<<

Geothermal systems in the Sunda volcanic

Investigations on the islands of and Bali,

Dissertation

zur Erlangung des Doktorgrads der Naturwissenschaften (Dr. rer. nat.) am Fachbereich Geowissenschaften Universität Bremen

vorgelegt von

Budi Joko Purnomo

Bremen March, 2015

ii

Gutachter: Prof. Dr. Thomas Pichler Prof. Dr. Peter La Femina

iii

iv

E r k l ä r u n g

Hiermit versichere ich, dass ich

i. die Arbeit ohne unerlaubte fremde Hilfe angefertigt habe, ii. keine anderen als die von mir angegebenen Quellen und Hilfsmittel benutzt haben und iii. die den benutzten Werken wörtlich oder inhaltlich entnommen Stellen als solche kenntlich gemacht habe.

______,den ______

______(Unterschrift)

v

vi

TABLE OF CONTENTS

Abstract ...... ix Zusammenfassung ...... xi I. Introduction ...... 1 I.1. Thesis outline ...... 2 I.2. Research objectives ...... 4 I.3. Theoretical backgrounds ...... 5 I.3.1. An overview of geothermal system ...... 5 I.3.2. Geothermal manifestations ...... 8 I.3.3. Geothermometry ...... 9 I.3.4. Boron isotope ...... 14 II. Geological setting, sampling and analyses ...... 17 II.1. Geological setting ...... 17 II.2. Sampling and analyses ...... 20 III. Geothermal systems on the island of Java, Indonesia ...... 23 Abstract ...... 24 III.1. Introduction ...... 25 III.2. Sampling locations ...... 25 III.3. Results ...... 26 III.4. Discussion ...... 30 III.4.1. General considerations about geothermal systems on Java...... 30 III.4.2. Geothermometry...... 37 III.4.3. The heat sources of the fault-hosted geothermal systems ...... 41 III.4.4. Oxygen and hydrogen isotope considerations ...... 44 III.5. Conclusions ...... 46 IV. Boron isotope variations in geothermal systems on Java, Indonesia ...... 49 Abstract ...... 50 IV.1. Introduction ...... 51 IV.2. Sampling locations ...... 52

vii

IV.3. Results ...... 52 IV.4. Discussion ...... 56 IV.4.1. Boron in thermal waters and seawater input ...... 56 IV.4.2. The δ11B of acid sulfate and acid chloride crater lakes ...... 60 IV.4.3. Processes affecting the δ11B value of thermal waters ...... 62 IV.4.4. The δ11B of fault-hosted and -hosted geothermal systems ...... 66 IV.5. Conclusions ...... 66 V. Geothermal systems on the island of Bali, Indonesia ...... 69 Abstract ...... 70 V.1. Introduction...... 71 V.2. Geological setting ...... 72 V.3. Results ...... 74 V.4. Discussion ...... 75 V.4.1. Geochemistry of thermal waters ...... 75 V.4.2. Phase separation and seawater input ...... 82 V.4.3. Geothermometry ...... 84 V.4.4. Oxygen and hydrogen isotope considerations ...... 86 V.5. Conclusions ...... 88 VI. Conclusions and outlook ...... 91 VI.1. Conclusions ...... 91 VI.2. Outlook...... 92 Acknowledgements ...... 95 References ...... 97

viii

Abstract

Geothermal systems located in a volcanic island arc are generally considered to be exclusively influenced by volcanic activities. Giving the complex geological setting on the island of Java, the geothermal systems probably is not that simple. The subduction of the Indo-Australian plate beneath the Eurasian plate formed three volcanisms, Paleogene, Neogene and Quaternary, and faults. The subduction also thinned the crust of the southern part of the island due to uplift and erosion. The presence of geothermal systems hosted by a volcano and a fault zone, thus named as volcano-hosted and fault-hosted, respectively, give an opportunity to study their different physicochemical characteristics. The influence of Quaternary magma in the fault-hosted geothermal systems distributed in two different volcanic belts, Quaternary and , were examined. Additionally, the contrast signatures of boron isotope between two different crater lakes, acid sulfate and acid chloride, which likely has been overlooked, were determined. Finally, the possibility of a carbonate sedimentary basement as the host-rock of geothermal systems on the volcanic island of Bali was investigated.

The physicochemical properties of geothermal waters were used in this study, including major anion and cation, trace elements and stable isotope (2H and 18O). The fractionation characteristics of boron isotope were used for further investigation on the contrasting fault-hosted and volcano-hosted geothermal systems. The distinct boron isotope composition of seawater was applied to confirm seawater input in some geothermal systems.

The volcano-hosted and fault-hosted geothermal systems were chemically different: - the former had higher HCO3 concentrations and Mg/Na ratios compared to the latter.

This condition was caused by CO2 magmatic gas supply in the volcano-hosted, which was insignificant or absent in the fault-hosted geothermal systems. The CO2 gas - supply produces slightly acid HCO3 thermal waters, hence together with its slower ascent due to the longer flow path would elevate the Mg2+ content. Geothermal systems hosted by faults located in the Quaternary magmatic belt were clearly supplied by magmatic fluid, thus could not be classified as a fault-hosted geothermal system. Although the Quaternary magmatic fluid input in the fault-hosted geothermal systems located in the Tertiary volcanic belt are absent, the reservoir temperature and lithium (Li) enrichment indicated a Quaternary magmatic heat source for the fault-

ix

hosted geothermal systems of Cilayu and Cisolok. Meanwhile, the other fault-hosted geothermal systems are likely heated by a deep-seated magma. Magmatic fluid input enriched the 2H and 18O isotope of some volcano-hosted geothermal systems, something that was not identified in any of the fault-hosted geothermal systems.

Boron isotope further distinguished the volcano-hosted and fault-hosted geothermal systems. The magmatic fluids input, favorable for minerals precipitation, and longer thermal water ascent in the former geothermal system promoted δ11B enrichment. In contrast, the fast thermal water ascent and absence of magmatic fluids input maintained a light δ11B signature in the fault-hosted geothermal systems. Two different types of thermal crater lakes, acid sulfate and acid chloride, had a contrast in B isotope compositions. The acid chloride crater lake had a light δ11B value representing magmatic origin, while the acid sulfate crater lakes had heavier δ11B values, produced by a sequence of processes: vapor phase separation in the subsurface, followed by evaporation and B adsorption into clay minerals on the surface. B isotope confirmed seawater input in two fault-hosted geothermal systems: Parangtritis and Krakal.

The geothermal systems on Bali were studied using the physicochemical properties of 2+ 2+ - surface thermal waters. The (Ca + Mg )/HCO3 of approximately 0.4 and the visual absence of 18O isotope enrichment ruled out a carbonate host rock type, instead the K/Mg ratios indicated water-rock interaction with a calc-alkaline magmatic rocks. The - 2+ 2+ 2+ + well correlation of HCO3 content with Ca , Mg , Sr and K revealed water-rock interaction influenced by carbonic acid. The B/Cl ratios revealed phase separation for the Bedugul and Banjar geothermal systems. The heavy δ11B of +22.5 ‰ and a Cl/B ratio of 820 confirmed seawater input in the Banyuwedang geothermal system.

x

Zusammenfassung

Es wird allgemein angenommen, dass Geothermale Systeme in vulkanischen Inselbögen ausschließlich durch vulkanische Aktivität angetrieben werden. Die komplexe geologische Struktur der Insel Java legt nahe, das dieses Erklärungsmodell dort wahrscheinlich zu einfach ist. Durch Subduktion der indo-australischen unter die eurasische Platte verursachter Vulkanismus, sowie dadurch entstandene Grabensysteme, lassen sich chronostratigraphisch dem Paläogen, Neogen und Quartär zuordnen. Die Mächtigkeit der Kruste ist durch Heraushebung und Erosion während des Subduktionsprozesses verringert. Die geographische Nähe geothermaler Systeme die sich an Vulkanen („volcano-hosted“) oder in Grabenbrüche („fault-hosted“) befinden, erlaubt es, die unterschiedlichen physiko-chemischen Eigenschaften beider Systeme zu untersuchen. In den vorliegenden Arbeiten wurden die Auswirkungen quartären Magmas auf “fault-hosted” geothermale Systeme untersucht. In zwei Kraterseen wurde die Isotopenverteilung des Elements Bor (δ11B) verglichen und die Konzentrationen von gelöstem Sulfat und Chlorid, die in vorherigen Arbeiten häufig nicht betrachtet werden, bestimmt. Zusätzlich wurde untersucht, ob karbonatisches Sediment als Untergrund in geothermalen Systemen der vulkanisch entstandenen Insel Bali möglich ist.

Die physiko-chemischen Eigenschaften geothermalen Wassers, einschließlich Anionen- und Kationenverteilungen, Konzentrationen von Spurenelementen sowie Isotopengehalte (δ2H, δ18O) wurden untersucht. Die Isotopenverteilung des Elements Bor (δ11B) wurde zur weiteren Charakterisierung der Unterschiede von „volcano- hosted“ und „fault-hosted“ geothermalen Systemen genutzt. Die bekannte Konzentration des δ11B in Meerwasser wurde zusätzlich genutzt, um das Einströmen von Meerwasser in einige geothermale Systeme zu bestätigen.

„Volcano-hosted“ und „fault-hosted“ geothermale Systeme lassen sich aufgrund ihrer - chemischen Eigenschaften unterscheiden. Erstere haben höhere HCO3 Konzentrationen und Mg2+/Na+-Verhältnisse verglichen mit Letzteren. Dieser Befund lässt sich auf den Einfluss des CO2-Gehalts in magmatischen Gasen zurückführen, der in “fault-hosted” Systemen vernachlässigbar gering bzw. nicht vorhanden ist.

Durch CO2-Eintrag wird in geothermalem „volcano-hosted“ Systemen Hydrogenkarbonat gebildet und durch das dissoziierende Proton eine Senkung des pH-Werts verursacht. Aufgrund der, durch die geologische Architektur bedingten,

xi

verlängerten Passage durch das Gestein erhöht sich der Mg2+-Gehalt. Die durch Grabenbildung in quartären Vulkangürteln entstandenen geothermalen Systeme werden eindeutig durch Wärmekonvektion magmatischer Flüssigkeit gespeist, was dazu führt, dass diese nicht als „fault-hosted“ geothermales System im eigentlichen Sinn beschrieben werden können. Obwohl ein Einströmen Quartär gebildeten Magmas in die “fault-hosted” geothermalen Systeme, Cilayu und Cisolok, die in einem im tertiär gebildeten Vulkangürtel lokalisiert sind, nicht gezeigt werden kann, kann durch die Reservoir-Temperatur sowie die Anreicherung von Lithium (Li) Quartär entstandenes Magma als Wärmequelle angenommen werden. Wahrscheinlich ist der Ursprung anderer “fault-hosted” Systeme dieser Region die Wärme tiefliegender Magmakörper. Durch magmatischen Einfluss werden 2H- und 18O-Isotope im Wasser von “volcano-hosted” Systemen angereichert. Dieses Phänomen konnte in keinem der “fault-hosted” Systeme beobachtet werden.

