-142516 Sida cf. fallax -23.3 120 ± 40 50–10–37–3817 Beta-235595 Chamaesyce sp. -11.3 590 ± 40 50–10–37–3817 Beta-237859 Sida cf. fallax -23.3 540 ± 40 50–50–17–1088 Beta-150615 Cocos nucifera∗∗ -23.9 750 ± 50 50–50–17–1089 Beta-150619 Cocos nucifera∗∗ -23.2 250 ± 40 50–50–17–1089 Beta-150618 cf. Metrosideros poly- -25.2 430 ± 40 morpha 50–50–17–1089 Beta-150617 Cordyline fruticosa -24.0 modern

∗ Conventional radiocarbon age (Stuiver and Polach 1977). ∗∗ Leaf midrib.

The basic elements of this site—an enclosure with an at- tached smaller enclosure, and alignments extending makai [seaward]—have been described previously, but thorough clearing and detailed plane table mapping at 1:100 scale re- veal additional details (fig. 2.11). The central enclosure has a large, relatively intact wall in the northeast corner, but Traditional Hawaiian Surface Architecture 123

the remainder is in very poor condition. What appeared initially to be a rough platform at the southern end of the mauka [inland] . . . wall looks like a section of wall that has collapsed down to the bottom course or two, with the sur- face of the remnant and the tumbled stone roughly level, but loose and not well set, possibly a goat or cattle path rather than a constructed platform. A row of set boulders, nearly parallel to the north wall, is either the foundation of a south wall or a retaining wall . . . The smaller structure abuts the exterior mauka [inland] wall of the larger enclo- sure, and though it appears to be an addition, the nature of the dry stacked stone renders the sequence equivocal. No south wall remains standing for this smaller structure, although some embedded stones about 3 m south of the north wall could be a foundation remnant; as it stands now, the feature is an L-shaped extension. A cupboard is built into the north wall and another cupboard accessible from this feature’s interior extends into the rear wall of the larger enclosure . . . The makai [seaward] edge of the larger enclosure may have been walled at one time, but if so it was thoroughly dismantled. A platform extending less than 2 m south of the northern wall appears to be a later addition, and is constructed of boulders with cobble in-fill creating a surface that is more solid and level than the aforementioned possible platform . . . (Major and Dye 2005)

A single test unit was excavated with a folding shovel to a max- imum depth of 28 cm below surface in the northeast corner of the central enclosure. Excavation revealed two layers. Layer I, a dark brown loam cultural deposit overlies the culturally sterile Layer II, a yellowish red loam. Layer I extended beneath the basal stones of the wall, where it was about 10 cm thick. A trowel was used to excavate into the sidewall of the test unit beneath the wall to collect suitable material for dating. This lateral excavation yielded 12 wood charcoal pieces that were identified as being from native trees and shrubs and a shrub introduced to the islands by Polynesians. A piece of Sida cf. fallax, a short-lived shrub known as ‘ilima, was submitted for 14C dating (table 2.2). Site 50–10–04–22248 consists of two enclosures whose northeast- ern walls both terminate at a common bedrock outcrop (fig. 2.12). It is 124 Thomas S. Dye

Figure 2.11 Plan of site 50–10–04–22201. The coast is a short distance west of the site. Original map drawn at a scale of 1:100 by M. Major.

located on the brow of a low hill directly overlooking the coast. The southern enclosure is the larger of the two. It is oriented with its long axis perpendicular to the slope, in common with most other habitation structures at Kaiholena, and is fronted on the west by a level court. A cupboard, now collapsed, runs the entire length of the northeast wall. A dense deposit of marine shells, along with traditional Hawaiian arti- facts such as volcanic rock and glass flakes, is visible on the surface inside the southern enclosure. The northern enclosure is smaller and nearly square. It likely represents an out-building for the habitation at the southern enclosure. A 1 m2 excavation unit was excavated with a trowel against the interior face of the larger enclosure’s west wall, revealing a simple two layer stratigraphic sequence. Layer I is a thin cultural deposit of dark grayish brown loam that overlies Layer II, a culturally-sterile yellowish Traditional Hawaiian Surface Architecture 125

Figure 2.12 Plan of site 50–10–04–22248. The coast is a short distance west of the site. Original plane-table map drawn at a scale of 1:100.

red loam. The base of the enclosure wall is in Layer I. A sample of cowry shell was collected from the cultural layer directly beneath the base of the wall by excavating into the face of the excavation unit with a trowel. The shell was submitted for 14C dating (table 2.2). Site 50–50–17–1089 was identified as site 50-Ma-A22-4 and de- scribed as a habitation by Soehren (1963), mapped as an open-ended enclosure by Rosendahl (1975, sheet 101), and most recently investi- 126 Thomas S. Dye gated by Dye et al. (2002, 35 ff.). It is a complex structure consisting of a narrow, rectangular enclosure with its long axis oriented northeast- southwest (fig. 2.13 a), a paved court on the southeast and southwest sides of the enclosure (fig. 2.13 b), and a small platform in the west corner (fig. 2.13 c). Three pieces of branch coral deposited on the small platform indicate that it was used as a domestic shrine. A small test unit was excavated at the interior face of the southeast wall of the enclosure. It revealed that the base of the wall was set 5–10 cm below the surface, which is built up somewhat by recent organic material and pebbles fallen from the wall fill. This surface deposit was designated Layer I. The wall was resting on the surface of Layer II, a dark gray gravelly silt rich in water-rounded pebbles, or ‘ili‘ili, typically used as a paving material. This layer, which was about 30 cm thick and loose, with abundant voids, yielded a variety of traditional Hawaiian cultural material. Beneath this was Layer III, a dark brown gravelly silt lacking ‘ili‘ili that extended to the base of excavation in culturally-sterile com- pact brown subsoil at 74 cm below surface. Layer III also contained traditional Hawaiian cultural materials. At its surface, the interface with Layer II, was a concentration of wood charcoal representing a range of native and Polynesian-introduced trees and shrubs, most of which are not present in the flora near the site today. No historically introduced taxa were present in the charcoal sample. Three pieces of identified charcoal from the concentration at the surface of Layer III were selected for dating, two of short-lived materials and one from a tree, ‘ohi‘a¯ lehua (M. polymorpha), known to live for a long time (Gerrish and Mueller-Dombois 1999). One of the short-lived samples, a piece of the k¯ı plant (C. fruticosa), returned a modern date (table 2.2). Although it is possible that a modern piece of k¯ı charcoal was intru- sive, having filtered in through the loose sediments of Layer II, the lack of historically-introduced taxa in the rest of sample suggests that this possibility is unlikely. Instead, it would appear that this is a relatively young sample that returned a statistically unlikely and unexpected result. The other two samples yielded results that met expectations. The coconut leaf midrib, a short-lived material dated as Beta-150619, yielded a relatively recent age, while the piece of ‘ohi‘a¯ lehua wood charcoal returned an older age, as might be expected from a tree that can live in excess of 200 years (Gerrish and Mueller-Dombois 1999). The structure at site 50–50–17–1088 was identified by Soehren (1963) as 50-Ma-A22-3 and described by him as a possible fishing shrine, mapped by Rosendahl (1975, sheet 93) as two platforms and an enclosure, and investigated most recently by Dye et al. (2002, 29 ff.). It Traditional Hawaiian Surface Architecture 127

Figure 2.13 Plan of site 50–50–17–1089: a, enclosure; b, paved court; c, platform. The coast is a short distance southeast of the site. Adapted from Dye et al. (2002, fig. 7).

is a complex architectural feature with three major and several minor components including a stepped platform (fig. 2.14 a), a paved terrace (fig. 2.14b), and an enclosure (fig. 2.14c). It is likely that the site is more complex architecturally than it appears today, but the evidence for this complexity is obscured by the deteriorated condition of the site and the fact that it appears to have been modified over time with little regard for the integrity of its original components. The platform has three and possibly four steps, which show best on the northeast side, where two nearly complete faces rise from the base of the platform 128 Thomas S. Dye and a third, incomplete face is at the top of the platform at a height of approximately 1.4 m. The small extension on the southwest end appears to be a recent addition built with boulders taken from the platform. The ground around the structure is irregularly paved to an undetermined extent with water-worn boulders. The paved terrace is offset 1–2 m northwest from the plan of the platform, but on the basis of surface evidence appears to be coeval with it. There are two additions to the terrace on the northwest, the later of which appears to post-date the enclosure wall, which was built after the paved terrace and is probably a historic-period addition.

Figure 2.14 Plan of site 50–50–17–1088 at ‘Alae Iki, Maui. The coast is a short distance southeast of the site. Adapted from Dye et al. (2002, fig. 6).

A single test unit was excavated with a shovel against the southeast face of the platform. It revealed the base of the platform at about Traditional Hawaiian Surface Architecture 129

15 cm below surface, resting on a cultural deposit containing charcoal, marine shell, volcanic glass, and one small piece of modern glass. A thin deposit of ‘ili‘ili pebbles at about 10–15 cm below surface in portions of the test unit could not be associated stratigraphically with the platform. A bulk sample collected directly beneath the stratigraphically lowest point of the platform at a depth of 15–20 cm below surface yielded charcoal from the native trees ‘ohi‘a¯ lehua (M. polymorpha), koa (Acacia koa), and loulu (Pritchardia sp.), the native shrub ko‘oko‘olau (Bidens sp.) and the Polynesian introduced trees niu (C. nucifera) and kukui (Aleurites moluccana). No historically introduced taxa were present. A piece of short-lived coconut leaf midrib charcoal returned the earliest age among the dated structures (table 2.2), several hundred years earlier than even the old ‘ohi‘a¯ lehua wood charcoal from site 50–50–17–1089. Each of the 14C dates from beneath the three Kaiholena and two ‘Alae Iki structures was calibrated using a simple model that estab- lishes a single group with early and late boundaries representing the cultural deposit beneath the structure. Models for four of the dates include two floating parameters, indicated in the model by φ, one to represent the construction event and the other, based on historical information, to establish a terminus ante quem for the construction event. This simple model can be expressed with the following formula:

αdep ≥ θ1 ≥ βdep > φconst > φhist (2.7)

where αdep and βdep are the older and younger boundaries of the cultural deposit beneath the structure, θ1 is the calendar date when the dated material was introduced to the deposit, φconst is the date of the construction event, φhist is the historical information used as a terminus ante quem, and > can be read as “is older than.” It was not possible to establish a terminus ante quem for the habi- tation site 50–50–17–1089 at ‘Alae Iki. The last term of inequality 2.7 is omitted, yielding the following inequality:

αdep ≥ θ1 ≥ βdep > φconst (2.8)

At Kaiholena, the single dated sample from beneath each of the structures was used to estimate θ1, and φhist was taken to be ad 1823, when the Reverend William Ellis toured Hawai‘i Island and noted that the coastal settlements south of Lapakahi, an area that includes Kaiholena, were abandoned (Ellis 1963). At ‘Alae Iki, θ1 for the heiau site 50–50–17–1088 was estimated with the single dated sample from 130 Thomas S. Dye

beneath that structure, and φhist was set at ad 1897, the date the land parcel in which the heiau is located was transferred to Kipahulu Sugar Company and sugarcane fields were established. At the ‘Alae Iki habitation structure, site 50–50–17–1089, the 14C age of the coconut leaf midrib was used to estimate θ1 because it was likely that the 14C age of the ‘ohi‘a¯ lehua wood charcoal reflected the effects of in- built age. The three wood charcoal dates were calibrated using an atmospheric curve (Reimer et al. 2004). The two dates on cowry shell were calibrated using a marine curve (Hughen et al. 2004) and a ∆R of 110 ± 80 (Dye 1994). The estimated ages of the construction events (fig. 2.15) reveal some interesting patterns. • The constraining effects of the historical termini ante quem can be seen in the right hand sides of the posterior probability distributions for the four structures that had this information (fig. 2.15a, c, d and e). All of these distributions end in the nineteenth century in contrast to the one structure that lacked such information (fig. 2.15b) and that shows a high probability of having been built in the decades leading up to ad 1950. • The habitation sites were all constructed relatively late. The 95% highest posterior density regions for the Kaiholena habitation structures are: ad 1725–1824 for the structure at site 50–10–04– 22119; ad 1720–1824 for the structure at site 50–10–04–22201; and ad 1720–1824 for the structure at site 50–10–04–22248. All of these structures were likely built in the eighteenth or nineteenth centuries, near the end of the traditional Hawaiian era. • The failure to collect a suitable dating sample from directly beneath the ‘Alae Iki habitation structure, site 50–50–17–1089, and the decision to date a sample from beneath an architectural component stratigraphically inferior to it, yielded less satisfying results than the other habitation structures. The 95% highest posterior density region for the pavement beneath the ‘Alae Iki habitation, site 50–50–17–1089, is ad 1600–1949. The goal during sample collection in a situation like this is to find the youngest datable material that is older than the structure. Al- though the dated sample from this site corresponded to the relations specified in the Bayesian model (equation 2.8), its strati- graphic position well below the base of the structure suggests that it was not the youngest sample that could have been col- lected. The effects can be seen in the long and heavy left-hand Traditional Hawaiian Surface Architecture 131

tail of the posterior probability distribution (fig. 2.15b), which shows relatively likely probabilities back into the sixteenth cen- tury. Still, the probability that the pavement was laid down sometime in the eighteenth century or later, as estimated by the Bayesian calibration, is 0.78. The enclosure present at the surface today was constructed after the pavement, so it is rela- tively certain that all four habitation structures were built in the eighteenth or nineteenth centuries. • The ‘Alae Iki heiau structure at site 50–50–17–1088 might be much older than the habitation structures. The 95% highest posterior density region for the heiau construction event spans the five century period ad 1380–1894 (fig. 2.15a). The wide range of this estimate is due to the lack of information that would provide a strong terminus ante quem for the construction event, such as might be provided by wood charcoal from an internal fire-pit that dates use of the structure. Still, even lacking a strong terminus ante quem, the probability that the heiau platform was built before the eighteenth century is 0.52, indicating that it is most likely older than the four habitation structures. The presence of loulu (Pritchardia sp.) wood charcoal supports this interpretation; pollen records indicate that the presence of this palm in lowland forests was declining precipitously by the fifteenth century, a process that demonstrably led to its local extirpation in other areas (Athens et al. 2002).