Eine weitere Möglichkeit der Unterscheidung geothermaler Systeme zwischen “volcano-hosted” und “fault-hosted“ besteht in der Messung des δ11B. Durch magmatische Flüssigkeit wird die Ausfällung von Mineralien unterstützt, und die oben beschriebene längere Passage führt zu einer höheren Anreicherung von δ11B. Im Gegensatz dazu führt eine schnelle Passage ohne Einströmen magmatischer Flüssigkeit nur zu einer abgeschwächten Konzentration von δ11B in „fault-hosted“ geothermalen Systemen. Durch geothermale Systeme entstandene Kraterseen können aufgrund ihrer δ11B-Konzetration der Genese charakterisiert werden. Kraterseen mit hoher Cl--Konzentration haben einen niedrigen δ11B-Wert und lassen auf magmatischen Ursprung schließen, während Kraterseen mit einer hohen Sulfat- Konzentration ebenfalls einen hohen δ11B-Wert aufweisen aber im wesentlichen mit Wasser meteorischen Ursprungs gefüllt sind. Dem hohen δ11B-Wert liegt der Ablauf folgender Prozesse zugrunde: unterhalb der Oberfläche kommt es zu einer Trennung der Dampfphase, durch anschließende Verdunstung adsorbiert Bor an Tonmineralien der Oberfläche. Die hohen δ11B Werte bestätigen den Meerwassereintrag in die „fault- hosted“ geothermalen Systeme von Parangtritis und Krakal.

Geothermale Systeme der Insel Bali wurden anhand der physiko-chemischen Eigenschaften des geothermalen Oberflächenwassers untersucht. Ein ([Ca2+]+ 2+ - 18 [Mg ])/HCO3 -Verhältnis von ~0,4 und eine nicht vorhandene Anreicherung von δ O schließen den Einfluss von Karbonat-Sediment aus. Die in den Untersuchungen bestimmten K+/Mg2+-Verhältnisse deuten hingegen auf eine Einbeziehung von xii

basischer Magma in den Reaktionskreislauf. Die beobachteten hohen Korrelationen - 2+ 2+ 2+ + der HCO3 Gehalte mit Ca , Mg , Sr und K lassen auf eine Kohlensäure vermittelte Reaktion an der Phasengrenze Wasser-Stein schließen. Eine Phasentrennung in den geothermalen Systemen von Bedugul und Banjar konnten durch die B/Cl-Ratios gezeigt werden. Ein δ11B von +22.5‰ und ein Cl/B-Verhältnis von 820 bestätigten den Meerwassereintrag in das geothermale System von Banyuwedang.

xiii

xiv

INTRODUCTION

I. Introduction

Indonesia is known for having the biggest geothermal resources in the world, mostly spread along the Sunda volcanic island arc, which extends from west to east, from the island of Sumatera to the island of Damar. The island of Java located in the middle of the island arc is the most geothermal exploited area due to the huge demand of electricity. Based solely on the geological association, hosted by either a volcano or fault, geothermal systems on Java can be divided into volcano-hosted and fault-hosted, respectively. Concerning their potential as an electricity power source, to date the geothermal development program has been only focused on the volcano- hosted geothermal systems due to the assumption of low potential energy in the fault- hosted geothermal systems. Apart from the energy issue, the presence of fault-hosted geothermal systems on Java provides an essential opportunity to determine the influence of the Quaternary volcanoes to the systems. In a volcanic islands arc, the geothermal systems are considered to be simply dominated by the Quaternary volcanic activities. Therefore, less attention is given to the fault-hosted geothermal system located in this geological setting. Several fault-hosted geothermal systems distributed in a Quaternary volcanic arc, for instances in the Liquine-Ofqui fault zone, Chile and in the Southern Apennines, Italy (Alam et al., 2010; Italiano et al., 2010), indicated contrasting geochemical processes compared to volcano-hosted geothermal systems. In the fault-hosted system deep circulated groundwater is simply conductive heated, while in the volcano-hosted system involves condensation of volcanic steam (Alam et al., 2010). On Java the geothermal systems are probably more complex due to the geological setting. The island is mainly developed by three main volcanisms: Paleogene volcanism formed the Tertiary volcanic belt, distributed along the southern part, followed by Neogene volcanism shifted northward to the middle of the island and finally Quaternary volcanism emerged along the Neogene volcanic (Soeria-Atmadja et al., 1994). The relatively impermeable Tertiary volcanic rocks might inhibit the extent of Quaternary magmatic fluids flow. The subduction in the south of Java developed several major and minor faults that are distributed in the two volcanic belts, Tertiary and Quaternary. The northward plates subduction has uplifted the southern part of the island, followed by erosion, hence produced a thinner crust (Clements et al., 2009; Hall et al., 2007). The thin crust enables deep penetrated groundwater through fault to extract the heat. Considering these geological settings, different characteristics

1

INTRODUCTION

between fault-hosted located in the Quaternary volcanic belt and the Tertiary volcanic belt could be expected. Two characteristics of boron (B) isotope fractionation might be beneficial for further investigation on the contrast between volcano-hosted and fault-hosted geothermal systems. Those characteristics are 1) the relatively conservative at temperatures >65 °C (Aggarwal and Palmer, 1995; Palmer et al., 1987) and 2) the significant fractionation due to B adsorption/coprecipitation onto clay, iron oxide and calcite and evaporate minerals (Agyei and McMullen, 1968; Hemming and Hanson, 1992; Lemarchand et al., 2007; Oi et al., 1989; Palmer et al., 1987; Schwarcz et al., 1969; Spivack and Edmond, 1987; Swihart et al., 1986; Vengosh et al., 1991a; Vengosh et al., 1991b; Vengosh et al., 1992). The fast ascent of fault-hosted thermal water through a fault and the absence of magmatic fluids input should result in a light δ11B signature. Conversely the slow ascent and the presence of magmatic fluids input might be favorable for δ11B enrichment. Additionally the presence of two contrasting thermal crater lakes, acid-sulfate and acid-chloride, on Java provide an opportunity to determine their distinct boron isotope signatures, which likely has been overlooked so far. Finally, the study on the characteristic of volcano-hosted and fault-hosted geothermal systems was extended to the volcanic island of Bali. Apart from this objective, the island is an ideal location to investigate the role of a carbonate rock basement to the geothermal systems. In such a geological setting, the carbonate rock was identified hosting the geothermal systems instead of volcanic rocks, e.g., in Vicano-Cimino and Sabatini-Tolfa (Cinti et al., 2011; Cinti et al., 2014).

I.1. Thesis outline This thesis consists of six chapters with three main bodies. Chapter 1, Introduction, provides the context of the research and overview of some theoretical backgrounds. Chapter 2 explains the geological setting of the study areas, sampling and analytical methods. Chapter 6 presents the overall conclusion answering the research objectives and outlook related to the results of this study. Below the three main chapters and my contributions on every chapter are briefly described.

Chapter 3 This chapter provides a classification of the geothermal systems on Java. The systems were classified into volcano-hosted and fault-hosted based solely on their

2

INTRODUCTION

geological association, either in a volcano or fault zone. Subsequently the physicochemistry of thermal waters clearly divided the two geothermal systems. The possible role of Quaternary magma as a heat source for fault-hosted geothermal system was investigated based on the reservoir temperatures and trace element enrichments. This research is published in Journal of Volcanology and Geothermal Research (Purnomo, B.J. & Pichler, T., 2014). My contributions on this article are: - sampling and measurement of some physicochemical parameters in the field, - preparing samples for laboratory analyses, - quality control on the laboratory results, - data interpretation, and - writing the first draft of the published paper.

Chapter 4 In this chapter boron isotope was used to further investigate the distinct geochemical characteristics between volcano-hosted and fault-hosted geothermal systems. The contrast in B isotope signatures between two distinct thermal crater lakes, acid-sulfate and acid-chloride, was determined. Boron isotope was also applied to confirm seawater input in some geothermal systems on Java. A part of the result was presented as a poster (Purnomo, B.J., Pichler, T. and You, C.-F, 2014) in American Geophysical Union (AGU) Fall Meeting 2014. The final version of this chapter has been submitted to Geochimica et Cosmochimica Acta (Purnomo, B.J., Pichler, T. and You, C.-F, submitted in November 2014, current status: under review). My contributions on this article are: - sampling and measurement of some physicochemical parameters in the field, - preparing samples for laboratory analyses, - data interpretation, and - writing the first draft of the submitted manuscript.

Chapter 5 This chapter presents an overview of geothermal systems on Bali that were classified into a volcano-hosted and fault-hosted. However, the main focus is to investigate the host-rock, either carbonate or volcanic, considering the Quaternary

3

INTRODUCTION

volcanoes have a carbonate sedimentary basement. The processes that govern the chemical composition of the surface thermal waters, seawater input and geothermometry of the reservoir were also determined. The result of this study has been submitted to Journal of Volcanology and Geothermal Research (Purnomo, B.J. and Pichler, T., submitted in February 2015, current status: under review). My contributions on this article are: - sampling and measurement of some physicochemical parameters in the field, - preparing samples for laboratory analyses, - quality control on the laboratory results, - data interpretation, and - writing the first draft of the submitted manuscript.

I.2. Research objectives The main purpose of this study is to characterize the system of geothermal fields located in the Sunda volcanic island arc. The geothermal system probably is complicated by the complex geological setting formed by the dynamic plates boundary. The presence of multiple volcanisms, fault zones and thin crust potentially influence the geothermal system, instead of exclusively dominated by Quaternary volcanic activities. Several specific objectives of the study are:

x to define the geothermal systems into a volcano-hosted or fault-hosted systems, x to calculate the reservoir temperatures of the geothermal systems, x to characterize the contrast between volcano-hosted and fault-hosted geothermal systems, including the physicochemical processes and fluids origin, x to investigate the role of Quaternary volcanoes to the fault-hosted geothermal systems, x to determine the possible heat sources of the fault-hosted geothermal systems, x to apply boron isotope for further investigation on the contrast between fault- hosted and volcano-hosted geothermal systems, x to investigate the boron isotope signatures of two distinct crater lakes, acid- sulfate and acid-chloride, x to confirm seawater input in some geothermal systems, considering the location close to sea, and

4

INTRODUCTION

x to determine the role of a carbonate sedimentary basement on the volcanic island of Bali.

I.3. Theoretical backgrounds I.3.1. An overview of geothermal system Geothermal system is a heat transfer process from the Earth’s crust to the surface. Generally the heat transfer involve meteoric water, hence it is called as hydrothermal system (Hochstein and Browne, 2000). A geothermal system consists of three main components: 1) heat source, 2) reservoir and 3) water (e.g., Goff and Janik, 2000; Kuhn, 2004). The heat sources for geothermal systems include magmas within the crust, intracrustal nonmagmatic and conductive heat flow (Hochstein and Browne, 2000). The first can present as a convective magma fluid input or conductive heat transfer from cooling magma, the second is present due to the anomaly thermal gradient in areas with relatively thin crust and the third is formed in the sedimentary basin where thick sediments generate heat and overpressure (Goff and Janik, 2000; Kuhn, 2004; Nicholson, 1993). Faulds et al. (2010) introduced a term of ‘amagmatic’ heat source for a deep-seated magma, to distinguish it with Quaternary magma. Apart from those heat sources, back to the experiment by Byerlee (1978), a frictional heat source was proposed. However, its presence in natural geothermal systems has never been proved, hence it is referred to ‘the stress-heat flow paradox’ (Lachenbruch and Sass, 1980). Several classifications of geothermal systems have been proposed. Simply based on the reservoir temperature geothermal system was classified into three (Hochstein and Browne, 2000):

1) High temperature (T >225 °C), 2) Intermediate temperature (125 to 225 °C), and 3) Low temperature (T <125 °C)

However the range of temperatures in the classification is not rigid, for an instance Benderitter and Cormy (1990) suggested a low temperature geothermal system accounted for temperatures below 100 °C.

Considering the geology, hydrology and engineering, Goff and Janik (2000) classified geothermal system into:

1) Young igneous systems

5

INTRODUCTION

Geothermal system hosted by Quaternary volcanic or intrusion. This system generally spreads along plate boundaries and spreading centers.

2) Tectonic systems Geothermal system associated with anomaly heat flows due to the relatively thin crust. The heat is extracted by a deep penetrated groundwater hosted by faults.

3) Geopressured systems This geothermal system distributed in sedimentary basins. Subsidences and thick sediments form the heat and overpressure.

4) Hot dry rock systems The heat stored in an impermeable rock is extracted by injecting water to the rock, subsequently the resulting thermal water is pumped out to the surface.

5) Magma tap systems The geothermal system taps the heat of a shallow magma by circulating water.

Nicholson (1993) proposed a classification of geothermal systems based on: heat transfer, temperature, fluid composition and topographic (Fig. 1.1). Heat transfer divides geothermal system into two, dynamic and static systems. The former involves circulation of fluids from depth, where meteoric water circulated into the depth, extract heat, then ascend to the surface. Meanwhile, in the latter system the thermal fluid is trapped in the deep strata. The dynamic geothermal systems are distributed in plate boundaries and spreading centers, while the static system are located in tectonic stabile areas, for instances in Russia, Eastern Europe and Australia. Considering the geological setting of the study areas is a subduction zone, only the dynamic system is further elaborated. The dynamic geothermal system is divided into a high temperature system, generally associated with Quaternary magmas, and a low temperature system, associated with conductive heat transfer of anomaly heat flow. The high temperature system is classified into a liquid-dominated and a vapor-dominated based on the dominant phase in the reservoir. The contrast topographic of the liquid- dominated geothermal systems divides the system into a low-relief and a high relief system. The former is hosted by a silicic volcano, while the latter by an andesitic volcano, which is a typical of volcanic island arc.

6

INTRODUCTION

Geothermal system

dynamic static (convective) (conductive)

low geopressured high low temperature temperature temperature

liquid- vapor- dominated dominated

low-relief high-relief

Fig. 1.1. Geothermal system classification by Nicholson (1993).

Exclusively based on the geological setting, Saemundsson (2009) classified geothermal systems into:

1) Volcanic-geothermal system Geothermal system associated with volcanoes, heated by intrusion or magma.

2) Convective system Geothermal system generated due to the deep circulation of meteoric water through fractures, extracted heat from an anomaly high temperature gradient.

3) Sedimentary system Geothermal system hosted by thick sedimentary layers, heated by regional thermal gradient.

Working in a complex of major fault and Quaternary volcanic arc geological setting in the Liquiñe-Ofqui Fault Zone (LOFZ), Alam et al. (2010) divided the

7

INTRODUCTION

geothermal systems into a volcanic type, associated with stratovolcanoes, and a structural (non-volcanic) type, hosted by the faults. This classification is adopted in this study, however, the terms of volcano-hosted and fault-hosted geothermal systems are preferred. Volcano rather than volcanic is used to avoid confusion with volcanic as a rock type, while fault is applied to clarify its geological structures. Although a volcanic arc is generally dominated by volcano-hosted geothermal systems, fault-hosted geothermal systems might be present, for instances in the Liquine-Ofqui fault zone, Chile and in the Southern Apennines, Italy (Alam et al., 2010; Italiano et al., 2010). In the fault-hosted geothermal system a deep circulated meteoric water through faults is conductive heated, while the volcano-hosted geothermal system involves condensation of volcanic steam (Alam et al., 2010). The heat sources of fault-hosted geothermal systems could be a deep seated magma. This was named as ‘amagmatic’ heat source to distinguish it with shallow magmas of Quaternary volcanoes, as has been applied in the Great Basin, USA and Western Turkey (Faulds et al., 2010).

I.3.2. Geothermal manifestations The existence of a geothermal system in the subsurface can be recognized by the surface features (manifestation), which include solfatara, fumarole, acid crater lake, hot spring, mud pool, steaming ground and altered ground (Fournier, 1989; Giggenbach, 1988; Giggenbach and Stewart, 1982; Goff and Janik, 2000; Hedenquist, 1990; Henley and Ellis, 1983; Hochstein and Browne, 2000; Nicholson, 1993). These manifestations are formed by a sequence of processes that briefly can be explained as follow (Giggenbach, 1988; Nicholson, 1993). Volatile magmas contain gases, H2O, CO2, H2S, HCl and HF, rise and react with deep circulated groundwater producing acid water. The acid water-rock interaction subsequently consumes H+ to form a neutral-chloride water, which then ascend to the surface. During the ascent separation of fluid phases might occur due to a rapid drop in temperature and pressure. The remaining chloride water discharges as a neutral chloride hot spring on the surface. In a low relief topography the water emerges above the upflow zone, close to the acid sulfate and bicarbonate hot springs, while in the high relief topography (stratovolcano) it discharges far away from the upflow zone to the flank of stratovolcano (Fig. 1.2). Meanwhile, the vapor phase resulted by the phase separation usually is rich in CO2 and H2S gases, hence the condensation and - reaction with O2-rich groundwater produce a slightly acid bicarbonate (HCO3 ) water

8

INTRODUCTION

- and an acid sulfate (SO4 ) water, respectively. In a vapor-dominated geothermal system, thermal water ascends exclusively as a steam phase that rich in CO2 and H2S gases, therefore, the resulting manifestations are mud pools, acid sulfate and bicarbonate hot springs (Fig. 1.3). A geothermal system generally produces three types of hot springs: 1) neutral chloride, 2) acid sulfate and 3) bicarbonate waters, however, mixtures between the individual groups are common (Hedenquist, 1990; Hochstein and Browne, 2000; Nicholson, 1993; White, 1957). The composition of these thermal springs is controlled by two main factors: 1) reservoir condition and 2) secondary processes during ascent. Host-rock type, temperature, residence time and magmatic fluid input govern the geothermal brines. The physicochemistry of thermal waters might change during the ascent due to phase separation, minerals precipitation, adsorption/desorption, water- rock interaction, groundwater dilution and seawater input (in a coastal area). The chemical composition of a thermal spring records the physicochemical processes in the subsurface. Chemically inert constituent (tracers), Cl, B, Li, Rb and Cs, can be used to track the source, while chemically reactive species (geoindicators), e.g., Na,

K, Mg, Ca and SiO2, record the physicochemical processes during thermal water ascent (Ellis and Mahon, 1977; Giggenbach, 1991; Nicholson, 1993).

I.3.3. Geothermometry Solute geothermometers are commonly applied to estimate the reservoir temperature of a geothermal system, by using either the absolute concentration, solute ratio or solute relation. The geothermometers were formulated based on the temperature dependency of a mineral-fluid reaction in the reservoir. This is preserved on the surface due to the slow re-equilibrium of the minerals in a low temperature system (Fournier, 1977). The successful application of solute geothermometers for hot springs relies on five basic assumptions: 1) exclusively temperature dependent mineral-fluis reaction; 2) abundance the mineral and/or solute; 3) equilibrium reaction in the reservoir; 4) no re-equilibrium; and 5) no mixing or dilution (Ellis, 1979; Fournier, 1977; Fournier and Truesdell, 1973; Nicholson, 1993; Truesdell, 1975). The last assumption can be overcome if the extent of dilution can be calculated. Below, solute geothermometers that were used in this study are listed.

9

Fig. 1.2. Conceptual model of liquid-dominated geothermal system in high relief topographic (stratovolcano) (adopted from Henley and Ellis, 1983).

INTRODUCTION

Fig. 1.3. Conceptual model of vapor-dominated geothermal system (adopted from Nicholson, 2000).

a) Silica geothermometer Silica geothermometer is deducted from the solubility of quartz in certain temperatures and pressures. The thermometer is based on the absolute silica concentration and thus sensitive to secondary processes such as mixing, boiling and precipitation (Fournier and Rowe, 1966). Several silica geothermometers were established for different conditions, such as:

- No steam loss T (°C) = {1309 / (5.19 – log SiO2)} – 273 (Fournier, 1977)

- Max. steam loss T (°C) = {1522 / (5.75 – log SiO2)} – 273 (Fournier, 1977)

In low temperatures the silica content is attributed to the solubility of chalcedony, cristobalite and amorphous silica (Arnorsson, 1970; Arnorsson, 1975; Fournier and

11

INTRODUCTION

Rowe, 1962; Fournier and Rowe, 1966). Therefore, the silica geothermometer, for an example chalcedony, is stated as:

- Chalcedony T (°C) = {1032 / (4.69 – log SiO2)} – 273 (Fournier, 1977)

The application of silica geothermometers for hot springs usually calculates underestimation temperatures, close to the discharge temperature, due to dilution by shallow water. This can be overcome by calculating the silica ‘parent’ using the silica mixing model of Truesdell and Fournier (1977) (Fig. 1.4).

Fig. 1.4. Silica vs. enthalpy diagram for silica parent calculation based on Truesdell and Fournier (1977).

b) Na/K geothermometer The Na/K geothermometer was calculated using the Na2+ and K+ ratios. The thermometer was formulated based on the temperature-dependent reaction of ion exchange between albite and K feldspar (Fournier, 1979; Giggenbach, 1988).

+ 3 + NaAlSi3O8 + K (aq) = KAlSi O8 + Na (aq) (albite) (K feldspar)

12

INTRODUCTION

The Na/K gethermometer formula for instances are:

T (°C) = {1217 / (log (Na/K) + 1.483} – 273 (Fournier, 1979)

T (°C) = {1390/ (log (Na/K) + 1.750} – 273 (Giggenbach, 1988)

This geothermometer is suitable for reservoir temperatures ranging from 180 to 350 °C, but not for temperatures below 120 °C (Ellis, 1979; Nicholson, 1993). However, the geothermometer has been indicated resulting in overestimation temperatures when applied to thermal waters with high Ca2+ concentrations. This is considered due to the competition of Ca2+, Na+ and K+ during ion exchange reaction (Nicholson, 1993). c) Na/K/Ca geothermometer The empiric Na/K/Ca geothermometer was developed by Fournier and Truesdell (1973) for Ca-rich thermal waters, i.e., (√Ca / Na) >1. The thermometer was defined as:

T (°C) = 1647 / {log (Na/K) + β[log(√Ca / Na) + 2.06] + 2.47} – 273

The thermometer is considered working well for reservoir temperatures above 180 °C. However, the application to hot springs could be troublesome. During thermal water 2+ 2+ ascent the Ca content might be depleted by CO2 release, Mg exchange and calcite precipitation. d) Na/K/Mg geothermometer Giggenbach (1988) developed a ternary diagram of K/100-√Mg-Na/1000 as a geothermometer and a tool to assess suitable hot springs for geothermometer application (Fig. 1.5). In the diagram thermal waters are divided into three groups, fully equilibrium, partially equilibrium and immature waters. Immature water potentially produce inaccurate reservoir temperatures calculation. e) Na/Li geothermometer The Na/Li ratios of geothermal water were identified having an inverse correlation with temperatures (Ellis and Wilson, 1960; Koga, 1970). The thermometer was formulated based on the theoretical exchange reaction of Fouillac and Michard (1981):

Clay-Li + H+ = Clay-H + Li+

13

INTRODUCTION

Fig. 1.5. Evaluation of Na-K-Mg geothermometer (Giggenbach, 1991).

with an equation:

T (°C) = 1000 / {log (Na/Li) + 0.389] – 273 (Fouillac and Michard, 1981)

Later, a new formula was established by Kharaka et al. (1982) as follow:

T (°C) = 1590 / {log (Na/Li) + 0.779] – 273

This thermometer was considered resulting the most reliable reservoir temperatures for geothermal systems with a carbonate reservoir type (Minissale and Duchi, 1988).