A More Complex Bayesian Model

The examples of the previous section were all relatively simple. A single parameter of interest—the construction date of some surface ar- chitectural component—was estimated with one 14C date on a suitable material recovered from a context below the base of the component. In most instances, this information was supplemented with historical information that specified a date after which the component was unlikely to have been built. In this section, the principles of Bayesian calibration are applied to a more complex example where the relative ages of two architectural components are known from stratigraphy, and where historical information drawn from Hawaiian tradition in- dicates a date after which one of the components, but not the other, was unlikely to have been built. Excavations at H¯apaiali‘i Heiau, site 50–10–37–3817, in Kahalu‘u, Kona, Hawai‘i Island were undertaken in preparation for restoration 132 Thomas S. Dye

Figure 2.15 Estimated ages of construction events: a, ‘Alae Iki heiau structure at site 50–50–17–1088; b, ‘Alae Iki habitation structure at site 50–50–17–1089; c, Kaiholena habitation structure at site 50–10–04–22119; d, Kaiholena habitation structure at site 50–10–04–22201; e, Kaiholena habitation structure at site 50–10–04–22248. The prior information for a, c, d, and e is summarized by (2.7). The prior information for b is summarized by (2.8). of the heiau platform (fig. 2.16). These excavations cleared up a long- standing discrepancy in the description of this structure. When it was first mapped by Stokes (1991) in 1906, the length of the structure, from its seaward end on the lava flat covered at half tide to its rear wall founded in sand and partially covered by brush, was 135 ft. and the width was 100 ft. When Henry E. P. Kekahuna mapped the structure as an “ancient house foundation” (Kekahuna 1952) in the early 1950s, he Traditional Hawaiian Surface Architecture 133 recorded the same width as Stokes did, 100 ft., but a length of 150 ft., some 15 ft. longer than Stokes had measured and mapped. During the most recent investigations of the structure, it was initially thought that Stokes’s more detailed map was likely also to be more accurate than Kekahuna’s map. Excavations focused on finding the foundation of the east wall that shows on Stokes’s map, the top of which was also partially visible, though much disturbed, on the surface. There was no surface evidence for the wall mapped by Kekahuna. The first test excavation, near the southeast corner of the heiau structure, did not find any evidence for the foundation of the east wall. Two other excavations did find evidence, however, one immediately north of the first excavation and another near the northeast corner of the heiau. Charcoal was isolated in a deposit below the base of the wall in the southern test unit and collected for 14C dating. Subsequently, removal of sand east of the heiau revealed the completely buried base of a second wall rising to about three courses, located in the approximate position of the east wall indicated on Kekahuna’s map. Rocks filled the space between the newly-discovered wall and the wall mapped by Stokes, indicating that the heiau platform once extended to this wall, which was part of the heiau structure. Once again, charcoal was isolated in a deposit under the base of this second wall and collected for 14C dating. Based on this structural evidence, it appears that both Stokes and Kekahuna mapped the structure accurately. When Stokes drew his map in 1906, sand must have covered the easternmost wall, hiding it from view, as it did when excavations started there in 2007. Sometime between 1906 and 1952, when Kekahuna mapped the structure, the sand behind the structure must have shifted, exposing the eastern- most wall. Subsequently, the sand moved back, completely covering the wall remnant. Wood charcoal potentially suitable for 14C dating was sent to Gail Murakami of the International Archaeological Research Institute, Inc. Wood Identification Laboratory for identification. In both cases, the plants identified from the wood charcoal collections are native taxa; no historically introduced taxa were identified. These findings support the idea that both walls were constructed in traditional Hawaiian times, before plants and animals introduced in the historic period drastically altered the landscape. In addition, both samples yielded wood charcoal from shrubs. In the sample from beneath the inner wall, wood charcoal from ‘akoko (Chamaesyce sp.), a shrub commonly identified from traditional Hawaiian fires, was identified. The sample 134 Thomas S. Dye

Figure 2.16 Plan of H¯apaiali‘i Heiau; a, the east wall mapped by Stokes (1991); b, the east wall mapped by Kekahuna (1952). At high tide, only the eastern wall of this coastal structure is not in the water. Original plane-table map drawn at a scale of 1:100 by K. Kalawe.

from the outer wall yielded wood charcoal from the shrubs ‘ilima (S. fallax) and ‘ulei¯ (Osteomeles anthyllidifolia). Wood charcoal from shrubs is preferred for 14C dating because shrubs have relatively short lives and thus yield 14C dates that closely approximate the age of the burning event. Traditional Hawaiian Surface Architecture 135

Two samples were submitted to Beta-Analytic for 14C dating. A single piece of ‘akoko wood charcoal from beneath the inner wall was processed normally in the laboratory and yielded sufficient carbon for accelerator mass spectrometry dating. It yielded a conventional 14C age of 590±40 bp. A single piece of ‘ilima wood charcoal from beneath the outer wall was also processed normally in the laboratory and yielded sufficient carbon for accelerator mass spectrometry dating. It yielded a conventional 14C age of 510±40 bp (table 2.2). When this information from the laboratory is calibrated using the most recent atmospheric calibration curve (Reimer et al. 2004), the 95% highest posterior density regions indicate that the dated ‘akoko shrub most likely grew sometime in the period ad 1293–1411 (fig. 2.17a), and that the ‘ilima grew somewhat later, in the period ad 1311–1441 (fig. 2.17b). As can be seen in figure 2.17, the probabilities for the age of the ‘akoko charcoal are skewed to the older end of the range, while those for the ‘ilima charcoal are skewed to the younger end of the range. The ages of the plants found beneath the walls of the heiau give indications of the ages of the walls, but do not date them directly. Although the wood charcoal must be older than the wall built on top of it, the information returned by the dating laboratory gives no indication of how much older the charcoal is than the wall. Deriving estimates of when the walls were built requires additional information in the form of a model that relates the ages of the wood charcoal samples and the two walls. This model can be expressed with two inequalities:

α1 > θ1 > β1 ≥ φinner > φouter > φkapu (2.9)

α1 > θ2 > β2 ≥ φouter > φkapu (2.10)

where: α1 is the beginning of the cultural deposit underlying the two walls; β1 is the land surface on which the inner, older wall was built; β2 is the land surface on which the outer, younger wall was built; θ1 is the calendar date when the ‘akoko wood charcoal was introduced to the deposit; θ2 is calendar date when the ‘ilima wood charcoal was introduced to the deposit; φinner is the construction date of the inner, older wall; φouter is the construction date of the outer, younger wall; and φkapu is ad 1819, the year that the traditional Hawaiian religious system was overthrown and H¯apaiali‘i Heiau was abandoned. 136 Thomas S. Dye

Figure 2.17 Posterior probability distributions for H¯apaiali‘i Heiau; a, age of ‘akoko wood charcoal beneath the inner wall; b, age of ‘ilima wood charcoal beneath the outer wall; c, construction date of the inner wall, using the prior information summarized by (2.9) and (2.10); d, construction date of the outer wall, using the prior information summarized by (2.9) and (2.10); e, construction date of the inner wall, using the prior information summarized by (2.9), (2.10), and (2.11); f, construction date of the outer wall, using the prior information summarized by (2.9), (2.10), and (2.11).

With this additional information, the calibration returns estimates of the construction dates for the two walls of ad 1389–1766 for the inner wall (fig. 2.17c) and ad 1567–1818 for the outer wall (fig. 2.17d). Comparing these with the calibrated ages of the dated shrubs (fig. 2.17a and b) it can be seen that the initial estimates of the construction dates are both younger than the age estimates for the charcoal collected Traditional Hawaiian Surface Architecture 137 beneath them. These initial estimates are a step in the right direction, but they don’t yield very precise estimates of when the walls were constructed. In particular, the age of the inner wall spans a period of 380 years, which is a disappointingly great range. Fortunately, it is possible to add more information to the model that will constrain this range considerably. Although local tradition does not indicate when H¯apaiali‘i heiau was built, it does note that it is older than Ke‘eku¯ heiau, which stands next to it on the shoreline. In 1906, Stokes learned from Malanui that H¯apaiali‘i

was built by Ma‘a, a kahuna of Maui, who left for Kaua‘i later. It was said to be much older than Ke‘eku¯ Heiau. The period of Ma‘a was said to be later than that of Pa‘ao. The heiau was for prayers only. (Stokes 1991, 73)

Ke‘eku¯ is one of two heiau luakini known for Kahalu‘u Ahupua‘a, the other being ‘Ohi‘amukumuku,¯ which was located about 400 m away to the northeast. The first traditional mention of Ke‘eku¯ heiau is in relation to the Hawai‘i Island chief Kalaunui‘ohua,¯ who sacrificed the priestess, or kaula¯ wahine, Wa‘ahia, there and for which he received war powers from the god, K¯anenuiakea (Malo 1996, 126–129, 300–302). Assuming that this information is correct, then H¯apaiali‘i heiau was built sometime before Kalaunui‘ohua¯ assumed the reins of power and sacrificed the kaula¯ wahine at Ke‘eku.¯ When was Kalaunui‘ohua¯ active? Hawaiian tradition doesn’t assign a range of calendar years to Kalaunui‘ohua’s¯ reign because, traditionally, Hawaiians used their calendar to measure time within the annual cycle and not as a basis for longer-term time measurement. Instead, time at this longer scale was related to important natural and social events well known by the people. Lacking a direct statement in calendar years, the typical way to estimate the antiquity of a ruling chief involves three steps: (i) count the number of generations between the chief of interest and a chief of the historic period whose calendar birth year can be fixed, (ii) multiply this figure by the length of a generation, which is “the age of the parent at the birth of the child next in the genealogy” (Stokes 1933), and (iii) subtract the resulting figure from the birth year of the historic period chief. This is the tactic followed here. The basis for each of the three steps is laid out in detail in what follows, where a mathematical model suitable for inclusion in the Bayesian calibration is derived. 138 Thomas S. Dye

The genealogies of ruling chiefs indicate that Kalaunui‘ohua¯ was born fifteen generations before Kamehameha (Hommon 1976, 98, fig. 12b; 317–321). The number of generations is clear because in most generations between Kalaunui‘ohua and Kamehameha, the ruling chief was succeeded by his son. In the four cases where the chief was not succeeded by a son, the generation of the usurper can be identified as the same as the son’s by various genealogical connections. Estimates of Kamehameha’s birth-year vary somewhat. The esti- mate by Kamakau (1992) of 1736 has been viewed by most subsequent researchers as too early. Stokes (1933) discusses this matter at length, concluding that the most likely birth year was 1756, twenty years later than the date cited by Kamakau. Perhaps more widely accepted today is an estimate of 1758 based on a statement by the Spaniard, Francisco de Paula Marin, that Kamehameha was aged 60 years and 6 months when he died in the spring of 1819; this is the date adopted herein. Although the length of a generation in traditional Hawai‘i has been estimated at 30 and 25 years, Stokes (1933) makes a convincing argument that, in the case of Hawaiian ruling chiefs, the best estimate is closer to 20 years. This is based primarily on the concern among high ranking families that a son choose a spouse who would provide offspring of the highest possible rank. The right of primogeniture, or the right of succession belonging to the first born child of the chief, being the rule for normal political succession, the first wife of a high ranked chief was carefully chosen, as the following account testifies: He mea hookaawale ia na [a]lii It was the practice to sepa- nui, i loaa mai na keiki alii, aole no rate the ali‘i nui [high chief] so e moe jua e me ka wahine i uuku that children of an ali‘i [chiefly] iho kona [a]lii ana, aole hoi e moe rank could be conceived. An ali‘i mua me ka wahine noa. [chief] did not casually sleep with a woman of a lesser rank nor with a woman free from any restric- tions. Aka, e imi mua ia ka wahine But, a woman of ali‘i nui [high alii nui, a kane alii niu paha, e chief] rank was first looked for ka poe kuauhau kupuna maloko o . . . by genealogists through the na mookuauhau, a ike pono ia ka genealogies. The woman was wahine, i nui ke alii, ke kane paha, carefully studied and when the a ike pono ia kona mau kupuna ali‘i [chief] was mature, perhaps a a mau makua, oia ka wahine, ke young man, and the background kane paha, a na [a]lii nui e moe of his ancestors and parents had Traditional Hawaiian Surface Architecture 139 mua ai. (Malo 1996, 30) been carefully studied, then this was the woman or man for the ali‘i nui [high chief] to sleep with. (Malo 1996, 174)

Because “[d]elay much beyond the age of puberty, among people not otherwise restricting sex-intercourse, would probably be imprac- ticable despite the prescribed segregation” (Stokes 1933, 53), Stokes believed that many chiefs sired first-born soon after the onset of puberty around age 16, thus setting a minimum figure for the length of a generation. For several reasons, Stokes (1933, 59–61) believed that this minimum was not a good estimate of the average, which he put at 20 years. One way to describe Stokes’s estimate mathematically is with a normal distribution that has a mean of 20 years and a standard deviation of two years (fig. 2.18). Under this model, 67 percent of the generations will fall in the range 18–22 years and 95 percent will fall in the range 16–24 years. Of the remaining 5 percent, 2.5 percent of the generations will be shorter than 16 years and 2.5 percent will be longer than 24 years. Using this model, Wa‘ahia was sacrificed on the altar of Ke‘eku¯ in ad 1488 ± 8.

Figure 2.18 Generation length model based on Stokes (1933).

This traditional evidence can be included in the Bayesian model by adding another inequality (2.11) to inequalities (2.9) and (2.10). Here, φkeeku is the year that Kalaunui‘ohua¯ sacrificed the kaula¯ wahine, Wa‘ahia, on the altar at Ke‘eku.¯

φinner > φkeeku (2.11) 140 Thomas S. Dye

There is a dramatic improvement in the construction date esti- mates when this traditional information is added to the Bayesian model. The age estimate for the inner, older wall is powerfully con- strained by the information from Hawaiian tradition and the 95% high- est posterior density region is ad 1369–1496 (fig. 2.17e), a decrease of 250 years compared to the range estimated without traditional information. Correspondingly, the calibration now indicates that the outer, younger wall might also have been built earlier than the initial estimate indicated (fig. 2.17f ). Now, the 95% highest posterior density region for the outer, younger wall is ad 1504–1818, an estimate that includes the early part of the sixteenth century. Based on the 14C and traditional information, using the model specified in inequalities (2.9), (2.10), and (2.11), one can be 95% certain that H¯apaiali‘i heiau was less than 118 years, or about six generations, old when Kalaunui‘ohua¯ sacrificed the priestess, Wa‘ahia, at Ke‘eku¯ heiau.

A Diachronic Settlement Pattern Hypothesis

Estimating the age of a construction event by dating material from an appropriate context is the first step in establishing the use-life of a structure. The second step, estimating the date of abandonment, is often more difficult because datable post-abandonment contexts are extremely rare. In some cases, as in four of the examples set out in this chapter, particular historical information provides a useful terminus ante quem for the abandonment event but does not specify the event directly. In the absence of this particular information, other tactics are required. Cordy (2007, 110–111), in an attempt to reconstruct population sizes of three traditional districts on O‘ahu and Hawai‘i Islands in ad 1778, when Captain and his crew made the first written records of Hawaiian contact with Europeans, has hypothesized that traditional Hawaiian habitation features standing today were all occupied at that time. Although this hypothesis seems intuitively plausible to most archaeologists working in Hawai‘i today, the dating evidence from Kaiholena suggests that it might assume a too long use-life for traditional Hawaiian structures. At the three Kaiholena habitation structures, for example, the probability that the structure was built before ad 1778 is 0.39 for site 50–10–04–22201, 0.32 for site 50–10–04–22119, and 0.34 for site 50–10–04–22248. Based on the available dating evidence it is more likely that each these structures was built after Cook visited the islands, and the probability Traditional Hawaiian Surface Architecture 141 that all three of them were built before ad 1778 is only 0.04. Counting these three structures in a census of Kaiholena for ad 1778 would, with a probability of 0.96, inflate the population estimate by including one or more structures that didn’t exist in ad 1778. Application of the research design set out in this chapter to the structures counted in Cordy’s paleodemographic exercise would greatly increase confidence in the population estimates that it yields. The evidence recovered at Kaiholena for the traditional practice of dismantling abandoned structures for their building materials (see pg. 103) raises the question of whether any of the relatively intact habi- tation structures there were built before ad 1778. Although traditional Hawaiian building materials—cobbles and small boulders of volcanic rock—appear to be ubiquitous on the landscape and thus something of a free good, an abandoned structure represents considerable labor expended on transport of materials, as well as skill in their selection for suitability and size. An abandoned structure located near the site of a new building project must have represented an attractive source of building materials and it is difficult to imagine that the traditional Hawaiian mason would ignore it. On this model, many if not most of the structures built before ad 1778 and in use at that time would have been disassembled for construction projects initiated after that date. It is difficult to know how widespread was the process of using abandoned structures as sources of building materials. Kaiholena, with its partially eroded landscape and sparse cover of low grass, is an ideal place to identify the traces of abandoned and dismantled structures. In areas that are more thickly vegetated, these low-lying traces of old structures are often passed over during survey and are recognized only after vegetation has been removed, as it is in the rare cases when a brush fire precedes the archaeological surveyor (Holm and Kirch 2007). Similarly, in areas where sediment deposition has outpaced erosion, these abandoned structures, which can outnumber those present at the surface today, are found only through removal of large volumes of sediment over broad areas, a scale of excavation that is extremely rare in Hawaiian archaeology (Williams 1992). It is certain that current inventory survey methods in Hawai‘i are poorly suited to discovering the remnants of dismantled structures, whether or not they are buried. On current evidence, such as it is, there is little if any evidential basis for the idea that intact habitation structures represent anything but the very end of traditional Hawaiian times, which in most parts of the islands stretched into the early decades of the nineteenth century. The idea that any of these habitation structures date to an 142 Thomas S. Dye early period requires a leap of faith that can hardly be justified given the availability of the relative and absolute chronological methods described in this chapter. In contrast to the relatively late construction dates for the habita- tion structures, the two construction date estimates for heiau indicate that these structures are likely to be relatively old; H¯apaiali‘i Heiau was built in the fifteenth century and the small heiau at ‘Alae Iki might have been built even earlier. These estimates are in line with con- struction date estimates for heiau structures on Maui Island, derived by techniques that involve less rigorous control over archaeological context and chronological analysis, which range from the thirteenth through the early seventeenth centuries (Kirch and Sharp 2005; Kolb 1991, 2006). These data suggest that there are two components to the di- achronic settlement pattern of traditional Hawaii, with a relatively fixed set of temple structure locations surrounded by a more transient set of habitation structures. Over time, the structures of individual heiau might change as old structures were repaired and remodeled, and new heiau structures might be built, but, once built, the locations of heiau wouldn’t change and their spatial relations to one another would stay the same. In contrast, the pattern of habitation structures almost certainly involved changes in location as old structures were abandoned and then dismantled for the materials needed to build new habitations nearby.