I.3.4. Boron isotope Boron is a trace element, thus can be used to track the thermal water origin.

Dissolve boron is mainly present as B(OH)3 (boric acid, trigonal species) and B(OH)4 (borate anion, tetrahedral species) (Dickson, 1990; Hershey et al., 1986). At low pH

(<7) only B(OH)3 is present, and conversely at pH >10 boron is found as a B(OH)4

14

INTRODUCTION

species. Boron has two stable isotopes, 10B and 11B, with an abundance of 19.8 ‰ and 80.2 ‰, respectively (e.g., Xiao et al., 2013; Bart, 1993). Boron isotope composition is reported as δ11B per mil (‰) relative to the standard of NIST-SRM 951 (Catanzaro et al., 1970), with a formula:

11 11 10 11 10 δ B (‰) = [( B/ B)sample / ( B/ B)NIST-SRM-951 – 1] x 1000

Geothermal waters have a wide range of δ11B composition, from -9.3 to +44 ‰ (Aggarwal et al., 2000; Aggarwal et al., 1992; Barth, 1993; Leeman et al., 1990; Musashi et al., 1988; Palmer and Sturchio, 1990; Vengosh et al., 1994b). The δ11B composition of thermal water is mainly controlled by: 1) host-rock type, 2) fluid mixing, 3) B isotope fractionation and 4) steam phase separation. The last factor generally only enriches the δ11B of up to 4 ‰ and thus is considered insignificant (Kanzaki et al., 1979; Leeman et al., 1992; Nomura et al., 1982; Spivack et al., 1990; Yuan et al., 2014). During thermal water-rock interaction, 11B is released into water, hence reduces the 11B composition of the rock (Musashi et al., 1991; Palmer and Sturchio, 1990). Carbonate rocks have a wider range of δ11B composition compared to volcanic rocks from island arc, respectively ranged from +1.5 to +26.2 ‰ (Hemming and Hanson, 1992; Vengosh et al., 1991a) and from -2.3 to +7 ‰ (Ishikawa and Nakamura, 1992; Palmer, 1991). B isotope fractionation at water temperatures above 65 °C was reported insignificant (Aggarwal and Palmer, 1995; Palmer et al., 1987), thus relatively conservative during thermal water ascent. Groundwater generally has heavier δ11B compositions than thermal water, thus groundwater dilution enriche the δ11B of thermal water (Palmer and Sturchio, 1990; Vengosh et al., 1994a). Seawater input has been reported elevating the δ11B value of thermal water, for instances at the Reykjanes and Svartsengi, Iceland as well as at the Izu-Bonin arc, Kusatsu-Shirane and Kagoshima, Japan (Aggarwal and Palmer, 1995; Aggarwal et al., 2000; Kakihana et al., 1987; Millot et al., 2009; Musashi et al., 1988; Nomura et al., 1982; Oi et al., 1993). Adsorption/coprecipitation of B into minerals leads to fractionation of the light 10B into minerals, thus increases the δ11B of water (Palmer et al., 1987; Schwarcz et al., 1969; Xiao et al., 2013). B can be adsorbed/incorporated by clay minerals and iron oxide (Lemarchand et al., 2007; Palmer et al., 1987; Schwarcz et al., 1969; Spivack and Edmond, 1987; Vengosh et al., 1991b), calcite (Hemming and Hanson, 1992; Vengosh et al., 1991a) and evaporite minerals (Agyei and McMullen, 1968; McMullen et al., 1961; Oi et al., 1989; Swihart et al., 1986; Vengosh et al., 1992).

15

INTRODUCTION

16

GEOLOGICAL SETTING, SAMPLING AND ANALYSES

II. Geological setting, sampling and analyses

II.1. Geological setting The geological features of the western part of the Indonesian archipelago was started by the collision of the Sibumasu and Indochina–East Malaya, indicated by the distribution of Triassic granite across the island of Sumatera (Hall, 2009). This event pushed the continental active margin, which then ceased in the early Cretaceous due to collision of the Australian microcontinental with the Java-Meratus subduction (Hall et al., 2007; Smyth et al., 2008). The subduction can be detected by the presence of metamorphic rocks, stretch along Sumatera to Central Java and turn to the north to the Borneo island (Hall, 2009) (Fig. 2.1). In the middle the subduction reactivated to the south due to the rapid northward movement, c.a., 6-7 cm/a, of the Australian plate (Hall, 2009; Hamilton, 1979; Müller et al., 2000; Schellart et al., 2006; Simandjuntak and Barber, 1996). This subduction generated a volcanic arc in the active margin, marked by the distribution of the Tertiary volcanic belt along the southern part of Java island (Soeria-Atmadja et al., 1994; Van Bemellen, 1949) (Fig. 2.2). In the early , the subduction ceased and resumed again in the middle Miocene, indicated by the low volcanic activities, due to the further northward movement of the subduction hinge (Macpherson and Hall, 1999; Macpherson and Hall, 2002; Smyth et al., 2008). The volcanic activities increased again in the late Miocene by the formation of the Neogene volcanic arc to the north of the Tertiary volcanic belt (Hall, 2002; Hall et al., 2007; Macpherson and Hall, 2002; Soeria- Atmadja et al., 1994). Later the Quaternary volcanic replaced the Neogene volcanic arc, hence the current features on Java only have two volcanic belts, the Tertiary in the south and the Quaternary in the middle (Hamilton, 1979; Soeria-Atmadja et al., 1994). The volcanisms on Java produced andesitic rocks, where the younger volcanic (Quaternary) rocks are more alkaline (Soeria-Atmadja et al., 1994). The Quaternary volcanism produced magmas ranging from tholeiites to high-K calc alkaline, named as ‘the normal island arc association’ (Whitford et al., 1979). The resulting magma type is associated with the distance to the subduction trench/the depth of Benioff zone (Wheller et al., 1987; Whitford et al., 1979). Volcanoes with deeper Benioff zones

17

GEOLOGICAL SETTING, SAMPLING AND ANALYSES

produce richer K volcanic rocks. On Java, the Muria volcano has the highest K content due to its location in the back arc (Wheller et al., 1987; Whitford et al., 1979). Apart from the location relative to the subduction trench, the melting materials and fluid flux also affect the typical magma produced in a volcanic island arc. Calc-alkaline magma is produced by a high fluid flux as the result of melting of the mantle wedge and sediment, while K-rich magma is associated with a lower fluid flux and melting of a deeper mantle (Cottam et al., 2010). The basement of the Sunda volcanic arc are vary from a continental crust in the West, accretionary complexes in the central to east Java and an oceanic crust on Bali to (Curray et al., 1977; Hamilton, 1979). Avraham and Emery (1973) predicted the crustal thickness on Java ranging from 20 to 25 km and thinner to the east, reached 15 km on Flores.

Fig. 2.1. Geographic and tectonic map of the Indonesian archipelago (after Hamilton, 1979 and Simandjuntak and Barber, 1996; the plates boundaries were based on Hall, 2009).

18

GEOLOGICAL SETTING, SAMPLING AND ANALYSESES

Fig. 2.2. Geological setting of the Java island with the old Tertiary volcanic belt, Quaternary volcanoes complex and faults.ts. (adopted from Hamilton,1979; Simandjuntak and Barber,1996; Hoffmann-Rothe et al., 2001; and Soeria-Atmadja et al., 1994).

GEOLOGICAL SETTING, SAMPLING AND ANALYSES

The subduction in between the early Miocene to thrust the Tertiary volcanic belt to the north by more than 50 km (Hall et al., 2007). The continuous subduction has also uplifted the southern part of Java. Followed by erosion, the process thinned the crust and exposed the Tertiary (Paleogene) volcanic belt. In Central Java the volcanic belt has been removed by excessive erosion, hence outcropped the Cretaceous basement (Clements et al., 2009; Hall et al., 2007). This block is the most thrust compared to the West Java and East Java due to the presence of a couple two major strike-slip faults, the Central Java fault and Citandui fault. The subduction pushed the East Java and West Java blocks to the north, hence uplifted the southern part of Central Java (Bahar and Girod, 1983; Satyana, 2007; Situmorang et al., 1976). Apart from these faults, two other major faults are present on Java, namely the E-W backarc-thrust of Barabis-Kendeng and the NE-SW strike-slip fault of Cimandiri (Hoffmann-Rothe et al., 2001). These faults have been generated since the Neogene time by compressional forces (Hall, 2002; Simandjuntak and Barber, 1996). The Cimandiri fault is an active fault with a slip rate of about 6 to 10 mm/a (Sarsito et al., 2011; Setijadji, 2010). Besides those major faults, there are several smaller faults, which include the E-W Lembang fault in West Java, the NE-SW Opak fault in Central Java and the NE-SW Grindulu fault in East Java.

II.2. Sampling and analyses The sampling of thermal and cold waters was performed in two periods, July to September 2012 and October to November 2013. The first sampling covered almost the whole area of Java, while the second was focused on Bali with some additional samples from Java. The water samples were taken from hot spring, cold spring, shallow thermal wells, deep geothermal well, steam vent, crater lake, freshwater lake and seawater (e.g., Fig. 2.3). Temperature, pH, conductivity, ORP and alkalinity, were measured in the field, either by probe or acid titration (HACH, 2007). The samples were filtered through a 0.45 μm nylon membrane. A part of the filtered sample was used for alkalinity measurement and two splits for the determination of anion, cation and isotopic compositions were stored in pre-rinsed polyethylene bottles and transported to the University of Bremen for further analyses. The cation split was acidified to 1% - 2- - concentrated HNO3 to avoid precipitation of metals. The anions, Cl , SO4 , NO3 and Br-, were analyzed by ion chromatography using an IC Plus Chromathograph (Metrohm). The cations, Ca2+, Mg2+, Na+ and K+, and Si were determined by

20

GEOLOGICAL SETTING, SAMPLING AND ANALYSES

inductively coupled plasma-optical emission spectrometry (ICP-OES) using an Optima 7300 instrument (Perkin Elmer). Trace elements of B and As were measured by using inductively coupled plasmamass spectrometry (ICP-MS) using an iCAP-Q instrument (Thermo Fisher). Stable isotopes (18O and 2H) were determined on a LGR DLT-100 laser spectrometer (Los Gatos Research). The isotopes results were reported in δ per mil (‰) relative to VSMOW with an analytical uncertainty of approximately ± 1 ‰ for δ2H and ± 0.2 ‰ for δ18O. The B isotope composition was analyzed using a multi-collector inductively coupled plasma mass spectrometer (MC-ICP-MS, Neptune, Thermo Fisher Scientific) at the Isotope Geochemistry Laboratory, National Cheng Kung University, Taiwan by following the procedure of Wang et al. (2010). A volume of 0.5 or 1 mL sample containing a minimum of 50 ng B was used in the measurement to ensure duplicated analysis. Prior to measurement, the HNO3 in the samples was substituted with H2O to minimize the memory effects. B was purified from the samples by micro-sublimation technique at 98±0.1 °C in a thermostatic hot plate rack. The 11B data were reported in δ per mil (‰) relative to the standard of SRM NBS 951 with an analytical uncertainty of < 0.2 ‰.