Implications for Population Estimation

This hypothesis—that habitation structures were used for relatively short periods of time—has immediate implications for attempts to reconstruct population histories with archaeological data. Leaving aside the vexed problem of absolute numbers of people and focusing instead on the more approachable and useful problem of estimating rates of change over time, there are currently two approaches used by archaeologists, both of which rely in their own ways on the association of absolute dates and surface structures. One of these is the “house- count” method, which assigns ages to habitation structures and counts how many houses are present in each of two or more generally 100 year periods. The other is the so-called “Dye-Komori” method, which treats the corpus of 14C dates as a sample from a charcoal population that is argued to be isomorphic with population size over time (Dye Traditional Hawaiian Surface Architecture 143 and Komori 1992b; Rick 1987). The two methods are often contrasted with one another, at least partially for differences in the way data are displayed graphically. Both methods construct histograms, though the ways in which they do so are problematic. The Dye-Komori method sums the posterior probabilities of the calibrated ages of the 14C dates, a technique that has been implemented in both the Calib (Stuiver and Reimer 1989) and OxCal software packages. This method of constructing histograms has no sound statistical basis, however, and the resulting histograms do not necessarily represent a valid estimate of the chronological distribution of the dated samples (Blackwell and Buck 2003) as is sometimes claimed (Dye and Komori 1992a). An alternative, often practiced by the house counters, is to use a subjective, often unspecified (e.g. Kirch 2007), procedure to assign a house to one or another 100 year interval bin based on a putatively associated and calibrated 14C age, a practice that effectively ignores any portion of the calibrated age range that extends beyond a bin boundary. Even if the criteria used to assign house structures to bins are fully specified, and a way is found to take into account the full calibrated age ranges, the possible distortions caused by binning are well-known and widely discussed by statisticians (see Chambers et al. 1983; Freedman and Diaconis 1981; Scott 1979, 1992). Both methods would benefit from use of a statistically defensible method for displaying and analyzing the chronological distribution of collections of 14C dates. Aside from these somewhat superficial differences, the two techniques are, practically speaking, much alike, at least as they are implemented in Hawai‘i, and they yield similar results. In fact, in the situation where all of the dates in the analysis are associated with habitation structures that can be counted and there are equal numbers of dates from each structure, the results are the same, except for possible distortions introduced by the ways they construct histograms. A more fundamental problem with the two approaches has to do with how the 14C ages are associated with archaeological events that are tallied. If the hypothesis is correct that habitation structures were used for a short period of time before they were dismantled for their building materials, then the house-count method is only counting houses built and occupied late in traditional Hawaiian times. Houses thought to have been occupied hundreds of years earlier, based on 14C dates, can be explained as a result of their having been dated with materials collected from pre-structure deposits that were carelessly associated with the surface structure, a problem that appears to crop 144 Thomas S. Dye up commonly in Hawaiian archaeology (Dye 2004), or with samples that did not control for the possible effects of old wood. The Dye- Komori method suffers from a similar, but slightly different, problem. Here, the crucial step in the analysis is assigning a household context to the absolute dates used in the analysis. Theoretically, this could be done in any number of ways, but when it is based on the size and form of surface architecture in the vicinity of a charcoal collection locality, the lack of demonstrated association of the surface architec- ture with the dated material means that there is no explicit reason to believe that the dated material derived from a household context. Thus, to the extent that the hypothesis that habitation structures are late is true, paleodemographers following both the house-count and Dye-Komori methods need to start over. The house counters need to demonstrate that the absolute dates in their analyses come from contexts that are confidently associated with use of the counted habitation structures. Followers of the Dye-Komori method must ensure that the criteria used to assign absolute dates to household activities don’t rely on mistaken associations of dates with habitation structures. Additionally, followers of the Dye-Komori method need to take into account the possibility that post-depositional processes similar in effect to that at work on the surface architectural record of Kaiholena are also having an effect on the population of charcoal available for 14C dating (Surovel and Brantingham 2007). Paleodemo- graphic estimates will only gain credibility when suitable controls are in place and archaeologists have a better handle on the diachronic settlement pattern.

Discussion and Conclusion

For many years, Hawaiian archaeologists have labored under the false assumption that it is possible to estimate the age of surface architectural features by dating materials found on, within, or near them. This has resulted in a corpus of 14C dates whose association with particular archaeological events is unclear; in the case of surface architecture this means that the archaeologist has not identified the stratigraphic relationship of the dated material to the base of the structure and so cannot be confident of the relationship between them. In the future, success in estimating the ages of traditional Hawaiian structures will be determined by the degree to which context has been identified, isolated, and then integrated into the analysis. This chapter Traditional Hawaiian Surface Architecture 145 has described a Bayesian statistical framework to accomplish this, along with a set of field procedures to support its data requirements. The Bayesian framework applied in this chapter to the calibration of 14C dates, is extremely broad in its scope (Buck et al. 1996). In theory, it can be applied to almost any archaeological situation and any dating material. This chapter has attempted to illustrate the power of the Bayesian framework with a series of real-world examples from Maui and Hawai‘i Islands, where information on the age of events was obtained from 14C dating, stratigraphy, Hawaiian tradition, and events recorded historically. The conclusion of this exercise—that the Bayesian approach yields tangible benefits—is not new. In Great Britain, the Scientific Dating Co-ordinator of English Heritage, the British equivalent of the Hawaii State Historic Preservation Division, has overseen the widespread adoption of Bayesian methods, which are now applied to around 85 percent of the 14C age measurements from excavation projects, the English equivalent of the data recovery project in Hawai‘i. Based on more than 2,000 14C age determinations, she has determined that the range of the highest posterior density region is reduced by about 35 percent by the use of Bayesian methods. “Looking at this from the point of view of a Civil Service accountant, one could say that the value-for-money of the English Heritage radiocarbon budget has been increased by around 25 percent” (Bayliss and Ramsey 2004, 26). The success of a Bayesian approach to estimating the age of sur- face architecture in Hawai‘i will hinge on the skill with which field- workers identify and isolate contexts that carry information useful in the analysis. The example of the ‘Alae Iki habitation site, where the dated context was isolated beneath a paving that was, itself, strati- graphically inferior to the wall whose age was the goal of estimation, is a good example of the important role that context plays. This example yielded the least precise construction date estimate among the habitation structures because the dated sample wasn’t taken from a context immediately beneath the structure, as in the other exam- ples. This deficiency might have been remedied in the field by a more intensive search for suitable dating material under the wall of the structure, perhaps by expanding the small test pit that was excavated there. As the other examples show, it is usually possible to isolate a suitable context with a relatively modest field effort, one that falls easily within the range of effort expended during inventory survey ex- cavations generally undertaken in Hawai‘i. The basic research design set out in this chapter could thus be applied quite generally in Hawai‘i 146 Thomas S. Dye and an inventory of dated structures might accumulate rapidly given the pace of archaeological research in the islands. Although the Bayesian method provides archaeologically useful estimates of construction dates with the 40 year standard deviations routinely produced by the 14C dating method today, the precision of its estimates will be improved as age estimation techniques become more precise. Application of 238U/230Th dating on branch coral pieces collected from Hawaiian archaeological sites is especially promising, yielding age estimates with standard deviations less than a decade (Kirch and Sharp 2005; Weisler et al. 2005). Unlike the ill-fated attempt in the 1970s to date volcanic glass by measuring its hydration rind (Olson 1983), a technique that practitioners erroneously claimed could yield similarly precise age estimates, 238U/230Th dating is based on a solid scientific footing. Incorporating high precision dates in Bayesian analyses will increase the precision of the dating estimates they yield. The contexts in which coral is found differ from those in which the charcoal and marine shells used for 14C dating are recovered and this provides both opportunity for refining dating analyses and potential for misinterpretation. Charcoal and pieces of marine shell that are used to date Hawaiian archaeological sites are often small; establishing a secure context for them typically involves careful field techniques for ruling out the possibility that they were secondarily deposited. In contrast, many coral pieces are relatively large and their potential for movement through the stratigraphic column is slight. This characteristic broadens the range of architectural contexts that can be dated to include the fill of platforms, terraces, and walls. In these situations, coral collected from the interior of an architectural component can augment or replace charcoal or marine shell collected from beneath the component to provide a terminus post quem for construction. Coral differs from charcoal and marine shell in another way that complicates its interpretation, however. Unlike charcoal and marine shell, which are the waste products of everyday activities, certain pieces of branch coral were invested with significance in old Hawai‘i where they were laid at the altars of temples as offerings to the gods. The depositional history of such a material is likely to be complex. The claim that coeval age estimates on branch coral collected from the surface of two temples on Moloka‘i indicates that offerings were made at both temples at about the same time (Weisler et al. 2005) ignores the possibility that coral offered at one temple was later taken from the altar and offered at the other temple, perhaps in a conscious attempt to link the mana, or access to supernatural Traditional Hawaiian Surface Architecture 147 power, of the two structures. The claim that both structures were in use around ad 1604–1605 is too strong; all that can be said with certainty is that corals of closely similar ages were deposited on the two structures. As a general principle, the age of the deposition event for a surface find can’t be determined solely from the age of the find itself. The age of a surface find provides instead a terminus post quem for the deposition event; some other line of evidence must be used to determine whether the item was deposited when it was young or old. The research design set out in this chapter is ideally suited to ad- dress the diachronic settlement pattern hypothesis—that the locations of heiau structures were fixed, while those of habitation structures changed frequently. Routine application of the research design during inventory survey will plausibly yield the information needed to accept or reject the hypothesis. If the hypothesis is correct, then short- lived dating materials collected beneath habitation structures will yield dates that indicate these structures were built late in traditional Hawaiian times, often after ad 1778 and the advent of written records. Older age estimates will come from contexts associated with remnant structures that were dismantled in traditional Hawaiian times and from buried structures no longer visible on the surface today. The pattern of dates from heiau structures should be more varied, with some relatively early dates from the remains of their original architec- tural components and a variety of younger dates from components that were added later or modified. Regardless of how well the hypoth- esis withstands repeated testing, the data yielded by research carried out according to the design will illuminate the diachronic settlement pattern of old Hawai‘i and provide, for the first time, a sound basis for carrying out paleodemographic studies of traditional Hawai‘i using archaeological data.

References

Athens, J. S., Tuggle, H. D., Ward, J. V., and Welch, D. J. (2002). Avi- faunal extinctions, vegetation change, and Polynesian impacts in prehistoric Hawai‘i. Archaeology in Oceania, 37:57–78.

Bayliss, A. and Ramsey, C. B. (2004). Pragmatic Bayesians: A decade of integrating radiocarbon dates into chronological models. In Buck, C. E. and Millard, A. R., editors, Tools for Constructing Chronologies: Crossing Disciplinary Boundaries, number 177 in Lecture Notes in Statistics, chapter 2, pages 25–41. Springer, London. 148 Thomas S. Dye

Blackwell, P. G. and Buck, C. E. (2003). The late glacial human reoc- cupation of north-western Europe: New approaches to space-time modeling. Antiquity, 77:232–240.

Bonk, W. J. (1968). An archaeological survey of a coastal tract in North and South Kohala, Hawaii. Typescript report prepared for Hawaii Division of State Parks, Department of Land and Natural Resources, Honolulu.

Bowman, S. (1990). Radiocarbon Dating. Interpreting the Past. Univer- sity of California Press, Berkeley, CA.

Bronk Ramsey, C. (2001). Development of the radiocarbon calibration program OxCal. Radiocarbon, 43(2A):355–363.

Bronk Ramsey, C. (2007). OxCal. http://c14.arch.ox.ac.uk/ embed.php?File=oxcal.html. Accessed February 22, 2007.

Buck, C. E., Cavanagh, W. G., and Litton, C. D. (1996). Bayesian Ap- proach to Interpreting Archaeological Data. Statistics in Practice. John Wiley & Sons, Chichester, UK.

Buck, C. E., Christen, J. A., and James, G. (1999). BCal: an on-line Bayesian Radiocarbon Calibration Tool. http://bcal.sheffield. ac.uk.

Chambers, J. M., Cleveland, W. S., Kleiner, B., and Tukey, P. A. (1983). Graphical Methods for Data Analysis. Wadsworth & Brooks/Cole, Pacific Grove, CA.

Cordy, R. (2007). Reconstructing Hawaiian population at European contact. In Kirch and Rallu (2007), chapter 7, pages 108–128.

Cordy, R. H. (2000). Exalted Sits the Chief: The Ancient History of Hawai‘i Island. Mutual, Honolulu.

Cox, R. T. (1946). Probability, frequency, and reasonable expectation. American Journal of Physics, 14:1–13.

Dixon, B., Conte, P. J., Nagahara, V., and Hodgins, W. K. (2000). Kahik- inui Mauka: Archaeological Research in the Lowland Dry Forest of Leeward East Maui. Prepared for Department of Hawaiian Home Lands. Historic Preservation Division, Department of Land and Nat- ural Resources, Kapolei, HI. Traditional Hawaiian Surface Architecture 149

Dye, T. S. (1989). Tales of two : Traditional historical and ar- chaeological interpretations of Hawaiian prehistory. Bishop Museum Occasional Papers, 29:3–22.

Dye, T. S. (1991). A reputation unmade: J. F. G. Stokes’s career in Hawaiian archaeology. In Stokes (1991), pages 3–20. Edited and introduced by Tom Dye.

Dye, T. S. (1994). Apparent ages of marine shells: Implications for archaeological dating in Hawai‘i. Radiocarbon, 36:51–57.

Dye, T. S. (2003). Revised Archaeological Inventory Survey at Kaiholena Ahupua‘a, North Kohala, Hawai‘i. Prepared for Pohaku Kea, LLC. T. S. Dye & Colleagues, Archaeologists, Honolulu.

Dye, T. S. (2004). How to fix the inventory survey rule. Hawaiian Archaeology, 9:123–132.

Dye, T. S., Carson, M. T., and Tomonari-Tuggle, M. J. (2002). Archeo- logical Survey of Sixty Acres of the K¯ıpahulu Historic District Within the K¯ıpahulu District of Haleakala¯ National Park, Maui. Prepared for Department of the Navy, Pacific Division, Naval Facilities Engi- neering Command. International Archaeological Research Institute, Honolulu.

Dye, T. S. and Komori, E. K. (1992a). Computer programs for creating cumulative probability curves and annual frequency distribution dia- grams with radiocarbon dates. New Zealand Journal of Archaeology, 14:35–43.

Dye, T. S. and Komori, E. K. (1992b). A pre-censal population history of Hawai‘i. New Zealand Journal of Archaeology, 14:113–128.

Ellis, W. (1963). Journal of William Ellis. Advertiser Publishing Com- pany, Honolulu.

Emory, K. P., Bonk, W. J., and Sinoto, Y. (1968). Hawaiian Archaeology: Fishhooks. Number 47 in B. P. Bishop Museum Special Publication. Bishop Museum Press, Honolulu.

Freedman, D. and Diaconis, P. (1981). On the histogram as a density estimator: L_2 theory. Zeitschrift f ur Wahrscheinlichkeitstheorie und verwandte Gebiete, 57:453–476. 150 Thomas S. Dye

Geman, S. and Geman, D. (1984). Stochastic relaxation, gibbs distribu- tions, and the bayesian restoration of images. IEEE Transactions on Pattern Analysis and Machine Intelligence, 6:721–741.

Gerrish, G. and Mueller-Dombois, D. (1999). Measuring stem growth rates for determining age and cohort analysis of a tropical evergreen tree. Pacific Science, 53:418–429.

Green, R. C. (1984). Settlement pattern studies in Oceania: An in- troduction to a symposium. New Zealand Journal of Archaeology, 6:59–69.

Hacking, I. (1975). The Emergence of Probability. Cambridge University Press, Cambridge.

Hacking, I. (2001). An Introduction to Probability and Inductive Logic. Cambridge University Press, Cambridge.

Harris, E. C. (1989). Principles of Archaeological Stratigraphy. Aca- demic Press, London, second edition.

Harris, E. C., Brown, M. R., and Brown, G. J. (1993). Practices of Archae- ological Stratigraphy. Academic Press, London.

Hartzell, L. L., Lebo, S. A., Lennstrom, H. A., McPherron, S. P., and Olszewski, D. I., editors (2004). Activities and Settlement in an Upper Valley: Data Recovery and Monitoring Archaeology in North Halawa¯ Valley, O‘ahu, volume 1. Anthropology Department, B. P. Bishop Museum, Honolulu.

Holm, L. and Kirch, P. V. (2007). Up in smoke: Assumptions of surface visibility and site identification. Hawaiian Archaeology, 11:83–100.

Hommon, R. J. (1976). The Formation of Primitive States in Pre-Contact Hawaii. PhD thesis, University of Arizona, Tucson, AZ.

Howson, C. and Urbach, P. (1989). Scientific Reasoning: The Bayesian Approach. Open Court, La Salle, IL.

Hughen, K. A., Baillie, M. G. L., Bard, E., Beck, J. W., Bertrand, C. J. H., Blackwell, P. G., Buck, C. E., Burr, G. S., Cutler, K. B., Damon, P. E., Edwards, R. L., Fairbanks, R. G., Friedrich, M., Guilderson, T. P., Kromer, B., McCormac, G., Manning, S., Ramsey, C. B., Reimer, P. J., Reimer, R. W., Remmele, S., Southon, J. R., Stuiver, M., Talamo, S., Taylor, F. W., van der Plicht, J., and Weyhenmeyer, C. E. (2004). Traditional Hawaiian Surface Architecture 151

Marine04 marine radiocarbon age calibration, 0–26 cal kyr bp. http: //www.radiocarbon.org/IntCal04.htm/Marine04.14c.

Hume, D. (1956). An Enquiry Concerning Human Understanding. Gate- way Editions; distributed by H. Regnery Co, Chicago.

Hume, D., Selby-Bigge, L. A., and Nidditch, P. H. (1980). A Treatise of Human Nature. Clarendon Press, Oxford, second edition.

Hunt, T. L. and Holsen, R. M. (1991). An early radiocarbon chronology for the Hawaiian Islands: A preliminary analysis. Asian Perspectives, 30:147–161.

Kamakau, S. M. (1992). Ruling Chiefs of Hawaii. Kamehameha Schools Press, Honolulu, revised edition.

Kekahuna, H. E. P. (1952). Map showing Kaha-lu‘u Beach. Dept. of Anthropology, B.P. Bishop Museum, Honolulu.

Kirch, P. V. (1973). Archaeological Excavations at Kahalu‘u, North Kona, Island of Hawaii. Prepared for Kamehameha Development Corpora- tion. Anthropology Department, B. P. Bishop Museum, Honolulu.

Kirch, P. V. (1985). Feathered Gods and Fishhooks: An Introduction to Hawaiian Archaeology and Prehistory. University of Hawaii Press, Honolulu.

Kirch, P. V. (2007). Paleodemography in Kahikinui, Maui: An archaeo- logical approach. In Kirch and Rallu (2007), pages 90–107.

Kirch, P. V. and Rallu, J.-L., editors (2007). The Growth and Collapse of Pacific Island Societies: Archaeological and Demographic Perspec- tives. University of Hawai‘i Press, Honolulu.

Kirch, P. V. and Sharp, W. D. (2005). Coral 230Th dating of the impo- sition of a ritual contral hierarchy in precontact Hawaii. Science, 307:102–103.

Kolb, M. J. (1991). Social Power, Chiefly Authority, and Ceremonial Architecture in an Island Polity, Maui, Hawaii. PhD thesis, University of California, Los Angeles.

Kolb, M. J. (2006). The origins of monumental architecture in ancient hawai‘i. Current Anthropology, 47(4):657–665. 152 Thomas S. Dye

Ladd, E. J. (1973). Kaneaki temple site—an excavation report. In Ladd, E. J., editor, Makaha Valley Historical Project: Interim Report No. 4, number 19 in Pacific Anthropological Records, report 1, pages 1–30. Anthropology Department, B. P. Bishop Museum, Honolulu.

Major, M. and Dye, T. S. (2005). Archaeological Data Recovery at the Coastal Portion of Kaiholena Ahupua‘a, North Kohala, Hawai‘i. Prepared for Pohaku Kea LLC. T. S. Dye & Colleagues, Archaeologists, Honolulu.

Malo, D. (1996). Ka Mo‘olelo Hawai‘i: Hawaiian Traditions. First People’s Productions, Honolulu. Translated by Malcolm Naea Chun.

McAllister, J. G. (1933). Archaeology of Oahu. Number 104 in B. P. Bishop Museum Bulletin. Bishop Museum Press, Honolulu.

Olson, L. (1983). Hawaiian volcanic glass applied “dating” and “sourc- ing”: Archaeological context. In Clark, J. T. and Kirch, P. V., edi- tors, Archaeological Investigations of the Mudlane-Waimea-Kawaihae Road Corridor, Island of Hawai‘i: An Interdisciplinary Study of an En- vironmental Transect, number 83–1 in Departmental Report Series, pages 325–340. Anthropology Department, B. P. Bishop Museum, Honolulu.

Ramsey, C. B. (1995). Radiocarbon calibration and analysis of stratig- raphy: The OxCal program. In Cook, G. T., Harkness, D. D., Miller, B. F., and Scott, E. M., editors, Proceedings of the 15th International 14C Conference, volume 37 of Radiocarbon, pages 425–430.

Reimer, P. J., Baillie, M. G. L., Bard, E., Bayliss, A., Beck, J. W., Bertrand, C. J. H., Blackwell, P. G., Buck, C. E., Burr, G. S., Cutler, K. B., Damon, P. E., Edwards, R. L., Fairbanks, R. G., Friedrich, M., Guilderson, T. P., Hogg, A. G., Hughen, K. A., Kromer, B., McCormac, G., Manning, S., Ramsey, C. B., Reimer, R. W., Remmele, S., Southon, J. R., Stuiver, M., Talamo, S., Taylor, F. W., van der Plicht, J., and Weyhenmeyer, C. E. (2004). IntCal04 Terrestrial Radiocarbon Age Calibration, 0–26 cal kyr BP. http://www.radiocarbon.org/IntCal04.htm/IntCal04.14c.

Rick, J. W. (1987). Dates as data: An examination of the Peruvian preceramic radiocarbon record. American Antiquity, 52:55–73.

Rosendahl, P. H. (1975). Archaeological base map, lower Kipahulu district, Haleakala National Park, Maui. Drawing 162/82003, Hawaii Traditional Hawaiian Surface Architecture 153

Field Office, National Park Service, U.S. Department of Interior. 105 sheets, scale 1 in. = 5 m.

Schilt, R. and Sinoto, A. (1980). Limited Phase I Archaeological Survey of Mahukona Properties, North Kohala, Island of Hawai‘i. Prepared for Belt, Collins & Associates. Anthropology Department, B. P. Bishop Museum, Honolulu.

Scott, D. W. (1979). On optimal and data-based histograms. Biometrika, 66:605–610.

Scott, D. W. (1992). Multivariate Density Estimation: Theory, Practice, and Visualization. Wiley series in probability and mathematical statistics. Wiley, New York.

Sinoto, A. (1979). Archaeological Reconnaissance Survey of Mahukona Properties, North Kohala District, Island of Hawai‘i. Prepared for Belt, Collins and Associates. Anthropology Department, B. P. Bishop Museum, Honolulu.

Sivia, D. S. (1996). Data Analysis: A Bayesian Tutorial. Clarendon Press, Oxford.

Soehren, L. J. (1963). An Archaeological Survey of Portions of East Maui, Hawaii. Prepared for National Park Service, U.S. Department of Interior. Anthropology Department, B. P. Bishop Museum, Honolulu.

Soehren, L. J. (1964). An archaeological reconnaissance of the Mahukona-Kawaihae highway, Kohala, Hawaii. Typescript report in Department of Anthropology, B. P. Bishop Museum, Honolulu.

Soehren, L. J. (1969). An archaeological reconnaissance of Parker Ranch coastal lands, North Kohala. In Pearson, R. J., editor, Archaeology on the Island of Hawaii, number 3 in Asian and Pacific Archaeology Series, pages 15–24. Social Science Research Institute, University of Hawaii, Honolulu.

Stein, J. K. and Farrand, W. R., editors (1985). Archaeological Sediments in Context. Peopling of the Americas. Center for the Study of Early Man, Institute for Quaternary Studies, University of Maine at Orono, Orono, ME.

Stokes, J. F. G. (1933). New bases for Hawaiian chronology. Annual Report of the Hawaiian Historical Society, 41:23–65. 154 Thomas S. Dye

Stokes, J. F. G. (1991). Heiau of the Island of Hawai‘i: A Historic Survey of Native Hawaiian Temple Sites. Number 2 in Bishop Museum Bulletin in Anthropology. Bishop Museum, Honolulu. Edited and introduced by Tom Dye.

Stuiver, M. and Polach, H. (1977). Discussion: Reporting of 14C data. Radiocarbon, 19:355–363.

Stuiver, M. and Reimer, P. (1989). Histograms obtained from comput- erized radiocarbon age calibration. Radiocarbon, 31:817–823.

Surovel, T. A. and Brantingham, P. J. (2007). A note on the use of tem- poral frequency distributions in studies of prehistoric demography. Journal of Archaeological Science, 34:1868–1877.

Taylor, R. E. (1987). Radiocarbon Dating: An Archaeological Perspective. Academic Press, New York.

Thien, S. (1979). A flow diagram for teaching texture-by-feel analysis. Journal of Agronomic Education, 8:54–55.

Vincent, C. H. (1988). Treatment of discrepancies in radiocarbon dating. Radiocarbon, 30(2):157–160.

Weisler, M. I., Collerson, K. D., Feng, Y.-X., Zhao, J.-X., and Yu, K.-F. (2005). Thorium-230 coral chronology of a late prehistoric Hawaiian chiefdom. Journal of Archaeological Science, 33(2):273–282.

Weisler, M. I. and Kirch, P. V. (1985). The structure of settlement space in a Polynesian chiefdom: Kawela, Molokai, Hawaiian Islands. New Zealand Journal of Archaeology, 7:129–158.

Williams, S. S. (1992). Early inland settlement expansion and the effect of geomorphological change on the archaeological record in K¯ane‘ohe, O‘ahu. New Zealand Journal of Archaeology, 14:67–78.

Abstract

This chapter develops the conceptual and methodological frame- works needed to investigate the relative and absolute dating of tradi- tional Hawaiian surface architectural features. Its goal is to provide a general research design for the investigation of the diachronic set- tlement pattern. Its main line of argument is that this is something Hawaiian archaeologists can achieve with only a modest field effort, well within the range of effort typically expended during a routine inventory survey. The two key changes in archaeological practice that Traditional Hawaiian Surface Architecture 155

it prescribes—a holistic, systematic approach to stratigraphy and a model-based approach to chronological analysis—are both situated in precisely formulated, well established, and actively developing lines of archaeological research. A series of worked examples illustrate the application of these stratigraphic and chronologic analyses, and pro- vide the data for a hypothesis on the diachronic settlement pattern of traditional Hawai‘i. The hypothesis is framed in such a way that it can be routinely tested under a wide variety of field conditions. Finally, an argument is made that testing and refinement of the hypothesis will have immediate implications for broader questions of traditional Hawaiian history, such as the history of population.

3

Lady Mondegreen’s Hopes and Dreams Three Brief Essays on Inference in Hawaiian Archaeology

Dave Tuggle International Archaeological Research Institute, retired

Original: The girl with kaleidoscope eyes Mondegreen: The girl with colitis goes by

Original from Lucy in the with Diamonds J. Lennon and P. McCartney The provisional outcome of the preliminary investigation indicates that the site was most likely a temporary use shelter cave feature . . .

Hawaiian archaeology report In recent years . . . “methodology” has been increasingly used as a pretentious substitute for “method” in scientific and technical contexts, as in “The oil company has not yet decided on a methodology for restoring the beaches.” This usage may have been fostered in part by the tendency to use the adjective “methodological” to mean “pertaining to methods,” inasmuch as the regularly formed adjective “methodical” has been preempted to mean “orderly, systematic.” But the misuse of “methodology” obscures an important conceptual distinction between the tools of scientific investigation (properly “methods”) and the principles that determine how such tools are deployed and interpreted.

American Heritage Dictionary

157 158 Dave Tuggle

Polonius: What is the matter, my lord? Hamlet: Between who? Polonius: I mean, the matter that you read, my lord. Hamlet: Slanders, sir; for the satirical rogue says here that old men have grey beards, that their faces are wrinkled, their eyes purging thick and plum-tree gum, and that they have a plentiful lack of wit, together with most weak hams; all which, sir, though I most powerfully and potently believe, yet I hold it not honesty to have it thus set down, for yourself, sir, shall grow old as I am, if like a crab you could go backward. Polonius: [Aside] Though this be methodology, yet there is madness in’t.