21

GEOLOGICAL SETTING, SAMPLING AND ANALYSES

(a) (b)

(c) (d)

(e) (f)

Fig. 2.3. (a) Kawah Sikidang acid crater lake, Dieng; (b) Cisolok geyser; (c) Kawah Kreta steam vent, Kamojang; (d) Dieng geothermal brines; (e) Yeh Panas hot spring, Bali; (f) Tirta Husada thermal shallow wells, Bali.

22

GEOTHERMAL SYSTEMS ON THE ISLAND OF JAVA, INDONESIA

III. Geothermal systems on the island of Java, Indonesia

Modified from:

Budi Joko Purnomo * & Thomas Pichler *

* Geochemistry & Hydrogeology, Department of Geosciences, University of Bremen, Germany

Journal of Volcanology and Geothermal Research (2014) DOI: 10.1016/j.jvolgeores.2014.08.004

23

GEOTHERMAL SYSTEMS ON THE ISLAND OF JAVA, INDONESIA

Abstract

This paper presents an overview of all known geothermal systems on the island of Java by presenting physicochemical data for associated hot springs, cold springs and acid crater lakes. A total of 69 locations were sampled and classified based on their position in either a volcanic complex (volcano-hosted) or a fault zone (fault- hosted). In particular the potential of a magmatic heat source for fault-hosted geothermal systems was investigated. Volcano-hosted geothermal systems had - higher HCO3 concentrations and higher Mg/Na ratios than fault-hosted geothermal systems. This geochemical difference is likely due to degassing and subsequent CO2- water reaction in the volcano-hosted systems, which is absent in the fault-hosted geothermal systems. The HCO3 vs. Cl and Mg/Na vs. SO4/Cl systematics indicated that fault-hosted geothermal systems located in the active Quaternary volcanic belt received shallow magmatic fluids, hence should be classified as volcano-hosted geothermal systems. The heat source of fault-hosted geothermal systems located in the old (Tertiary) volcanic belt were investigated by a combination of Li enrichment and calculated reservoir temperatures. There a shallow magmatic heat source was only indicated for the Cilayu and Cisolok geothermal systems. Thus, a deep seated magma was considered to be the heat source for the fault-hosted geothermal systems of Cikundul, Pakenjeng, Parangtritis and Pacitan. In ten of the volcano-hosted geothermal systems, 2H and 18O isotopes enrichments were found, but not in any of the fault-hosted geothermal systems. Stable isotope enrichment due to evaporation was recognized in the Kawah Candradimuka and Kawah Sileri, Kawah Hujan and Candi Gedong Songo geothermal systems. A combination of intensive evaporation and magmatic gases input produced very heavy stable isotopes in the hot acid crater lakes of the Kawah Kamojang, Kawah Sikidang and Kawah Putih geothermal systems. The addition of substantial amounts of andesitic water to the geothermal fluid was observed in the Candi Songgoriti, Banyuasin and Pablengan geothermal systems. Contrary to established belief fault-hosted geothermal systems on Java could be considered a potential source for geothermal energy.

Keywords: Java, volcano-hosted and fault-hosted geothermal systems, shallow and deep magmatic heat sources, geochemistry, stable isotope

24

GEOTHERMAL SYSTEMS ON THE ISLAND OF JAVA, INDONESIA

III.1. Introduction At least 62 geothermal fields with the potential for exploitation are present on the island of Java (Setijadji, 2010). Following Alam et al (2010) geothermal fields can be divided into volcano-hosted and fault-hosted geothermal systems based on their geologic association. The former is a geothermal system related to a volcanic complex and the latter is a geothermal system located in a fault zone. To date, seven volcano-hosted geothermal fields were developed and five of them produced electricity. Fault-hosted geothermal fields were not developed and are rarely explored, due to the assumption of insufficient energy. However, considering the geology of Java, a volcanic (magmatic) influence on the fault-hosted geothermal systems is likely. In other volcanic arcs around the World, fault-hosted geothermal fields which are located close to volcanic areas indicate a heating of deep circulated meteoric water, e.g., in the Liquiñe-Ofqui fault zone of Chile and in the Southern Apennines of Italy (Alam et al., 2010; Italiano et al., 2010). Using a trend of B enrichment, Alam et al (2010) suggested for the Liquiñe-Ofqui fault zone (a) heating of meteoric water in fault-zone hosted geothermal systems and (b) condensation of volcanic steam in volcano-hosted geothermal systems. However, the authors did not indicate the heat source of the fault hosted geothermal system. Arehart et al. (2003) identified a magmatic heat source for the Steamboat geothermal system (Nevada, USA), based on trace metal and gas data. Historically this geothermal system was considered as an extensional geothermal type with anomalous heat flow as the heat source (Wisian et al., 1999). Anomalous heat flow in the Alpine fault, New Zealand, for example, is considered to be caused by uplift and erosion (Allis and Shi, 1995; Shi et al., 1996). Here physicochemical processes, fluid sources and reservoir temperature of volcano and fault-hosted geothermal systems on Java were examined, using chemical and isotope (2H and 18O) data. The data indicated a magmatic influence on the fault- hosted geothermal systems, and thus a hidden energy potential for some of the fault- hosted geothermal systems on Java.

III.2. Sampling locations Water samples were collected from July to September 2012, the end of the dry season on Java. The samples were taken from 25 geothermal systems: (1) Cisolok, (2) Cikundul, (3) Batu Kapur, (4) Ciater, (5) Maribaya, (6) Tampomas, (7) Patuha, (8) Pangalengan, (9) Darajat, (10) Kamojang, (11) Cipanas, (12) Kampung Sumur, (13)

25

GEOTHERMAL SYSTEMS ON THE ISLAND OF JAVA, INDONESIA

Ciawi, (14) Cilayu, (15) Pakenjeng, (16) Slamet Volcano, (17) Dieng, (18) Kalianget, (19) Ungaran, (20) Candi Dukuh, (21) Parangtritis, (22) Lawu, (23) Pacitan, (24) Arjuna-Welirang and (25) Segaran (Fig. 3.1). In total 70 samples were collected, 61 from hot springs, 4 from cold springs, 4 from hot crater lakes and 1 from the Indian Ocean (Table 3.1). The locations of the 4 cold spring samples were chosen based on their proximity to those hot springs which were sampled during this investigation.

III.3. Results The results of the field and laboratory measurements are presented in Table 3.1. Cold water springs in Java were slightly acid to slightly alkaline (pH= 6.2 to 7.8) and conductivity ranged from 86 to 324 μS/cm. Compared to the hot spring samples, the concentrations of Ca2+, Mg2+, Na+, K+ and Cl- of the cold spring waters were low (≤ - 2- 31 mg/L). These cold spring waters had HCO3 and SO4 contents of 19.5 to 115.9 mg/L and 2.7 to 40.6 mg/L, respectively. The volcano-hosted hot springs had a larger variety of temperature, pH, - 2- - + conductivity, major anions (HCO3 , SO4 , and Cl ) and two major cations (Na and Mg2+), but relatively a similar range of K+ and a smaller range of Ca2+, compared to the fault-hosted hot springs. The temperatures of the volcano-hosted hot springs ranged from 22 to 95 °C and those of the fault-hosted hot springs ranged from 47 to 102 °C. The volcano-hosted hot springs were very acid to slightly alkaline (pH= ~ 1 to 8.4), while of the fault-hosted hot springs were slightly acid to slightly alkaline (pH= 5 to 8.1). The conductivity of the volcano-hosted hot springs varied from 86 to 14600 μS/cm, compared to 1500 to 17340 μS/cm of the fault-hosted hot springs. The - concentration of HCO3 in the volcano-hosted hot springs ranged from below detection 2- - to 1634.8 mg/L, SO4 ranged from below detection to 3005.5 mg/L, and Cl ranged - from 6.9 to 8084 mg/L; and those of the fault-hosted hot springs had HCO3 2- concentration ranged from 22 to 1085.8 mg/L, SO4 ranged from below detection to 1284.5 mg/L, and Cl- ranged from 122.1 to 6184.5 mg/L. The concentration of Mg2+ in the volcano-hosted hot springs ranged from 2.6 to 211.9 mg/L, Na+ ranged from 2.2 to 2979 mg/L, K+ ranged from 1.4 to 119.8 mg/L, and Ca2+ ranged from 4.9 to 510.7 mg/L; while the concentration of Mg2+ in the fault-hosted hot springs ranged from below detection to 97.7 mg/L, Na+ ranged from 115.8 to 1797.4 mg/L, K+ ranged from below detection to 94.2 mg/L, and Ca2+ ranged from 32.8 to 2047.6 mg/L.

26

Fig. 3.1. The distribution of sampled geothermal systems on Java, i.e., (1) Cisolok, (2) Cikundul, (3) Batu Kapur, (4) Ciater, ((5)5) Maribaya, (6) Tampomas, (7) Patuha, (8) Pangalengan, (9) Darajat, (10) Kamojang, (11) Cipanas, (12) Kampung Sumur, (13) Ciawi,awi, (14) Cilayu, (15) Pakenjeng, (16) Slamet Volcano, (17) Dieng, (18) Kalianget, (19) Ungaran, (20) Candi Dukuh, (21) Parangtritis, ((22)22) Lawu, (23) Pacitan, (24) Arjuna-Welirang and (25) Segaran. Geological structures and volcanic belts were based on Hamilton (197799),), Simandjuntak and Barber (1996), Hoffmann-Rothe et al. (2001) and Soeria-Atmadja et al. (1994).

Table 3.1. Sampling locations, physicochemical and stable isotope compositions of cold springs, hot springs and hot acid crater lakes on Java.

18 2 Sample Location Geo. Temp. pH Ec TDS Ca Mg Na K Cl HCO3 SO4 NO3 Br Si B Li As O H ID Type (°C) (uS/cm) (mg/L) (μg/L) (‰)‰) Hot Springs J1 Pancuran 3 (Slamet volc.) V 46.3 6.2 4070 3995 189.6 204.2 358.0 75.4 732.5 652.7 599.6

Table 3.1. (continued)

18 2 Sample Location Geo. Temp. pH Ec TDS Ca Mg Na K Cl HCO3 SO4 NO3 Br Si B Li As O H ID Type (°C) (uS/cm) (mg/L) (μg/L) (‰)) Hot Springs J42 Pakenjeng F 59.9 7.4 1960 1367 225.5

GEOTHERMAL SYSTEMS ON THE ISLAND OF JAVA, INDONESIA

Both the volcano- and fault-hosted hot springs were characterized by relatively large variation of B, Li and As concentrations. The B concentration of the volcano- hosted systems ranged from 0.03 to 94.4 mg/L, Li ranged from 0.4 μg/L to 11.06 mg/L and As ranged from 0.3 μg/L to 9.5 mg/L. In the fault-hosted systems, the B concentrations ranged from 0.42 to 58.2 mg/L, Li ranged from 23.5 μg/L to 2.23 mg/L and As ranged from 1.2 μg/L to 3.5 mg/L. Most of the hot springs with high B, Li and As concentrations were chloride water. This phenomenon is common in geothermal systems, because neutral chloride waters ascend directly from the reservoir and thus are generally enriched in selected trace elements, i.e., the geothermal suite of elements (Goff and Janik, 2000; Nicholson, 1993; White et al., 1971). In addition to physicochemical parameters, stable isotopes of 2H and 18O were determined. The stable isotope composition of hot spring waters from the volcano- hosted systems had a larger variation than those from the fault-hosted systems. The 18O isotope composition of the cold springs ranged from -8.8 to -6.1 ‰, the volcano- hosted hot springs waters ranged from -9.3 to 7.9 ‰ and the fault-hosted hot springs waters ranged from -7.5 to -4.3 ‰ (Table 3.1). The 2H isotope composition of the cold springs ranged from -55.6 to -42.0 ‰, the volcano-hosted hot springs waters ranged from -62.3 to -4.1 ‰ and the fault-hosted hot springs waters ranged from -50.1 to -4.2 ‰ (Table 3.1).

III.4. Discussion III.4.1. General considerations about geothermal systems on Java As mentioned above, geothermal systems on Java were classified into volcano- hosted and fault-hosted. Based on this classification, from a total of 25 sampled geothermal systems, 8 were considered fault-hosted (i.e., Pacitan, Maribaya, Batu Kapur, Pakenjeng, Cilayu, Cikundul, Cisolok, and Parangtritis) and 17 were considered volcano-hosted (i.e., Segaran, Arjuna-Welirang Volcano, Lawu Volcano, Ungaran Volcano, Candi Dukuh, Dieng, Kalianget, Slamet Volcano, Ciawi, Kampung Sumur, Tampomas, Cipanas, Ciater, Darajat, Kamojang, Pangalengan, and Patuha) (Fig. 3.1). All of the volcano-hosted geothermal systems were in the Quaternary volcanic belt, while most of the fault-hosted geothermal systems were in the Tertiary volcanic belt (Fig. 3.1). Several of those fault-hosted geothermal systems are located in major fault zones, e.g. the Cisolok and Cikundul geothermal systems in the Cimandiri fault, the Maribaya geothermal system in the Lembang fault, the Parangtritis geothermal

30

GEOTHERMAL SYSTEMS ON THE ISLAND OF JAVA, INDONESIA

system in the Opak fault, and the Pacitan geothermal system in the Grindulu fault. The Batu Kapur, the Pakenjeng and the Cilayu geothermal systems are associated with minor faults (Fig. 3.1). Contrast with the other fault-hosted geothermal systems, the Maribaya and the Batu Kapur were located in the active Quarternary volcanic belt, thus probably are heated by volcanic activity. According to the geologic maps of Java (Alzwar et al., 1992a; Effendi et al., 1998a; Samodra et al., 1992; Silitonga, 1973a; Sujatmiko and Santosa, 1992a), the Pacitan, the Pakenjeng, the Cilayu, the Cisolok and the Parangtritis geothermal systems are situated close to the zones of the Tertiary intrusive rocks. The position of a fault-hosted geothermal field nearby an intrusive rock could be an indication of a magmatic heat source, an assumption that was further investigated using geochemical tools. The origin and physicochemical history of hydrothermal fluids can be explored in a Cl, SO4 and HCO3 ternary diagram (Chang, 1984; Giggenbach, 1991; Giggenbach, 1997; Nicholson, 1993). Based on their position in the diagram, hydrothermal waters can be divided into neutral chloride, acid sulfate and bicarbonate waters, but mixtures between the individual groups are common. On Java, bicarbonate was the dominant water type for the volcano-hosted hot springs, followed by acid sulfate and neutral chloride waters, while for the fault-hosted hot springs the water types were distributed more or less evenly (Fig. 3.2). The occurrence of different water types in a given hydrothermal system is common, indicating the different physicochemical processes, such as, phase separation and mixing in the shallow subsurface (Ellis and Mahon, 1977; Giggenbach, 1997; Hedenquist, 1990; Henley and Ellis, 1983; McCarthy et al., 2005). Although the bicarbonate water type seems to be more abundant in the volcano-hosted hydrothermal systems, a definite difference between the volcano- and fault-hosted systems is difficult to be assessed in a ternary diagram alone. Thus, following the procedure of Valentino and Stanzione (2003), the hot waters from Java were plotted in HCO3 vs. Cl and Mg/Na vs. SO4/Cl diagrams (Figs. 3.3 and 3.4). - These diagrams show that the volcano-hosted thermal waters have a higher HCO3 content and a higher Mg2+/Na+ ratio. This observation could result due to magmatic degassing and thus addition of CO2 to the volcano-hosted hot springs, which is likely absent or minor in the fault-hosted hot springs. The reaction between H2O and CO2 increases acidity and thus intensifies water-rock interaction (Giggenbach, 1984; Giggenbach, 1988; White, 1957). Acid conditions and low temperature, due to slow upward migration or a long flow path of the ascending thermal waters, increases the solubility of Mg2+ (Allen and Day, 1927), hence producing Mg-rich waters.

31

GEOTHERMAL SYSTEMS ON THE ISLAND OF JAVA, INDONESIA

Fig. 3.2. SO4-HCO3-Cl ternary diagram of cold and hot springs. Most of the volcano- - hosted hot springs were of the HCO3 water type, whereas the fault-hosted hot springs 2- - - were distributed evenly between the SO4 , HCO3 and Cl types. Open circle = cold spring, gray filled triangle = volcano-hosted hot spring and open square = fault-hosted hot spring.

32

GEOTHERMAL SYSTEMS ON THE ISLAND OF JAVA, INDONESIA

Fig. 3.3. HCO3 vs. Cl (in mg/L basis) diagram of cold and hot springs for: (A) and (B) formed in the margin of the ‘primary neutralization’, where (A) are closer than (B), (C) thermal waters from shallow depth, thus undergone major dilution by groundwater, (D) volcanic H2S gas oxidized by O2-rich groundwater, (E) and (F) fault-hosted hot springs, where (E) were less diluted by groundwater than the (F). (E) is likely influenced by seawater input. (Symbols are as in Fig. 3.2).

Fig. 3.4. Mg/Na vs. SO4/Cl (in meq/L basis) diagram of cold and hot springs. The interpretation of groups A, B, C, D, E and F are similar to those in Fig. 3.4. (Symbols are as in Fig. 3.2).

33

GEOTHERMAL SYSTEMS ON THE ISLAND OF JAVA, INDONESIA

Four groups of the volcano-hosted thermal waters are shown in the HCO3 vs. Cl and the Mg/Na vs. SO4/Cl diagrams, i.e., A, B, C and D (Fig. 3.3 and 3.4). The group - - A samples had the highest HCO3 and Cl concentrations, but the lowest SO4/Cl ratios. - - Conversely, the group D samples had the lowest HCO3 and Cl concentrations but the - - highest SO4/Cl ratios. Meanwhile, the group B and C had HCO3 and Cl concentrations and SO4/Cl ratios in between of the group A and D. However, the - - group C was split from the group B due to its lower HCO3 and Cl contents. The groups A and B are thought to have formed at the margin of the ‘primary neutralization’ zone (Giggenbach, 1988). There separation of CO2 and its reaction - with groundwater produces HCO3-rich thermal waters, while the Cl content remains high due to the lesser dilution by groundwater. The group A samples originated closer to the ‘primary neutralization’ zone than the group B samples and thus had a higher - Cl concentration. In the Mg/Na vs. SO4/Cl diagram, two acid thermal waters, J51 2- - (Kawah Putih) and J36 (Ciater), are exceptions in group B. The moderate SO4 /Cl ratios in these two samples were likely caused by H2S and HCl addition to shallow groundwater (Delmelle and Bernard, 1994; Delmelle et al., 2000; Giggenbach, 1988; Sriwana et al., 2000). The resulting low pH enhances water-rock interaction, thus causing the high Mg2+/Na+ ratio in those two samples. The group C samples were considered thermal waters formed in the shallow subsurface and thus were diluted by - - groundwater, which lowered their HCO3 and Cl contents. The group D samples were interpreted to be thermal waters influenced by primary H2S-rich magmatic vapor in the shallow subsurface, hence producing acid sulfate waters. In this group, J65 (Dieng) was an exception (Fig. 3.4), because the hot spring had a neutral (pH= 7) and was Cl- poor (14.7 mg/L). This thermal water is likely a mixture of HCO3-rich and SO4-rich water, which can occur in low relief liquid-dominated geothermal systems (Kuhn, 2004). Meanwhile, the thermal waters of the fault-hosted geothermal systems were divided into two groups, E and F, where the former had a higher Cl- content than the latter. Chloride is a conservative element and thus the Cl- variation in the fault-hosted thermal waters could be caused by either varying degrees of mixing with shallow groundwater or reflect the initial Cl- content of the hydrothermal fluid. Based on that assumption, group E thermal waters should have undergone less mixing with shallow groundwater than group F. This is also indicated in the Na-K-Mg ternary diagram (Giggenbach, 1988), where the group E thermal waters (J45 and J58) plot closer to equilibrium than the group F thermal waters (Fig. 3.6). The high Cl- concentration in sample J58 from the Parangtritis hot spring is unusual, but can be explained by

34

GEOTHERMAL SYSTEMS ON THE ISLAND OF JAVA, INDONESIA

seawater addition to the geothermal system, because of its proximity to the Indian - Ocean. Most of the acid sulfate waters had HCO3 concentrations, which were below detection and thus less samples could be plotted as group D in the HCO3 vs. Cl diagram than in the Mg/Na vs. SO4/Cl diagram. Conversely, many neutral chloride 2- waters were below the detection limit of SO4 and therefore more thermal waters plotted as group A in the HCO3 vs. Cl diagram than in the Mg/Na vs. SO4/Cl diagram.

Several anomalies and discrepancies were found in both the HCO3 vs. Cl and

Mg/Na vs. SO4/Cl diagrams. In the HCO3 vs. Cl diagram, the fault-hosted hot springs of Maribaya (J35) and Batu Kapur (J37 and J38) plot within the group B of the volcano-hosted thermal waters, whereas the volcano-hosted hot springs of J27 (Lawu) and J61 (Dieng) plot as a group of fault-hosted thermal waters (Fig. 3.3). As - mentioned previously, the HCO3 abundance in the Maribaya and Batu Kapur hot springs was likely caused by addition of magmatic CO2, due to their location within the - active Quaternary volcanic belt. Meanwhile, the lower HCO3 concentrations in samples J27 (Lawu) and J61 (Dieng) were likely the result of carbonate mineral precipitation. That sample J28 plots below group B in the Mg/Na vs. SO4/Cl diagram should be due to the formation of clay minerals and the associated depletion of Mg2+.