Misquoting Polonius Hamlet, Act 2, scene 2

Introduction

Archaeologist, crime scene investigator, or Shakespearean scholar— anyone who attempts to know something about the past faces the problem of equifinality. This can also be called the problem of the ar- chaeological mondegreen, which I prefer because I just like the sound of that phrase. When I read conclusions in archaeological reports, they often rattle around in my head like archaeological versions of “the girl with colitis goes by.” The problem of deriving coherent archaeo- logical inferences about the Hawaiian past is the subject of the three brief essays in this paper. The first considers inference about the patterning of archaeological remains where the ancient sky functions as one historical known, the second considers the case of the refuge cave—where inference has been largely built on a ruling hypothesis (Chamberlin 1965; Elliot and Brook 2007), and the third examines the peculiar history of Hawaiian volcanic glass dating—which may be characterized as a tale of fantasy in which the discipline engaged in collective suspension of disbelief. Each of these has a methodological component, including how patterns are identified and interpreted, how site function is approached as either a definition or a hypothesis, and how data require some framework for interpretation. Shakespearean scholar is mentioned above not only as a rhetori- cal connection to the misquote of Polonius, but also momentarily to place this discussion in the larger context of how people reason—and so often reason erroneously (Fischer 1970). An instructive case is the never-ending controversy about the authorship of Shakespeare’s plays, as recently presented so clearly by James Shapiro (2010). This in- volves many interesting approaches to investigating the past, not only Lady Mondegreen’s Hopes and Dreams 159 academic historical research, but also cryptology, séance, and even archaeology when tombs are opened to look for hidden manuscripts. Those who search for an author other than Shakespeare often initiate their arguments with a flawed deduction. The problem is to find who wrote the plays attributed to Shakespeare, because, it is deduced, he could not have done so, because he was of the wrong class . . . or not well-learned . . . or not well-traveled . . . or several other nots. This question arises in part because there is very little artefactual evidence, that is, few documents, about Shakespeare the person. One of the few existing documents is his will. The will lists no books or manuscripts, thus demonstrating that this man had no library, and so could not have been the author of the plays that are full of literary allusions. The difficulty here is the misunderstanding of what this artifact was in its historic context. A will was only one of the estate documents. Another was the inventory of properties, where books would have been listed. This is a missing Shakespeare artifact: his property inventory has never been found (Shapiro 2010, 50). At first glance the practice of Hawaiian archaeology might seem quite simple because sites are relatively shallow, there is limited stratification, portable artifacts are few in number, and there is ex- tensive surface architecture. But it is precisely these characteristics that make the practice of Hawaiian archaeology quite difficult. With- out question there was a great deal of complexity and diversity in Hawaiian cultural behavior, but there is an enormous gap between that behavioral reality and the relative homogeneity of the Hawaiian archaeological landscape—the archaeological reality. Or perhaps this is only apparent homogeneity because of the relative homogeneity of our descriptive, and perhaps inferential, terminology. This derives in part, I suspect, from the conflation of method and methodology. Most reports in Hawaiian archaeology have a section devoted to a description of how the work was conducted. This section is variously labeled Field Techniques, Field Methods, or frequently Methodology. If there is any additional attention to methodological matters, reports will include a perfunctory section on the criteria by which site or fea- ture function is inferred. Even if authors were inclined to cite the body of work that establishes the methodological constructs of Hawaiian archaeology they would have a hard time doing so because there is no such body of work. First Principles of Methodology in Hawaiian Archaeology has yet to be written. Methodological studies have of course been carried out in Hawai‘i, but in large part they have not been subjected to rigorous evaluation and refinement within the com- 160 Dave Tuggle munity of scholars, and no codified set of methodological principles has emerged. Archaeological conclusions ultimately tumble out as prescriptive sentiments based largely on ad hoc deductive construc- tions or conventional representations of what we think we were told by Menzies and Ellis, Kamakau and Malo. We find methodological comfort in being able to employ the richness of Hawaiian traditions in the island version of the direct historical approach to inference. But this comfort is costly. The possible multiverses of the past are ignored.

Stars: Practical

[T]he po‘e kilo hoku () of Oahu and Kauai . . . were very skilled in reckoning the months of the year and in discerning the ways of the , the , and the stars, as well as knowing the configurations of the earth . . . [They] went to stay . . . on Kauai to make their observations. [On Kaua‘i] for the setting . . . of the sun from Ka‘ula to Kawaihoa on , the season was called Kau. It was also called Kau for Kaulana-a-Kane, the resting place of Kane [on Ka‘ula].

Kamakau (1976, 14) There are fifteen chiefly stars and they are the chiefs of the commoner stars. The priestly stars, they are the guides for the land stars, the stars of the months and the three stars of the gods, which are always encircled by all the other stars for an entire year.

N¯amakaokeahi (2004, 8)

In the traditional Hawai‘i revealed by the ethnographic literature, the sky was a critical part of the perception of the world that found expression in traditions, ritual, place names, and architecture, and the sky is the only component of the past that we can reconstruct with certainty.1 Given these two factors, the argument here is sim- ple: archaeoastronomy, largely neglected in Hawai‘i, should be part of the methodological concerns of the discipline; archaeologists in Hawai‘i should routinely include observations and analysis relevant to as a part of site description and investigation; and those observations should be made based on pattern analysis protocols developed as a part of archaeoastronomical methodology relevant to

1This essay is a revised version of the paper Practical Archaeoastronomy, pre- sented at the Annual Conference of the Society for Hawaiian Archaeology, Maui, October 2006. My recognition of the importance of this subject was ignited by Bruce Masse and Rubellite Johnson, and my understanding continues to develop from collaboration with them. Lady Mondegreen’s Hopes and Dreams 161 the nature of Hawaiian archaeological sites. At the same time, there is a corollary to this argument, which is that the definition of site should be reconsidered. If there are evident reasons why little attention is given to ar- chaeoastronomy in Hawai‘i, may be instructive. Stone- provides a critical case of an ancient monument that was dedi- cated to celestial events and that was used as an architectural means to view those events, but it is also a type-site for the history of excess and skepticism concerning archaeoastronomy. Before the early 1960s, archaeologists and astronomers had given sporadic attention to ar- chaeological sites for their astronomical potential, but it was the study of Stonehenge in 1963 by (1964, 1965) that brought this interest into focus. Hawkin’s study ignited research on this subject and essentially led to the creation of the field of ar- chaeoastronomy. However, astronomer Hawkins erred in excess when he disregarded relevant archaeological data and overstepped the po- tential of the astronomical information. His critics, archaeologists and others, erred in skepticism when they underestimated the capabilities of the builders of Stonehenge and when they, the critics, stepped over the archaeological data that spoke to astronomy. In a classic petulant statement, historian Jacquetta Hawkes said of Hawkin’s research “Ev- ery age has the Stonehenge it desires—or deserves” (Hawkes 1967). But that truism about Stonehenge and archaeology in general is a philo- sophical and epistemological dead-end, so we can take the lessons of Stonehenge and move toward methodological realism, which of course leads us to the tools for Practical Archaeoastronomy. First, this obviously involves some basic knowledge of astronomy, the relevant component of which is sometimes called spherical as- tronomy or, in a more restricted sense, positional astronomy, that is: understanding how things move—or appear to move—across the visible celestial sphere. For the Practical Archaeoastronomer, a recom- mended resource is a book by astronomer James Kaler (1996) titled The Ever-Changing Sky: A guide to the Celestial Sphere. Those unfa- miliar with this book may be surprised to learn that it was written just for us, anthropologists who seldom look up. Kaler was preparing a seminar on astronomy for graduate students in anthropology when he realized that he lacked modern class material: no astronomer had written a useful book on in over 50 years. So he wrote this book, and recognizing a common deficiency within our discipline, he wrote it as a comprehensive, but non-mathematical guide to the subject. 162 Dave Tuggle

Second, the tool that has truly made practical Practical Archaeoas- tronomy is the computer sky-simulation program—our time machine. There are a number of these available: a program called TheSky2 is considered one of the best. These programs provide accurate and precise simulations of celestial positions from any location on earth for any moment into the past up to 4000 years BCE. How this in- formation on spherical astronomy is being applied in archaeo- and ethnoastronomy can be investigated in the many articles and books on the subject. Third, knowledge of Hawaiian ethnoastronomy is a necessary tool to use for structuring questions and making observations, but of course analytically and not prescriptively. Bryan and Crowe’s (2002) Stars Over Hawaii, Johnson and Mahelona’s (1975) Na Inoa Hoku, Makemson’s (1941) The Morning Star Rises, and N¯amakaokeahi’s (2004) The History of Kanalu are fundamental works. Finally, there is the matter of what observations to make and how to make them. These are related to the nature of the celestial phenomena of interest, to pertinent ethnoastronomical elements, the types of sites involved, and the archaeological problems to be solved— in other words, there is not a prescriptive answer to this, but rather this is the domain of methodological enquiry. Celestial phenomena and ethnoastronomical perceptions and in- terpretations can be divided into two main categories. One of these is the category of events that were important to demonstrate that order was being maintained in the cosmos and these include calendrical references, that is solar positions, phases of the moon and the moon’s 18 year standstill, position of the , important individual stars, and the rising and setting of the .3 The other is disorder or the deviation from order, noted in such things as , meteors, and supernovae: this is the realm of omens, the portents of great change—disastrous or joyful, and the elements of prediction.4 Both

2This application is currently marketed by Software Bisque as TheSkyX at www. bisque.com. 3The rising and setting of the Pleiades is a seasonal marker not only in Hawai‘i, but also in much of the tropical world. 4Among the many ethnographic examples in Hawai‘i is the association of the birth of Kamehameha with a “new star” almost certainly Halley’s , and there are that can be interpreted to represent episodic celestial events. It should be noted that Hawaiian observation of the sky for omens was not limited to astronomical phenomena, but also included what is today termed atmospheric, that is, such things as cloud formation, moon halos, and . These events are noted in traditions and special occurrences may have been recorded in petroglyphs. Lady Mondegreen’s Hopes and Dreams 163 of these domains of sky watching need to be explored when the tradi- tional astronomy of an area is being reviewed and research questions are being formulated, and in addition they need to be considered within the topographic setting. Rigor in recording for archaeoastronomical purposes certainly involves careful mapping, a precise description of site orientation, but with particular attention to terrain and terrain features. This last is seldom carried out if there is no concern with archaeoastronomical components of a site. But, in part, this inattention to topography may reflect the common view that structural positioning is a simple function of terrain and needs no careful recording. For example, re- garding temple orientation, this argument was advanced a century ago by Stokes (1991, 34, 35) when he made the generalization that temples located along the coastline are parallel to the coast and those located inland are oriented to the contour of the ground. It might also be argued that because of the northwest-southeast axis of the islands themselves, temple orientation in most locales will necessarily be in the direction of the rising or setting sun, so solar archaeoastronomy is meaningless, but this is a fallacious example of an alignment tautol- ogy, as it can be pointed out that a structure has a single orientation, whereas the sun’s position on the horizon in Hawai‘i has a range of more than 50 degrees. In contrast to these generalizations, Kirch (2004) recently empirically addressed the question of terrain position- ing with data from the temples of Kahikinui, Maui and concluded that the “heiau in Kahikinui clearly have preferred orientations” (Kirch 2004, 112) which he hypothesizes to be associated with “particular set[s] of astronomical or landscape phenomena” (Kirch 2004, 112). To evaluate astronomical potential, one can begin with field ob- servation of site location and orientation and compare this with the celestial possibilities based on the astronomy itself and on events of traditional interest. There are a number of sites in Hawai‘i where inves- tigation has indicated astronomical orientation, such as Pu‘uomahuka, the large temple on Oahu’s north shore, whose long axis is directed toward the position of the setting sun of summer (Kyselka 1991), and Ahu a ‘Umi, the isolated temple on the upland flank of Hual¯alai on the island of Hawai‘i, that provides a guide to the position of the Pleiades (Silva and Johnson 1982). Alternatively, traditions and place names associated with locales or with temples provide indications of what celestial observations should be focused on in these cases. Many temples have the same names as stars, and others have names invoking the sun or moon, such as 164 Dave Tuggle

Kuaokal¯a, “back of the sun,” at Ka‘ena, also known as Moka‘ena Heiau, and Hikinaakal¯a, “rising of the sun,” on Kaua‘i; others have specific traditions indicating their association with the celestial, such as the temple at K¯aneloa on O‘ahu, which was dedicated to the sun (Nakuina in Sterling and Summers 1978, 98). On Maui, the religious structures of Haleakal¯a provide a substantial potential for astronomical association because of the traditions of the demi-god Maui [M¯aui] snaring and stopping the sun, which can be read as reference to controlling the (see Beckwith 1940, 227). As indicated in many sources, the name Haleakal¯a is not the pre- contact one; the many variations recorded as the traditional name or names all relate to Maui and the sun, including: (i) Alehekal¯a, sun snarer; (ii) Aheleakal¯a, rays of the sun; (iii) Kal¯ahaku, sun master; and (iv) Ala Hea Kal¯a, the calling of the sun. The religious structures of Haleakal¯a include one major temple located on the south rim and several shrines within the crater. These are all difficult to get to and have not been studied archaeoastronomically, but a review of the existing data suggests enormous potential (Tuggle and Tomonari- Tuggle 2006). In fact, I would argue from a review of island-wide data that Haleakal¯a may have been the most important center of astronomy in the archipelago. In addition to the story of Maui and sun-snatching, there are many other Haleakal¯a traditions that could be explored for architectural expression of celestial traditions, such as those associated with Kana, a stretching-tree kupua, and perhaps a variation of Maui, and with one of the shark brothers of Pele who had a home in the crater. The turning of the great shark of the Milky Way could be ideally positioned against landscape features of the rim from locations within the crater, and at the same time archaeological locations in the crater could be predicted based on the movement of the Milky Way. This discussion has moved from the orientation of temples to the matter of geographic location and view planes. Temple orientation is important in many cases, but this is too narrow a focus. Many factors may determine where a temple is placed and what its orientation is, and if it was used as a place for sky-viewing or for ritual related to celestial phenomena this may not be expressed in an architecturally obvious manner. This is said in view of that fact that we are seldom able to recon- struct all of the many structures and features there were on the temple platform that is recorded, and thus information on the orientation and use of superstructure elements has many possibilities. If we look at Lady Mondegreen’s Hopes and Dreams 165 world , celestial observation itself is essentially a cultural universal, but in a great many cases this involved no structures at all, or at most something as simple as crossed sticks for a viewing aid. In Hawai‘i, of the three solid references in traditional literature to priests marking the movement of the sun, including two descriptions by Kamakau, all mention the use of landmarks, but not one of them clearly indicates that the priest-observer is positioned at a temple. So a review of the possible important celestial events related to a project area should include consideration of possible viewing positions that might be relatively non-descript structures. This takes us to the matter of observations not commonly made: this is the recording of the landscape profile and prominent features as seen from a site or project area. For at least major sites such as temples and chief’s residences, locations were traditionally selected by priests, and it can be certain that relationship to topographic features was one of the important factors that guided the priests. In most places in Hawai‘i, the tra- ditional names of many topographic features are known from the ethnographic sources.5 The recording of the cultural landscape, and associated computer modeling, may suggest the potential for astro- nomical associations with prominent points, and this can be evaluated by modeling the ancient sky via computer simulation, set against the landscape features and profile. Modern readings on prominent points is a beginning— establishing the location of star rising over a landscape above the horizontal plane requires direct observation because of the angle of rise at this latitude. To establish the location of rising in the past is most easily done with computer simulation of precession, as can be shown with the different rising positions of the Pleiades a millennium apart. Now, with these data, celestial phenomena can be reviewed for association with landscape features within the view plane of the site in question. Further, through simulation, transient historical phenomena with well-known positions, such as comets and supernovae, can also be placed within this landscape-horizon. Traditions indicate that these were extremely important events as portents to people the world over, certainly including Hawai‘i. The of 1054 was one of the most spectacular celestial events of the last 2,000 years, and it was

5Ethnographic sources include many place name variations, as noted in the discussion of Haleakal¯a. 166 Dave Tuggle recorded in many ways in cultures around the world, from written records in China to pictographs in the Americas. It can be recognized in images because of its relation to the moon on a particular day of its short span. There are a number of suggestions in Hawaiian traditions and place names that this event was recorded here and as suggested by Rubellite Johnson in a personal communication may be the source of the name of Hual¯alai. For a moment, let us suggest that some reflection by means of Impractical Archaeoastronomy, that is gaining the appreciation of the sky at night seen from the places where we are doing research, appreciation that should contribute to a sense of place where the people of old were watching sky. This will be enhanced with percep- tual transformations that include thinking in terms of ancient names, concepts, and traditions: for example, looking at the not as Greco-Roman images, but as they were configured in traditional Hawai‘i, such as Hok¯ u‘iwa¯ rather than Boötes, such as the turning of the great shark (and perhaps mo‘o in the background) across the winter sky, and as the Water of K¯ane, and by looking at the positions of the sun on the horizon as the Kukulu¯ o Kahiki, or pillars of Kahiki, and at the projection of the ecliptic as the path of the spider. Awakening from reverie, practical archaeoastronomers will im- mediately take advantage of this and recognize that the view of the landscape, in its broadest sense, is part of the site they are standing on and will begin to construct archaeoastronomical associations of archaeological remains, landscape, and sky. Does the argumentation differ from that of other areas of research in archaeology? No; there is the interplay of data and ideas and the generation of inferences, hypotheses, and models. Yet there is one difference in the argumentation structure: archaeoastronomy has a set of data that we know to represent real conditions in the past. It is the only time-machine we will ever have. Why would we not want to use it? To reconstitute the statement of Jacquetta Hawkes, if we don’t explore the potential of site astronomy, we are not giving the Hawaiian past the full attention it deserves.

Postscript

A review of the hundreds of archaeological reports produced in the last 35 years of Hawaiian archaeology would reveal, I believe, a curi- ously confined, static, and uninquisitive archaeological world-view. Lady Mondegreen’s Hopes and Dreams 167

The problem of this world-view is one of the themes of these three essays, and is also implicit in the chapter by Dye on “Traditional Hawaiian Surface Architecture.” In all of these cases, that world-view is challenged. Placing ourselves in the dynamic environment of the past as much as possible is one way to accomplish that—and ar- chaeoastronomy is one such means to this end. Another approach to this is the relatively new field of , that is the study of the acoustical properties of archaeological sites and locales (e.g. Loose 2008; Scarre and Lawson 2006). A moment’s reflection will bring the realization that where sound is a vital part of past behavior, as it is for music, ceremonial performance, recitation, and channeling the voice of deities, acoustic characteristics may have been the most important reason for choosing certain locales that are now archaeological sites.

Caves: Seeking Methodological Shelter

Hawaiian archaeological projects commonly list research problems that include determination of the function of a site or surface features of a site.6 How is this problem of function solved? In many instances this begins with a set of identifying criteria and the archaeological remains are matched against the criteria. A reasonable fit results in a conclusion of function. There are numerous methodological com- ponents to this process, one of which is addressed here: establishing criteria of function with a case example of the refuge cave, a modified lava tube. The refuge cave, a place of wartime escape for non-combatants, is a commonly identified functional site type or model on the island of Hawai‘i, where lava tubes are found almost exclusively in the districts of Kona, Puna, and Ka‘u¯ (e.g., Bollt 2005; Bonk 1969; Cordy 2000; Kennedy and Brady 1997; Kirch 1985; Kolb and Dixon 2002). This identification is made primarily on the basis of entrances to lava tubes that have obstacles and/or constricted passageways, and secondarily on the possible concealment of entrances. These caves also, by the literature model, have interior evidence of habitation, usually structures and , but generally of limited quantity. Beyond this point, the model begins to disintegrate—a few

6This essay is a revision of the paper “Refuge Caves” of Kona: A Reconsideration of the Pre-Contact Hawaiian Refuge Cave Model by M.J. Tomonari-Tuggle and Dave Tuggle, presented at the Annual Conference of the Society for Hawaiian Archaeology, Kona, Hawai‘i, 2004. Special thanks to Tom Wolforth, MaryAnne Maigret, and Keola Lindsey for providing perspective. 168 Dave Tuggle cases report some weapons, but this is rare; in some cases refuge caves are described as being components of habitation complexes, but there are other cases where this is not true. Some archaeologists have interpreted refuge caves as used by commoners (e.g. Cordy 2000), and others have interpreted them to have been for elite use (e.g. Kennedy and Brady 1997). In most discussions of refuge caves, the identification of “refuge” is frequently tautologically inherent in the description of the cave and its structures. Such statements as “construction of features within the tube [were] to enable defense” (Schilt 1984, 117) or that a mound of stones at a cave entrance is a “refuge barrier” (Schilt 1984, 102) are typical. In other words, artificially constricted cave entrance = refuge cave has become a conventional interpretation. Of course, all disciplines have a body of equivalent knowledge regarding the accepted understandings of the world, but these are usually developed by means of scientific investigation, and the relevant literature is cited when such points are important in a larger argument. I had employed this simple equation more than once before realizing that I knew no source in the archaeological literature that could be cited as a definitive argument for the functional equation constricted cave entrance = refuge cave. Some efforts have been made to find an ethnographic reference to refuge caves (e.g. Bollt 2005; Schilt 1984), but no unequivocal examples have been found. One paper that seems to head in the direction of an explicit evaluation of the refuge cave model is that of Kennedy and Brady (1997) who refer to the broad cross-cultural significance of caves for ritual and other non-refuge activities—with implications that not all modified caves are for refuge, but inexplicably the paper drops this perspective when specifically describing Hawaiian caves. So why do we archaeologically categorize these caves with arti- ficially constricted entrances as wartime refuge caves? What is the origin of the refuge model and why have no alternative models been seriously evaluated? In reviewing the literature, a distinction needs to be made between unmodified caves used for hiding for various reasons and caves modified with defensive features that constitute the identified function of wartime refuge. The historical and ethnographic literature includes a few refer- ences to the use of caves for hiding. Ellis refers to a cave in Kona used as a hiding place during war7 and in another example mentions

7Ellis’s remarks on the Kona cave are discussed on page 170. Lady Mondegreen’s Hopes and Dreams 169

a “small cave” where, after the battle of Kuamo‘o, a woman hid after “drawing a piece of lava over its mouth” (Ellis 1963, 78). Two other examples of hiding in non-modified caves are from Monsarrat (cited in Summers 1971, 192) and Pukui (1983, 291), and one general statement about cave hiding is in Handy and Pukui:

Deep and long caves with several openings, in the upland and at the beaches, with a good draught in places where smoke of fires would be quickly dispersed by the wind, were used as places of refuge in time of war. There were places inside of these where sweet potatoes were stored in preparation for just such happenings. These were kept renewed from time to time, so that when war came, all could find refuge and sustenance. Fuel was kept in readi- ness near fireplaces, and salt for the fish or meats. The locations of these caves of refuge were not talked about to outsiders. (Handy and Pukui 1972, 14)

The lesson we learn from this and similar cases in the historical literature is that “hiding in lava tubes was well-established in the popular mind” (Bollt 2005, 15). However valuable this may be, it tells us little about refuge caves modified for wartime use, no references to which have been found in the ethnographic or historical sources. The earliest archaeological reference to caves as hiding places found to date is 1933, when Emory mentions that “[i]n Kona, at Ka- haluu, miles of underground lava tubes served as refugee [sic] caves or hiding places in times of war” (Emory 1999, 233), but there is no reference to fortified entrances. No references to modified caves as refuge sites have been found in the early archaeological survey re- ports from Kona (e.g. Reinecke 1930; Stokes 1991), but refuge caves are recorded as sites in the reports of the Bishop Museum in the 1950s (e.g. Bonk 1969; Emory 1970; Kelly 1986). Kelly mentions “caves of refuge,” and describes them as follows:

In areas where large lava tubes are found, Hawaiians some- times used the particularly well-located tubes in which to hide from pursuers. They sometimes built stone walls across the lava tube caves near the entrances, allowing space for passage of only one person at a time . . . these caves were mainly defenses . . . (Kelly 1986, 137) 170 Dave Tuggle

In a manuscript written in the late 1950s, Bonk describes a cave in the South Point area:

Access was . . . limited to one person at a time, which as- sured occupants of an easily defended position. A small number of men, young or old, or perhaps even women, would have had little difficulty defending such an unas- sailable position with little more than a club, a spear, or large stone. The individual who found himself forced to crawl or pull himself through the narrow entryway was completely at the mercy of any or all who may have waited within, beyond the opening. (Bonk 1969, 75)

On the basis of available information, it appears that these state- ments represent archaeological inference. No ethnographic or his- torical examples are cited, and there is also no indication that other explanations were considered. Detailed museum catalog descriptions from the 1950s have yet to be consulted, but it appears that the equa- tion of artificially constricted cave entrance = refuge cave had become a given by this time. It would seem that this equation had the appeal of a logical tem- plate to archaeologists working in Hawai‘i, so much so that it endured for another 50 years with no more than a fleeting consideration of alternatives (Kennedy and Brady 1997). The appeal was strong enough that archaeologists seem willing to take any evidence of hiding in caves as support for the equation, suggesting that it had become a ruling hypothesis. For example, Schilt writes: “Ellis . . . reported that refuge caves had been known as places of hiding, functioning as forts during warfare and in times of social conflict” (Schilt 1984, 38). Bollt says somewhat more candidly: “The closest description of a refuge cave that we possess also comes from Ellis” (Bollt 2005, 102). An examination of the Ellis descriptions suggest that he does not provide support for the archaeological refuge cave. Most citations to Ellis refer to the cave of Lani¯akea, in Kailua, which was an “appendage” to the massive wall he describes as a fort. Ellis remarks:

In this cavern, children and aged persons were placed for security during an assault or sally from the fort . . . They [the natives] asserted . . . that the cavern, if not the fort also, was formerly surrounded by a strong palisade. (Ellis 1963, 62) Lady Mondegreen’s Hopes and Dreams 171

In Kelly and Barrère it is noted that this place described by Ellis “must have been an ancient pu‘ukaua and pu‘uhonua (fortification and refuge)” (Kelly and Barrère 1980, 16). Kelly and Barrère do not clarify their interpretation of this as both fort and place of refuge, versus Ellis’s use of the term fort rather than pu‘uhonua, a term and concept he well knew and described in detail for Honaunau¯ (Ellis 1963, 113). Summers (1971, 92) suggests that a place could be used both as a pu‘ukaua and pu‘uhonua, and quotes Kamakau in support. However, the discussion of pu‘ukaua and pu‘uhonua in Kamakau (1964, 17, 18) is less than edifying, and it is not clear that he means what Summers takes him to mean. If one reads the total description in Ellis, it could very easily be understood as a pu’uhonua like that of Honaunau,¯ but there is no evidence that anyone told Ellis what it actually was, that is, fort and/or pu‘uhonua, and he may have simply deduced his interpretation. If Ellis’s drawing of the great wall at this place is taken literally, it is a large two-sided enclosure very similar to the enclosing pu‘uhonua wall at Honaunau.¯ The embrasures may be simply collapsed areas or this could in fact have been a traditional structure modified to hold canons during the time of Kamehameha. Thus, this may have been a traditional pu‘uhonua, abandoned during the time of Kalani‘opu‘u¯ and re-fitted as a military fort under Kamehameha, or, parsimoniously, this was simply a post-contact fort. Be that as it may, if this is “the closest description of a refuge cave that we possess” (Bollt 2005, 102) then we possess no description of a refuge cave that matches the archaeological site type. The cave had no crawlway entrance, and surrounded by a palisade and associated with a fort, it could hardly have been hidden. This Lani¯akea property was given to the missionary Thurstons in 1823, and it has been assumed that the stones of the fort were used to build the enclosure for the Thurston house lot—archaeological investigations have not identified any evidence for cave modification (Kelly 1983, 12).8

Hawaiian Terminology For Caves And Refuge

The term pe‘e pao has been noted in recent archaeological reports as the Hawaiian term for hiding in a cave during warfare and thus as a reference to refuge caves (e.g. Kolb and Dixon 2002, 526).9 The

8I thank Tom Wolforth for bringing this to my attention during discussions in 2004 on cave modifications. 9For further consideration of hiding as shameful, see Tregear (1891, 332). 172 Dave Tuggle

Hawaiian Dictionary translates pe‘e pao as “to hide in a cave,” or “secret cave,” as well as “term of contempt for vagabonds or cowards” (Pukui and Elbert 1971). There is nothing in the translation that suggests a wartime refuge cave and it also has the connotation of shameful hiding. In archaeological reports written in the 1950s, the term ana pu‘uhonua is used for refuge cave (e.g. Bonk 1969; Kelly 1986). When the Kelly report was re-published in 1986, a footnote was added based on comments from Mary Pukui, who said that the term ana pu‘uhonua is known, but that the accepted term for refuge cave is “ana pe’e kaua” [not pe‘e pao] (Kelly 1986, 137)—that is, a “war hiding cave,” without the implication of pu‘uhonua. Neither of these phrases can be found in Hawaiian dictionaries (e.g. Andrews 1865; Pukui and Elbert 1971), although the 1865 Andrews dictionary’s definition of ana includes “cave for the retreat of the vanquished; a place where the conquered are found.” Again, this does not suggest a refuge cave consistent with the archaeological model, but rather a post-battle hiding place and, one would assume, the accompanying shame. Finally, in Tregear (1891, 10), under the Hawaiian citation for ana is the sentence “O ka poe maloko o na pakaua a me na ana; Those which are in the forts and caves.”10 The implication is that forts and caves are the two places where people take refuge, presumably during war.

Evaluation of the Refuge Model: The Practical Questions

Terminology, ethnographic information, and historical accounts all indicate that seeking safety by hiding in caves was an element of Hawaiian warfare. There are no references in this material to indicate that such caves were fortified or artificially concealed, but that ques- tion aside, what is the inherent validity of the model of the refuge cave? The most explicit ethnographic statement about caves used as wartime refuge is from Handy and Pukui (1972, 14), quoted on page 169, that provides some criteria for a cave to be used for wartime survival. Few of the archaeologically-identified refuge caves meet these criteria: many have no air flow, are very hot, have no water supply, and lack secondary escape routes. Most are not places where one could survive for any length of time.