The discrepancy that samples J23 and J53 plot in group B in the HCO3 vs. Cl diagram and in group A in the Mg/Na vs. SO4/Cl diagram can be attributed to the removal of

SO4 due to precipitation of sulfate minerals. The same process is the likely cause for the location of sample J57, a dilute thermal water, in group B in the Mg/Na vs. SO4/Cl diagram. The Cl/B ratios of hydrothermal waters can be used to identify subsurface processes, such as, water-rock interaction, magma degassing and seawater feeding in a geothermal system (Arnorsson and Andresdottir, 1995; Valentino and Stanzione, 2003). The Cl/B ratios found in the samples from Java indicated three dominant processes for the volcano-hosted thermal waters, i.e., groundwater mixing, water-rock interaction with the andesitic host rock and phase separation. On the other hand, the Cl/B ratios of fault-hosted hot springs were generally affected by water-rock interaction with the andesitic host rock (Fig. 3.5). Geothermal water in the Dieng, Kamojang and Darajat geothermal fields had lower Cl/B ratio than the andesitic rock (Fig. 3.5), something that can be caused by phase separation in high temperature (>300 °C) reservoirs (Truesdell et al., 1989). This process removes B from the geothermal reservoir, thus relatively increasing the Cl- concentration of the remaining hydrothermal fluid (Arnorsson and Andresdottir,

35

GEOTHERMAL SYSTEMS ON THE ISLAND OF JAVA, INDONESIA

1995; Truesdell et al., 1989). One volcano-hosted hot spring, J36 (Ciater), and three fault-hosted hot springs, J11 (Pacitan), J12 (Pacitan) and J58 (Parangtritis), plotted closed to the seawater-precipitation line. The J58 (Parangtritis) sample also plotted in the HCO3 vs. Cl and Mg/Na vs. SO4/Cl diagrams in a way which would indicate seawater addition. However, seawater addition is not likely for the J12 and J13 hot springs due to their low Cl- concentration. Hence, the low B/Cl ratio of these hot springs could have been caused by B removal due to adsorption by clay minerals, particularly illite (Harder, 1970).

Fig. 3.5. Cl vs. B diagram of cold waters and geothermal waters. The Cl/B ratio of seawater from the Indian Ocean (this research) and andesitic rock from Trompetter et al. (1999) are used. The Cl/B ratio of volcano-hosted and fault-hosted hot springs were considered to be controlled by water-rock interaction. The Dieng, Kamojang and Darajat volcano-hosted geothermal systems underwent phase separation. Though J58 (Parangtritis), J11 and J12 of Pacitan and J36 (Ciater) plotted close to seawater, but only for J58 seawater mixing was indicated. (Symbols are as in Fig. 3.2).

36

GEOTHERMAL SYSTEMS ON THE ISLAND OF JAVA, INDONESIA

III.4.2. Geothermometry Solute geothermometers, such as listed in Table 3.2, can provide powerful tools to estimate subsurface conditions. Their successful application has been extensively discussed in the geothermal literature and relies on five basic assumptions: 1) exclusively temperature dependent mineral-fluid reaction; 2) abundance of the mineral and/or solute; 3) chemical equilibrium; 4) no re-equilibration; and 5) no mixing or dilution (e.g., Nicholson, 1993). The no mixing or dilution assumption, however, can be circumvented if their extent and/or influence on solute ratios (e.g., Na/K) are known. Several geothermometers were applied in order to analyze the physichochemical processes encountered by the hydrothermal fluids during their ascent to the surface. These processes include dilution by shallow water, conductive cooling, adiabatic cooling, mineral precipitation, adsorption/desorption, water-rock interaction and re-equlibrium (Fournier, 1977; Kaasalainen and Stefánsson, 2012). Different geothermometers record different equilibria and disagreement does not immediately eliminate the use of one or the other. Careful application and evaluation of calculated temperatures may provide important clues to the overall hydrology of the geothermal system (Pichler et al., 1999). Silica geothermometers, which are commonly applied to hot springs (Fournier, 1977) predicted a temperature range from 100 to 140 °C for most samples (Table 3.2). Lower temperatures were calculated for J60 (Kampung Sumur), J31 (Candi Dukuh), and J24 and J28 (Lawu). The J60 and J31 hot springs were located at the edge of a pool and a lake, respectively and thus, should be diluted by surface water. The two hot springs, J24 and J28, were Cl-rich, which would preclude dilution by surface or groundwater dilution. That lead to the conclusion that the J24 and J28 hot springs lost silica due to the precipitation of silicate minerals during ascent, causing the low silica geothermometer temperatures. When applied to hot springs, silica thermometers are known to predict closer to the discharge temperature, rather than the reservoir temperature (e.g., Pichler et al., 1999). This inherent problem can be overcome by calculating the silica ‘parent’ concentration using the silica mixing model of Fournier (1977). After application, the silica geothermometers predicted much higher reservoir temperatures of 258 °C for Slamet, 188 °C for Ciawi, 180 °C for Batu Kapur and 221 °C for Pangalengan. These temperatures were in the range of those, predicted by the Na/K and Na/K/Ca geothermometers (see below).

37

GEOTHERMAL SYSTEMS ON THE ISLAND OF JAVA, INDONESIA

Table 3.2. Calculated reservoir temperatures.

Location Samples Geothermal T Field Geothermometers (°C) o Na-K- ID Types ( C) Quartz Quartz Chalcedony Quartz Na/K Ca (steam (parents) loss) J1 46.3 128 125 101 290 217 J2 52.1 131 128 104 286 215 Slamet V 258 J33 40.5 110 110 81 380 241 J34 53.3 122 120 94 326 232 J3 43.2 129 126 102 313 223 Ciawi V 188 J4 53.4 132 128 105 308 222 J6 46.2 116 115 87 287 203 Cipanas J7 V 48.3 117 116 88 nd 284 205 J8 49.3 118 117 90 283 202 J13 48.3 122 119 93 304 219 J14 45.7 114 113 85 305 217 J16 46.1 102 102 72 318 218 Arjuna- J17 V 42.3 101 102 71 nd 319 213 Welirang J18 46.4 128 125 101 187 169 J19 28.4 129 126 101 188 180 J20 41.5 125 122 97 189 168 J21 44.9 121 119 93 282 222 Segaran V nd J22 22.3 118 117 90 283 221 J23 34.3 104 104 74 133 128 J24 38.4 95 96 64 150 155 J26 36.4 101 102 71 127 147 Lawu J27 V 34.4 104 104 74 nd 289 203 J28 33.2 79 83 48 125 126 J29 39.8 118 116 89 217 176 J30 41.0 120 118 92 199 166 Candi Dukuh J31 V 35.9 89 91 58 nd 204 165 J39 68.9 135 131 108 371 232 Pangalengan J40 V 39.8 118 117 90 221 250 180 J41 54.1 133 129 106 323 219 J46 38.9 112 111 83 305 216 Kalianget V nd J47 40.0 115 114 86 310 219 Patuha J52 V 32.9 125 123 98 nd 301 205 J53 64.3 132 128 105 212 172 Tampomas V nd J54 51.4 127 124 99 171 158 Kampung J57 V 35.0 95 96 64 nd 263 196 Sumur J60 57 101 102 71 306 213 J61 54 139 134 113 431 261 J62 70 136 132 109 599 296 Dieng J63 V 56 132 129 105 nd 354 233 J64 60 126 123 98 349 236 J66 32 107 107 77 474 248 J69 27 104 104 74 481 242 J11 51.3 56 62 24 Pacitan F nd nd nd J12 37.3 54 63 24 Maribaya J35 V 46.5 127 124 99 nd 299 203 J37 56.3 130 127 103 278 221 Batu Kapur V 180 J38 40.9 122 120 93 250 195 J42 59.9 75 79 44 Pakenjeng F nd nd nd J43 43.1 74 79 43 J44 70.3 125 122 97 177 176 Cilayu F nd J45 45.1 127 124 99 167 167 Cikundul J48 F 50.5 89 91 58 nd 111 111 J49 102.0 115 114 87 143 133 Cisolok F nd J50 100.0 110 109 80 139 128 Parangtritis J58 F 39.2 74 78 42 nd 88 55 *V= volcano-hosted, F= fault-hosted, nd= not defined

38

GEOTHERMAL SYSTEMS ON THE ISLAND OF JAVA, INDONESIA

Reservoir temperature calculations with the Na/K and Na/K/Ca geothermometers were conducted according to Giggenbach (1988), by first evaluating if equilibrium between host-rock and hydrothermal fluid was attained. Only six samples, J24 and J26 (Lawu volcano), J44 and J45 (Cilayu), J48 (Cikundul), and J58 (Parangtritis) were in partial equilibrium, and four samples, J23 and J28 (Lawu), and J49 and J50 (Cisolok) were close to partial equilibrium with their respective host rocks (Fig. 3.6). The Na/K geothermometer was then applied for these ten samples and calculated reservoir temperatures were 127 to 150 °C for Lawu, 167 to 177 °C for Cilayu, 111 °C for Cikundul, 88 °C for Parangtritis and 139 to 143 °C for Cisolok.

Fig. 3.6. Na-K-Mg ternary diagram Giggenbach (1988) for the Java hot springs. Several fault-hosted and a few of volcano-hosted hot springs were close to or in partial equilibrium with the host-rock. (Symbols are as in Fig. 3.2).

39

GEOTHERMAL SYSTEMS ON THE ISLAND OF JAVA, INDONESIA

Fournier (1989) respectively proposed the use of the Na/K geothermometer and the Na-K-Ca geothermometer for the prediction of the highest and the lowest reservoir temperatures in a geothermal system. Following this approach, calculated temperatures ranged from 205 to 301 °C in Patuha and 219 to 323 °C in Pangalengan geothermal systems. These results were in good agreement with the predicted reservoir temperatures by direct measurement, i.e., 209 to 241 °C for Patuha (Layman and Soemarinda, 2003) and 250 to 300 °C for Pangalengan (Abrenica et al., 2010; Layman and Soemarinda, 2003). Reliable reservoir temperature prediction with those two geothermometers was also indicated in the Dieng geothermal system, where calculated temperatures ranged from 236 to 349 °C. These temperatures were relatively similar to the predictions by Prasetio et al. (2010) (i.e, 240 to 333 °C), who used gas geothermometers. Therefore, based on those observation, the two geothermometers were applied to the remaining geothermal systems and calculated temperatures in Arjuna-Welirang ranged from 217 to 305 °C, in Cipanas from 202 to 277 °C, in Segaran from 221 to 283 °C, in Kalianget from 216 to 305 °C, in Tampomas from 172 to 212 °C and in Maribaya from 203 to 299 °C. The K+ concentration in the Pacitan and Pakenjeng hot springs waters were below detection limit, thus the Na/K and Na/K/Ca geothermometers could not be applied. Silica geothermometers resulted in a maximum temperature of 63 °C for Pacitan and 79 °C for Pakenjeng geothermal systems. Those temperatures although likely lower than the actual reservoir temperatures indicated reservoir temperatures below 100 °C. A compilation of all calculated reservoir temperatures on Java are presented in Table 3.3. In the Darajat and Kamojang geothermal systems, the only surface expressions were acid sulfate-type hot springs. That type of hydrothermal fluid reacts extensively with near surface rocks, hence chemical geothermometers could not be applied (e.g. Nicholson, 1993). A reservoir temperature of 280 °C was predicted by Hadi (1997) for the Darajat geothermal system, based on the alteration minerals observed in drill cores and Sudarman et al. (1995) reported a shallow reservoir temperature measurement of 232 °C for the Kamojang geothermal system.