10The original source of the sentence is unknown. Lady Mondegreen’s Hopes and Dreams 173

Of the two main identification criteria in the archaeological model, defensive construction of the entryway and entrance concealment, the effectiveness of the latter is hard to evaluate, but concealment as defense is certainly reasonable. However, a crawlway or constricted passage is a dubious means of effective defense. One of the definitive expressions of the defense argument is that of Bonk describing access as “limited to one person at a time, which assured occupants of an easily defended position,” (Bonk 1969, 75).11 This is not a credible argument from a practical standpoint. It is unlikely that a group of armed warriors pushing through a constricted entrance would be deterred for long. But even assuming that warriors could be held back, once a cave as a hiding place has been discovered, it ceases to be protection and instead becomes a trap. People inside caves can be routed by fire and smoke, or simply out-waited and allowed to suffer from hunger or thirst. A Moloka‘i example described by Monsarrat in 1884, one of the few specific historical instances of a cave used for refuge, is instructive:

a lot of war canoes from Hawaii were coasting along the shore and when passing the mouth of the cave [Ananalu- awahine, Kalawao] at sea saw two old women sitting in the cave. Hawaii men built a fire over the cave on the flat above and put water in so that it caved in. The cave at that time was filled with the people of Kalaupapa and Kalawao who were hiding from the Hawaii men. After making a hole into the cave the Hawaii men entered and killed all the people in the cave. (quoted in Summers 1971, 192)

Alternative Models of Caves with Constricted Entrances

The ethnographic and historical literature certainly indicates that people did, on occasion, hide in caves for refuge during warfare. The ethnographic and historical literature also has information on cave use for habitation, burial, religious practices, and collection of resources such as water. Caves, or underground passages, are also prominent in traditions regarding the mythopoeic view of the world. However, there is no information to indicate that any of these activities or beliefs are associated with cave entrance modification consistent with the archaeologically identified refuge cave. Thus, to determine the

11This passage is quoted in full on page 170. 174 Dave Tuggle purpose for the construction of the constricted entrances is a matter of inference based on archaeological data and cultural context. Of the six possible purposes noted above, which are not neces- sarily exhaustive, five can be modeled to match the elements of the archaeological refuge cave.12 A constricted, and possibly concealed, entrance could serve: (i) to protect people during warfare, as the prevailing equation posits; (ii) to protect burials; (iii) to provide a hidden and/or symbolic setting for religious rituals, or a place to hide ritual paraphernalia and images; (iv) to represent the presence of a deity or event associated with a deity; and (v) to hide and protect a resource. There are a number of general considerations for alternative models. A crawlway may serve as a practical defense and as symbolic entrance for chambers dedicated to burial, religious ceremony, or ritual object storage. A crawlway, it can be argued, may be ineffective against an army, but serve as a reasonable defense against one or two bone thieves or intruders. The symbolism of a passageway into earth is common cross-culturally as access to the underworld or the afterlife, or as representations of the womb of Mother Earth (Campbell 1970, 160; Walker 1983, 154–156; Baring and Cashford 1991, 18–20), and may have been conceived as such in Hawai‘i. Most of these traditional Hawaiian references are to caves in the ocean associated with sharks, the underworld, death, and the deity Kanaloa (e.g., Tava and Keale 1989, 21, 74; Valeri 1985, 15; Beckwith 1970, 61). However, some of these caves also open at various places on land, and there is traditional information on terrestrial caves that pro- vides some suggestion of possible meanings, which include hiding and protection and sometimes associated sorcery (e.g., Beckwith 1970, 218, 244, 263; see Robins et al. 2000), and association with female and birth, particularly in the specific combination of Ku¯ with Hina, and in the form of the mother of Moloka‘i. A cave on Moloka‘i is called Keanaohina, and is the kumu (root) of the island, and the place of residence of the deity mother of Moloka‘i, known as Moloka‘inuiahina and also Hinalauae (Summers 1971, 127; Beckwith 1970, 219). Handy and Pukui write:

Believed to be local in origin were other forebears [in addition to Pele and her family]: that one from whose navel grew a great gourd vine, originating in a certain cave, which spread over and peopled seven districts in Ka‘u¯ . . . (Handy and Pukui 1972, 27)

12The one exception is that of daily habitation. Lady Mondegreen’s Hopes and Dreams 175

Pele herself was continually renewed by coming from openings in the body of Haumea, and the cult of Pele is a likely candidate as one group that might hold cave rituals. If the elaborately constructed entrances of these caves are associ- ated with burial or ritual activity, the caves may be more appropriately called ana huna, a secret cave, or cave where things are hidden (R. Johnson, pers. comm. 2004) rather than ana pu’uhonua, ana pe’e kaua, or pe’e pao. A full review of these alternative models, not attempted in this brief essay, involves an evaluation of the archaeological cases of caves with constricted entrances. But, unfortunately, like so many sites, most caves have been looted of the artifacts that would most inform on the specific and perhaps unique activities within them—artifacts such as weapons, deity images, and burial items. This is an element of the archaeological reality that archaeologists in Hawai‘i must address in solving methodological problems.

Conclusion

It is argued here that the constricted entrance = refuge cave equation is a conventional interpretation, and not an inference based on anal- ysis of relevant data. There is no convincing evidence to make this equation any stronger than other models to explain the archaeological features of such caves, including protection and symbolism associated with burial, ritual, and/or the activities of deities.

Glass: The Suspension of Disbelief

[I]t was agreed, that my endeavours should be directed to persons and characters supernatural, or at least romantic, yet so as to transfer from our inward nature a human interest and a semblance of truth sufficient to procure for these shadows of imagination that willing suspension of disbelief for the moment, which constitutes poetic faith.

Samuel Taylor Coleridge (Biographia Literaria, 1817)

Archaeologists often treat chronometric dates as facts about the past, when, in fact, they are inferences based on the facts of instru- ment data, and may or may not be valid statements about the past. The dates are also probabilistic statements. In both dimensions, the inferential and the probabilities, the methodological bridges to those statements about the past are ignored. We struggle to make sense 176 Dave Tuggle out of 14C dates—hoping that something actually happened at a site within that 200 year span that glares at us from a Beta Analytic print- out. For sites of special importance we keep reviewing the 14C dates, and proveniences, and labs, and material, and calibrations, and sta- tistical methods—hoping that in the end we have some semblance of historical truth (Dye 1992; Kirch and McCoy 2007; Tuggle and Spriggs 2000). Within our limited reach of historical empiricism, we probably do have some semblance of historical truth. But there is another case of so-called chronometric age determination in the history of Hawaiian archaeological investigations that, in retrospect, consisted of nothing but hopes and dreams: this is the history of hydration-rind dating of Hawaiian volcanic glass. Hawaiian hydration-rind dating flourished for some 15 years, from 1970 to 1985, then fell into disuse. The 1985 publication of the now classic introduction to Hawaiian archaeology, Feathered Gods and Fishhooks included some hydration-rind dates, but did so with reser- vations (Kirch 1985, 51) that included the citation of two studies that questioned this dating method (Olson 1983; Tuggle and Olson 1978). Hydration dating of obsidian was developed in 1960 (Friedman and Smith 1960) and in 1969 the concept was applied to Hawaiian volcanic glass (Morgenstein and Riley 1974) and a dating service emulating radiocarbon labs was created, called Hawaii Marine Research, Inc. This was accepted with great enthusiasm by archaeologists working in Hawai‘i and began to be employed nearly universally. As Kirch notes, this enthusiasm was such that

many archaeologists virtually abandoned the use of radio- carbon dating for sites excavated in the mid-to-late 1970s, relying almost exclusively on the volcanic-glass dating method. (Kirch 1985, 51)

As one of those glassy-eyed enthusiasts, I not only employed glass-hydration dating as my main source of absolute chronometry for Hawaiian sites, albeit with a few backup radiocarbon dates, but I also began a research program at the University of Hawai‘i-M¯anoa to provide baseline data, such as source locales and geochemical variation, for volcanic glass research. However, in the process of working with Hawaiian volcanic glass it became clear that there were serious problems with hydration dating, not the least of which was measurement of the hydration rinds. This also led to a review of the research that lay behind the practice of dating of Hawaiian volcanic Lady Mondegreen’s Hopes and Dreams 177 glass, including the empirical basis for the hydration rate then being used. The results of this background study were summarized in one primary paper (Tuggle and Olson 1978), never published, with some additional information published in a later paper (Olson 1983). The geochemical research reviewed at that time indicated three important points: (i) the nature of the hydration process was a matter a considerable uncertainty; (ii) proposing and promulgating a rate of hydration was premature; and (iii) the proposed rate was based on a very small data set and a questionable inferential chain (Tuggle and Olson 1978, 7–18).13 This is three ways to say that the hundreds of hydration-rind “dates” generated in Hawaiian archaeology are nothing but scientific noise. As far as I know, no additional volcanic glass dates have been generated from Hawaiian specimens since the mid-1980s, and after Kirch (1985) volcanic glass dates in the previous literature have been rarely cited. How did most archaeologists in Hawai‘i so quickly come to em- brace volcanic glass hydration dating? Kirch provided the following assessment:

it promised to be an extremely valuable chronometric tool. Volcanic-glass flakes are ubiquitous in most archaeological contexts . . . the technique is relatively inexpensive . . . and also appeared to be quite accurate. (Kirch 1985, 51)

Other factors were involved as well. The dates were reported by the laboratory with a very small error range, much smaller than that of the average radiocarbon date, so that the dates appeared to be precise. In addition, the target event, the date of site activity, could be inferred to be much more closely associated with the dated event, which was the flaking of a piece of volcanic glass, than with the radiocarbon- dated event (death of the plant or animal)—there was no “old wood” problem. What appeared to be almost too good to be true proved in re- ality to be absolutely so. To rephrase the previous question, how did archaeologists uncritically perceive promise as realization? The most immediate answer is suggested in the nature of the Hawaiian archaeological record, which offers so very little for chronological

13The very small data set is two cases of deep-sea palagonitization—an altered form of a basaltic glass known as sideromelane, involving millions of years and huge error factors in the geological dating. The inferential chain involved generation of a rate and applicability to Hawaiian terrestrial volcanic glass. 178 Dave Tuggle construction: no meaningful stylistic artifact temporal variability, shallow sites that are often poorly stratified, and a relatively short time-span of island and site occupation. This is a precondition for the infomercial appeal of a quick and cheap fix. At the same time, as Tom Dye has suggested in a personal commu- nication, this must also be understood as a consequence of the general methodological vacuum of Hawaiian archaeology where radiocarbon dates were being interpreted, as they still are, with no coherent and community-wide methodological framework. The fitting end to this story is that that was the end of the story. Research on volcanic glass hydration as a dating method has contin- ued in other places (e.g. Liritzis et al. 2008; Mulrooney et al. 2009), but to my knowledge, not in Hawai‘i.

Conclusion

The larger context for understanding these three essays—the little historical episode of the rise and fall of hydration dating, the matter of the conventional interpretation of cave configuration, and the unrealized potential of archaeoastronomy—is the history of Hawaiian archaeological research, where such limited attention has been paid to methodological problems, notably since about 1970, that is, the era of the National Historic Preservation Act and bureaucratized archaeology called cultural resource management. The nature of Hawaiian archaeological remains presents enormous methodological challenges, challenges that should be addressed as research problems if meaningful advances are to be made in our understanding of the past—not simply the continuous accumulation of data replicating the known. This brings two conclusions: first, that First Principles of Method- ology in Hawaiian Archaeology is a much-needed volume; second, that archaeological reports could be structured as models of the ar- chaeological process of inference—that is, structured as scientific presentations and not expositions in the humanities. Such an organi- zation would be explicit about the methodological foundation, where the missing pieces are in that foundation, and what the strengths and weaknesses are in the structure of the inferences. As a result, we should be able to recognize and discard the archaeological monde- greens before they become part of conventional descriptions of the past. Lady Mondegreen’s Hopes and Dreams 179

References

Andrews, L. (2003 [1865]). A Dictionary of the Hawaiian Language. Island Heritage, Waipahu, HI.

Baring, A. and Cashford, J. (1991). The of the Goddess: Evolution of an Image. Penguin, London.

Beckwith, M. (1970). Hawaiian Mythology. University of Hawaii Press, Honolulu.

Beckwith, M. W. (1940). Hawaiian Mythology. Yale University Press, New Haven, CT.

Bollt, R. (2005). Tricks, traps, and tunnels: A study of the refuge caves of Hawai‘i Island. Hawaiian Archaeology, 10:115–129.

Bonk, W. (1969). Lua Nunu o Kamakalepo: A cave of refuge in Ka‘u Kona. In Pearson, R., editor, Archaeology on the Island of Hawaii, number 3 in Asian and Pacific Archaeology Series. Social Science Research Institute, University of Hawaii at Manoa.

Bryan, E. H. and Crowe, R. A. (2002). Stars Over Hawaii. Press, Hilo, HI.

Campbell, J. (1970). The Masks of God: Creative Mythology. Viking Press, New York.

Chamberlin, T. C. (1965). The method of multiple working hypotheses. Science, 148(3671):754–759.

Cordy, R. H. (2000). Exalted Sits the Chief: The Ancient History of Hawai‘i Island. Mutual, Honolulu.

Dye, T. S. (1992). The South Point radiocarbon dates 30 years later. New Zealand Journal of Archaeology, 14:89–97.

Elliot, L. P. and Brook, B. W. (2007). Revisiting Chamberlin: Multiple working hypotheses for the 21st century. BioScience, 57(7):608–614.

Ellis, W. (1963). Journal of William Ellis. Advertiser Publishing Com- pany, Honolulu.

Emory, K. P. (1970). Inventory of Archaeological and Historical Sites in the Districts of Kona and Ka‘u and in Anaehoomalu, South Kohala, Island of Hawaii. Prepared for the Planning Department, County of Hawaii. Anthropology Department, B. P. Bishop Museum. 180 Dave Tuggle

Emory, K. P. (1999). Warfare. In Ancient Hawaiian Civilization. Mutual Publishing, Honolulu. First published 1933.

Fischer, D. H. (1970). Historians’ Fallacies: Toward a Logic of Historical Thought. Harper & Row, New York.

Friedman, I. and Smith, R. L. (1960). A new dating method using ob- sidian: Part I, The development of the method. American Antiquity, 25:476–522.

Handy, E. S. C. and Pukui, M. K. (1972). The Polynesian Family System in Ka-‘u, Hawai‘i. Charles E. Tuttle, Tokyo.

Hawkes, J. (1967). God in the machine. Antiquity, 41(163):174–180.

Hawkins, G. (1964). Stonehenge: A computer. Nature, 202:1258–1261.

Hawkins, G. (1965). Stonehenge Decoded. Doubleday, Garden City, NY.

Johnson, R. K. and Mahelona, J. K. (1975). Na¯ Inoa Hok¯ u:¯ A Catalogue of Hawaiian and Pacific Star Names. Topgallant, Honolulu.

Kaler, J. B. (1996). The Ever-Changing Sky: A Guide to the Celestial Sphere. Cambridge University Press, Cambridge.

Kamakau, S. M. (1964). Ka Po‘e Kahiko: The People of Old. Number 51 in B. P. Bishop Museum Special Publication. Bishop Museum Press, Honolulu.

Kamakau, S. M. (1976). The Works of the People of Old: Na Hana a ka Po‘e Kahiko. Number 61 in B. P. Bishop Museum Special Publication. Bishop Museum Press, Honolulu. Translated from the Newspaper Ke Au ‘Oko‘a by Mary Kawena Pukui. Arranged and edited by Dorothy B. Barrère.

Kelly, M. (1983). Na Mala o Kona: Gardens of Kona, A History of Land Use in Kona, Hawai‘i. Number 83–2 in Departmental Report Series. Anthropology Department, B. P. Bishop Museum, Honolulu.

Kelly, M. (1986). The concept of asylum. In Bryan, Jr., E. H. and Emory, K. P., editors, The Natural and of Honaunau,¯ Kona, Hawai‘i, number 86–2 in Departmental Report Series, report 11. Anthropology Department, B. P. Bishop Museum, Honolulu. Lady Mondegreen’s Hopes and Dreams 181

Kelly, M. and Barrère, D. (1980). Background History of the Kona Area, Island of Hawai‘i. Prepared for Department of Transportation, State of Hawai‘i. Department of Anthropology, B. P. Bishop Museum, Honolulu.

Kennedy, J. and Brady, J. (1997). In the netherland of island Earth: A re-evaluation of refuge caves in ancient Hawaiian society. Geoar- chaeology, 12:641–655.

Kirch, P. V. (1985). Feathered Gods and Fishhooks: An Introduction to Hawaiian Archaeology and Prehistory. University of Hawaii Press, Honolulu.

Kirch, P. V. (2004). Temple sites in Kahikinui, Maui, Hawaiian Islands: Their orientations decoded. Antiquity, 78(299):102–114.

Kirch, P. V. and McCoy, M. D. (2007). Reconfiguring the Hawaiian cultural sequence: Results of re-dating the Halawa Dune Site (MO-A1- 3), Molokai Island. Journal of the Polynesian Society, 116(4):385–406.

Kolb, M. J. and Dixon, B. (2002). Landscapes of war: Rules and con- ventions of conflict in ancient Hawaii (and elsewhere). American Antiquity, 67(3):514–534.

Kyselka, W. (1991). Solar in Hawai‘i. Hawaiian Journal of History, 25:31–51.

Liritzis, I., Laskaris, N., and Bonini, M. (2008). Nano- and micro-scale resolution in ancient obsidian artefact surfaces: The impact of AFM on the obsidian hydration dating by SIMM-SS. Physica Status Solidi, 5(12):3704–3707.

Loose, R. (2008). Tse’biinaholts’a yałti (curved rock that speaks). Time and Mind, 1:31–49.

Makemson, M. (1941). The Morning Star Rises: An Account of Polyne- sian Astronomy. Yale University Press, New Haven, CT.

Morgenstein, M. and Riley, T. J. (1974). Hydration-rind dating of basaltic glass: A new method for archaeological chronologies. Asian Perspectives, 17:145–159.

Mulrooney, M. A., Ladefoged, T. N., Stevenson, C. M., and Haoa, S. (2009). The myth of ad 1600: New evidence from Honga Ho‘onu, Rapa Nui. Rapa Nui Journal, 23(2):94–105. 182 Dave Tuggle

N¯amakaokeahi, B. K. (2004). The History of Kanalu. First People’s Productions, Honolulu.

Olson, L. (1983). Hawaiian volcanic glass applied “dating” and “sourc- ing”: Archaeological context. In Clark, J. T. and Kirch, P. V., edi- tors, Archaeological Investigations of the Mudlane-Waimea-Kawaihae Road Corridor, Island of Hawai‘i: An Interdisciplinary Study of an En- vironmental Transect, number 83–1 in Departmental Report Series, pages 325–340. Anthropology Department, B. P. Bishop Museum, Honolulu.

Pukui, M. K. (1983). Olelo Noeau: Hawaiian Proverbs and Poetical Sayings. Number 71 in B. P. Bishop Museum Special Publication. Bishop Museum Press, Honolulu.

Pukui, M. K. and Elbert, S. H. (1971). Hawaiian Dictionary. University of Hawaii Press, Honolulu, third edition.

Reinecke, J. E. (1930). Survey of sites on West Hawaii. Typescript in State Historic Preservation Division Library, Kapolei, HI.

Robins, J. J., Colins, J., Chiogioji, R., Carlson, I., Masterson, I., and Ham- matt, H. H. (2000). An Archaeological Inventory Survey of an Ap- proximately 803-Acre Subject Parcel in the Ahupua‘a of Honokaohau¯ I and II, North Kona District, Island of Hawai‘i. Prepared for Lanihau Partners. Cultural Surveys Hawai‘i, Kailua, HI.

Scarre, C. and Lawson, G., editors (2006). Archaeoacoustics. McDonald Institute for Archaeological Research, Cambridge.

Schilt, R. (1984). Subsistence and Conflict in Kona, Hawai‘i: An Ar- chaeological Study of the Kuakini Highway Realignment Corrridor. Number 84–1 in Departmental Report Series. Anthropology Depart- ment, B. P. Bishop Museum, Honolulu.

Shapiro, J. S. (2010). Contested Will: Who Wrote Shakespeare? Simon and Schuster, New York.

Silva, A. M. D. and Johnson, R. K. (1982). Ahu a ‘Umi Heiau: A native Hawaiian astronomical and directional register. Annals of the New York Academy of Science, 385:313–331.

Sterling, E. P. and Summers, C. C. (1978). Sites of Oahu. Departments of Anthropology and Education, B. P. Bishop Museum, Honolulu. Lady Mondegreen’s Hopes and Dreams 183

Stokes, J. F. G. (1991). Heiau of the Island of Hawai‘i: A Historic Survey of Native Hawaiian Temple Sites. Number 2 in Bishop Museum Bulletin in Anthropology. Bishop Museum, Honolulu. Edited and introduced by Tom Dye.

Summers, C. C. (1971). Molokai: A Site Survey. Number 14 in Pacific Anthropological Records. Anthropology Department, B. P. Bishop Museum, Honolulu.

Tava, R. and Keale, M. K. (1989). Ni‘ihau: The Traditions of an Hawaiian Island. Mutual, Honolulu.

Tregear, E. (1891). The Maori-Polynesian Comparative Dictionary. Whit- combe and Tombs, Christchurch, NZ.

Tuggle, H. D. and Olson, L. (1978). A review of “hydration dating” of Hawaiian volcanic glass. Typescript on file, Historic Preservation Division, Department of Land and Natural Resources, Kapolei, HI.

Tuggle, H. D. and Spriggs, M. (2000). The age of the Bellows dune site, O18, O‘ahu, Hawai‘i, and the antiquity of Hawaiian colonization. Asian Perspectives, 39(1–2):165–188.

Tuggle, H. D. and Tomonari-Tuggle, M. J. (2006). Archaeological Overview for Haleakala¯ National Park, Island of Maui. Prepared for National Park Service. International Archaeological Research Institute, Inc., Honolulu.

Valeri, V. (1985). Kingship and Sacrifice: Ritual and Society in Ancient Hawaii. University of Chicago Press, Chicago.

Walker, B. G. (1983). The Woman’s Encyclopedia of and Secrets. Harper, San Francisco.

Abstract

This paper addresses the point that archaeological data can be organized spatially and temporally into many patterns, and each pat- tern can be interpreted as a result of multiple possible behaviors. It is asserted that Hawaiian archaeology does not have a set of methods established through rigorous methodological research that allows the multiple possibilities to be identified and evaluated, and that as a re- sult reports provide conclusions that sound coherent largely because they are expressed in conventional language. The paper provides three essays related to the methodological foundations of research. The first discusses the potential of astronomy in Hawaiian archaeology—the use 184 Dave Tuggle

and neglect of a historical known. The second reviews the interpre- tation of refuge cave as a ruling hypothesis. The third examines the use of volcanic glass dating in Hawaiian archaeology as a failure of methodological interest. Index

‘Alae Iki, Maui, 119, 128–131, 142, unreduced, arguments for, 145 14–16 14C dating, 44, 46, 47, 70, 99, zones, 7–9, 12, 13, 17, 18, 100, 110–112, 118, 119, 40–43, 50–52 123, 125, 129, 131, 133, agricultural intensification, 4, 135, 143–146, 176 9, 38, 52 composition, 5, 47, 50, 133, Ahu a ‘Umi Heiau, 163 134 Anahulu Valley, O‘ahu, 20 expansion events, 44–47 aqueduct, Waimea, Kaua‘i, 14, interpretation, 94 16 processing, 49 archaeobotanical analysis, 48 provenance, 5, 46, 49, 50, archaic state, 3, 14, 27, 28, 36– 111 38, 52, 53 230 Th coral dating, 46, 100, 112, bureaucracy, 31 146 definition, 27–28 Hawaiian, 28–30, 32 adze C7727, 69 bureaucracy, 29, 30, 34, agricultural expansion, 2, 4–14, 36, 38, 39, 53 16–20, 22–25, 27, 37– large-scale, 30 41, 44–48, 50–52, 58, stratified control hierar- 60, 61, 92 chy, 29, 34, 35, 39 contiguous, 7 territorial units, 34 discontiguous, 6, 8 primary, 3, 28, 31 phases 1 and 2, 5, 10, 13, stratified control hierarchy, 23, 24, 38–40, 50, 51, 31, 32, 52 58 astronomy, 160–164, 166, 183 significant reduction in, 3, 4, 9, 18 stages, 7–9, 12, 51 Bayes, Thomas, 114

185 186

Bayesian statistics, 107, 110, 112– Dye-Komori curve, 23, 60, 142– 114, 117, 119, 131, 137, 144 145, 146 revised version, 5, 6, 23, BCal software package, 95, 112, 25, 37, 38, 53, 56 117, 118 Boötes, 166 Earth, Mother, 174 botanical evidence, 45–47 El Niño-Southern Oscillation (ENSO), bottle gourd (Lagenaria sicer- 71–73 aria), 70 weather anomalies, 72 breadfruit (Artocarpus altilis), weather conditions, 72 70 Ellis, Reverend William, 129 Brigham, William T., 96 Ellis, William, 160, 171 Bryan, E. H., 162 environmental limits on agri- burial, 173–175 cultural expansion, 17– 22 census aridity, 17, 47 Hawai‘i, 53–57 drought risk, 18 Marquesas Islands, 55, 57 filled alluvial lands, 19–20 Tahiti, 55, 57 flood-water systems, 20– Central East Polynesian long- 22, 38, 41, 42, 51, 52 distance exchange net- leached soils, 18 work, 68–69 slopes without systems, 18– Chenopodium, 46, 48, 49 19 chiefly ranks, 32–34 equifinality, 158 colonization, late Polynesian, ‘Ewa Plain, O‘ahu, 6n, 23, 59–61 see late Polynesian col- exchange network, 69 onization expansion, agricultural, see agri- conquest warfare, 3, 25, 27, 30, cultural expansion 36–39, 52 control, population, see popu- genealogical chronology, 5n, 62n, lation control 63, 65, 69 Cook, Captain James, 140 genealogical series Cordy, Ross H., 141 Hawaiian and Maori, 64 cores, salubrious, see salubri- Green, Roger C., 94 ous cores corvée labor, 2, 30, 39, 53 Hok¯ u‘iwa,¯ 166 Crowe, R. A., 162 Honaunau,¯ Hawai‘i, 171 H¯alawa Valley, Moloka‘i, 21 dating, volcanic glass, see vol- Haleakal¯a, Maui, 164, 165 canic glass dating Halele‘a District, Kaua‘i, 14–16 Dye, Thomas, 167, 178 Halley’s comet, 162 Index 187

Hamlet, 158 Ku,¯ 174 Harris Matrix, 95, 101, 107–110, Ka‘u¯ District, Hawai‘i, 167 119 Ka‘ena, O‘ahu, 164 Haumea, 175 Kahalu‘u, Hawai‘i, 98, 131, 137 Hawaiian Biocomplexity Project, Kahiki, 166 13, 15, 18, 42 chiefs, 62–64, 96 Hawkes, J., 166 connection, 61 Hawkes, Jacquetta, 161 definition, 61 Hawkins, Gerald, 161 Kahikinui District, Maui, 10, 13, Hikinaakal¯a Heiau, 164 17, 18, 23, 59–61, 163 Hina, 174 Kaho‘olawe Island, 27, 42, 47, Hinalauae, 174 48, 69 house-count method, 142–144 Kaiholena, Hawai‘i, 103, 119, 120, Hual¯alai, Hawai‘i, 163, 166 124, 129, 130, 140, 141, Hume, David, 115 144 Kalani‘opu‘u,¯ 171 index, Late Expansion (LEX), see Kalauniu‘ohua¯ (Hawai‘i Island Late Expansion (LEX) chief), 137–140 Index Kalaupapa, Moloka‘i, 19, 21, 23 indirect evidence for limits on Kaler, James, 161 agricultural expansion, Kamakau, Samuel M., 160, 165, 22–39 171 adze production, 22 Kamehameha, 138, 171 emergence of archaic states, Kanaloa, 174 27–39 Kawela, Moloka‘i, 18, 21 famine foods, 22 Kealakekua, Hawai‘i, 23, 60, 61 population control, 24 Kekaha kalana, 18, 34 population decline, 23 Kekahuna, Henry E. P., 132, 133 population growth rule of Kirch, Patrick V., 97, 98 thumb, 23, 58 Kohala Field System, Hawai‘i, roots of war, 24–27 6–13, 17, 18, 40–45, 51, intensification, agricultural, see 70 agricultural intensifi- Kona District, Hawai‘i, 167 cation Kona Field System, 18, 30, 43 inventory survey, archaeologi- Ku-k¯ ¯a‘ili-moku, 25 cal, 95, 103, 119, 141, Kula District, Maui, 22 145, 147

Johnson, Rubellite, 160, 162 Ladd, Edmund J., 97 Lani¯akea, Hawai‘i, 170, 171 K¯aneloa Heiau, 164 Lapakahi, Hawai‘i, 129 188

Late Expansion (LEX) Index, 10– Marin, Francisco de Paula, 138 12, 39, 40, 50, 51 Masse, Bruce, 160 threshold value, 10, 12, 40, Mauna Kea adze quarries, 22, 41, 50–52 51 variables, 10–11, 40, 41 McCartney, Paul, 157 late Polynesian colonization, 66– Menzies, Archibald, 160 68 method, 159, 176, 178 Easter Island (Rapa Nui), methodology, 158–162, 167, 175, 68 178, 183, 184 Enderby Island, 67 Migratory Period, 62, 63 New Zealand (Aotearoa), 66 Moka‘ena Heiau, 164 Norfolk Island, 67 Moloka‘inuiahina, 174 Raoul Island, 67 mondegreen, 158, 178 Rapa, 67–68 Monsarrat, M. D., 169, 173 Late Voyaging Period, 36, 62, Murakami, Gail, 133 66, 68–73 N¯amakaokeahi, B. K., Lennon, John, 157 162 Necker Island, 21 Libby, Willard, 94 network Limited Land Hypothesis, 3–6, exchange, see exchange net- 9, 10, 12–14, 20, 22, work 23, 25, 27, 36, 38–41, small-world, 31 48–50, 52, 58, 59, 61 social, see social network confirmation, 12, 13, 20, 40, Ni‘ihau Island, 22, 23, 35 41, 51, 59 Nihoa Island, 21 reporting, 49 North H¯alawa, O‘ahu, 94, 107 testing, 39 variables, 39 OxCal software package, 95, 117, 143 M¯akaha Valley, O‘ahu, 97 Mahele¯ records, 15 Pele, 164, 174, 175 Mahelona, J. K., 162 person-year, 58–60 Major, Maurice, 103, 121 person-years Makaha Valley, O‘ahu, 21 defined, 59 Makemson, M., 162 Pleiades, 162, 163, 165 Malanui (Stokes informant), 137 Polonius, 158 Malo, David, 160 Polynesian leadership dimen- map sions, 35 detailed plan, 97–99 Polynesian lineage, erosion of, isometric projection, 98, 36, 62 99 Polynesian rat (Rattus exulans), oblique schematic, 97–99 63, 67 Index 189 population superposition, stratigraphic, see ancient Hawaiian maximum, stratigraphic superpo- 53 sition control, 14, 16, 22, 24, 52 sweet potato, 2, 11, 17, 18, 46, growth 47, 69, 70, 73 depopulation rate, 54 near-zero, 14, 23, 24, 58, Tahitic loans in Hawaiian, 65– 59, 61 66 rate equation, 9 taxes, 2, 3, 25, 29, 30, 39, 52, 53 history, 95, 142 terminus ante quem, 5, 58 size estimation, 140–142 volcanic glass dating, 100, 146, population growth 158, 176–178, 184 near-zero, 58 potato, sweet, see sweet potato Wa‘ahia (priestess), 137, 139, 140 Pu‘uomahuka Heiau, 163 Wailau Valley, Moloka‘i, 7, 9, 11, Pukui, Mary K., 172 42, 43, 51 Puna District, Hawai‘i, 167 Waimea, Hawai‘i, 23, 61 warfare, see also conquest war- rank, chiefly, see chiefly ranks fare, 14, 27 rat, Polynesian, see Polynesian warfare, conquest, see conquest rat warfare refuge cave, 158, 167–174, 184 weedy taxa, 46, 48 Rule of Six, 32, 34 salubrious core, 2, 8, 36 Schilt, Rose, 103 settlement pattern, diachronic, 94, 95, 142, 144, 147 Shakespeare, William, 158, 159 Shapiro, James, 158 Sinoto, Akihiko, 103 six, rule of, see Rule of Six social network, 31, 32 state, archaic, see archaic state Stokes, John F. G., 93, 96, 97, 99, 133, 137, 139 Stonehenge, 161 stratigraphic superposition, 96, 97, 108 Summers, Catherine, 171