40

GEOTHERMAL SYSTEMS ON THE ISLAND OF JAVA, INDONESIA

Table 3.3. Compilation of calculated geothermal reservoir temperatures on Java.

Geothermal Geothermal T (oC) Geothermometer systems Types Slamet M. V 258 to 380 Si parent and Na-K Ciawi V 188 to 313 Si parent and Na-K Cipanas V 202 to 287 Na-K and Na-K-Ca Arjuna-Welirang M. V 217 to 305 Na-K and Na-K-Ca Segaran V 221 to 283 Na-K and Na-K-Ca Lawu M. V 127 to 150 Na-K Candi Dukuh V 165 to 204 Na-K and Na-K-Ca Pangalengan V 221 to 323 Si parent and Na-K Kalianget V 216 to 310 Na-K and Na-K-Ca Patuha V 205 to 301 Na-K and Na-K-Ca Tampomas V 172 to 212 Na-K and Na-K-Ca Kampung Sumur V 196 to 263 Na-K and Na-K-Ca Dieng V 236 to 349 Na-K and Na-K-Ca Pacitan V <100 Si Maribaya F 203 to 299 Na-K and Na-K-Ca Batu Kapur F 180 to 278 Si parent and Na-K Pakenjeng F <100 Si Cilayu F 125 to 177 Na-K Cikundul F 111 Na-K Cisolok F 139 to 143 Na-K Parangtritis F 88 Na-K *V= volcano-hosted, F= fault-hosted

III.4.3. The heat sources of the fault-hosted geothermal systems The Quaternary volcanic arc could be the heat source for the fault-hosted geothermal systems. That case was investigated by comparing the enrichment of conservative trace elements and reservoir temperatures between the volcanic and fault-hosted geothermal fields. A similar procedure was applied to the Steamboat geothermal system (Arehart et al., 2003), which pointed towards a magmatic heat source rather than just enhanced heat flow. Lithium was considered as the most

41

GEOTHERMAL SYSTEMS ON THE ISLAND OF JAVA, INDONESIA

conservative trace element in this study, because the B concentration in several hot spring waters were affected by phase separation. The Li vs. Cl diagram shows that some of the fault-hosted hot springs have similar high trends of Li enrichment as the volcano-hosted hot springs (Fig. 3.7). The slopes of Li enrichment of the volcano-hosted hot springs are ~0.003, but ~0.003 and ~ 0.0001 for the fault-hosted hot springs. The similarity of high Li enrichment between the fault-hosted and volcano-hosted geothermal fields could be an indication of the same type of heat source, i.e, magmatic. Nevertheless, this assumption has to be corroborated, because the trace element enrichment in hot springs can be generated by several other processes, i.e., (1) high trace element concentration of the host rock, (2) intermediate age and (3) high temperature of the geothermal system (Arehart et al., 2003). The first point was ruled out, because the geothermal host rocks on Java are basically identical, i.e., andesitic rocks. However, the second and the third points were evaluated by considering the calculated reservoir temperatures. The similarity of the high Li enrichment trend and the similarity high reservoir temperatures of fault- hosted and volcano-hosted geothermal systems was considered an indication of a magmatic heat source for both. Meanwhile, an intermediate age of a fault-hosted geothermal system can be concluded when its reservoir temperature is lower than that of a volcano-hosted geothermal system and its Li enrichment is as high that of a volcano-hosted geothermal system. An intermediate age geothermal system has a relatively higher trace element concentration due to the extended period of water-rock interaction. Young geothermal systems have less time of water-rock interaction and old geothermal systems have already leached most trace elments from their host rocks, thus both systems have relatively low trace element concentrations (Arehart et al., 2003). Considering those assumptions, the fault-hosted geothermal systems of Cilayu, Batu Kapur, Maribaya and Cisolok should be heated by a magmatic heat source. In contrast, such heat source was not likely for the Cikundul and Pakenjeng fault-hosted geothermal systems. Hence, the high Li enrichment of these two fault- hosted geothermal systems was caused by their intermediate age. All of the low temperature and Li-poor fault-hosted geothermal systems, Cikundul, Pakenjeng, Parangtritis and Pacitan, are located in the southern part of the Java island. This area consists of the Tertiary volcanic belt, where volcanism ceased in the last Paleogene. The volcanism then shifted northward forming the Neogene and Quaternary volcanic belts in the central part of the island (Soeria-Atmadja et al., 1994). Under those geological conditions, the heat source of those fault-hosted

42

GEOTHERMAL SYSTEMS ON THE ISLAND OF JAVA, INDONESIA

geothermal system was likely similar to those described as ‘amagmatic’ heat source in the Great Basin (USA) and Western Turkey (Faulds et al., 2010). An ‘amagmatic’ heat source is a deep seated magma, which remained after volcanism ceased. The name was used to distinguish this heat source from the shallow magmatic heat sources of the Quaternary. As indicated by the exposure of the Cretaceous basement, the southern part of the Java island underwent uplift and erosion due to the subduction of Indo-Australia and Eurasian plates (Clements et al., 2009). As a result the crust became thinner, which in turn increased the heat gradient, causing thermal circulation of groundwater along faults, thus generating the fault-hosted geothermal systems. The same heating mechanism was suggested for geothermal systems along the Alpine fault, New Zealand (Allis and Shi, 1995; Shi et al., 1996).

Fig. 3.7. Li vs. Cl diagram of volcano-hosted and fault-hosted hot springs. Most of fault-hosted hot springs had similar trends of Li enrichment to those of volcano- hosted hot springs. (Symbols are as in Fig. 3.2).

43

GEOTHERMAL SYSTEMS ON THE ISLAND OF JAVA, INDONESIA

III.4.4. Oxygen and hydrogen isotope considerations The deuterium and oxygen isotopic composition of hot springs can be used to investigate their origin (Arnason, 1977; Craig, 1963; Craig et al., 1956; Giggenbach, 1978; Giggenbach et al., 1983; Majumdar et al., 2009; McCarthy et al., 2005; Pichler, 2005). All of the fault-hosted hot springs and most of the volcano-hosted hot springs plotted close to the Local Meteoric Water Line (LMWL), indicating meteoric water as the source of the hydrothermal fluids (Fig. 3.8). Ten volcano-hosted geothermal waters, J9, J10, J19, J24, J26, J32, J51, J59 and J65, show stable isotope enrichment. This could be caused by either evaporation, water-rock interaction, the input of magmatic fluids input or any combination of the above (Craig, 1966; D'Amore and Bolognesi, 1994; Giggenbach and Stewart, 1982; Ohba et al., 2000; Varekamp and Kreulen, 2000). Using the formulas from Gonfiantini (1986) and Varekamp and Kreulen (2000), a theoretical evaporation line was calculated for 90 °C lake temperature, 22 °C ambient temperature, 80% atmosphere humidity and a δ18O of -6.9 ‰ and δ2H of -45 ‰ for meteoric water. Samples J10, J26, J32 and J65 plotted close to the evaporation line although only J10, J32 and J65 were affected by evaporation, while the stable isotope enrichment in J26 was likely caused by addition of andesitic water of Taran et al. (1989) and Giggenbach (1992). The hot acid crater lakes of Kawah Kamojang (J9), Kawah Putih (J51) and Kawah Sikidang (J59) have the heaviest isotope composition and plotted above the field of andesitic water (Fig. 3.8). Connecting the hot crater lakes with their respective meteoric water (δ18O = -6.9 ‰ and δ2H = -45 ‰) generates a line, which is flatter than the evaporation line (Fig. 3.9). The slope of this line is close to the slope of other hot acid crater lakes lakes, such as, Khusatsu-Shirane volcano (Ohba et al., 2000), Poas volcano (Rowe Jr, 1994), (Varekamp and Kreulen, 2000) and Kawah Ijen (Delmelle et al., 2000). Thus, the isotopic composition indicated that the hot crater lake fluids likely underwent substantial evaporation and some reaction with magmatic gas. The presence of andesitic water in the geothermal systems was indicated for samples J19 (Candi Songgoriti 2), J24 (Banyuasin) and J60 (Kawah Sileri), which plot on or near the mixing line between local groundwater and andesitic water (Fig. 3.8). However, Figure 3.10 only indicates andesitic water input for J19, J24 and J26, but not for J60. The plot of J26 on the theoretical evaporation line in Figure 3.9 is caused by its lighter stable isotope composition compared to the other hot springs with andesitic water mixing. Andesitic water input was also found, for example, for the

44

GEOTHERMAL SYSTEMS ON THE ISLAND OF JAVA, INDONESIA

Meager Creek (Clark et al., 1982), Larderello (D'Amore and Bolognesi, 1994), Geyser (D'Amore and Bolognesi, 1994), Tongonan (Gerardo et al., 1993), El Chicon volcano (Taran et al., 2008) and Tutum Bay (Pichler et al., 1999) geothermal systems (Fig. 3.9). In addition, the elevated Cl- concentration in samples J19, J24 and J26 corroborate the presence of andesitic water in those geothermal systems. In contrast to these three samples, the presence of andesitic water could not be confirmed for sample J60, due to its low Cl- concentration. Hence, similar to the J10, J32 and J65 hot springs, stable isotope enrichment of J60 should have been caused by evaporation. The fact that J60 plots below the evaporation is likely due to a lighter stable isotope composition of its meteoric source water compared to that of the meteoric source water, which was used for the calculation of the theoretical evaporation line.

Fig. 3.8. δ2H and δ18O compositions of cold and hot springs. LMWL and GMWL were taken from (Wandowo et al., 2001) and (Craig, 1961), respectively. All of fault-hosted and most of volcano-hosted hot springs had a meteoric water origin. Three stable isotope enrichments were identified, i.e., evaporation, combination of magmatic gases input with evaporation and andesitic water input. (Symbols are as in Fig. 3.2).

45

GEOTHERMAL SYSTEMS ON THE ISLAND OF JAVA, INDONESIA

Fig. 3.9. δ2H and δ18O correlation lines between hot springs (gray filled) to their associated cold springs (open) from Java. The dashed lines are stable isotope correlation between thermal waters and their respective meteoric waters from other locations around the World.

III.5. Conclusions Based on the geological setting, two types of geothermal systems were identified on Java, volcano-hosted and fault-hosted. Contribution of CO2 to the geothermal fluid in volcano-hosted geothermal system, which was absent/minor in fault-hosted geothermal system, led to a different water chemistry. Volcano-hosted hot - 2+ + springs had higher HCO3 concentration and a higher Mg /Na ratio than fault-hosted host springs. While the B concentration of fault-hosted hot springs was only affected by water-rock interaction, volcano-hosted hot springs were influenced by phase separation, water-rock interaction and groundwater mixing. Seawater addition was

46

GEOTHERMAL SYSTEMS ON THE ISLAND OF JAVA, INDONESIA

identified in the Parangtritis, which was considered a fault-hosted geothermal system. Calculated reservoir temperatures of volcano-hosted geothermal systems ranged from 125 to 338 °C and those of fault-hosted geothermal systems ranged from 74 to 299 °C. Several geothermal systems on Java, although fault-hosted are likely heated by shallow magmas, i.e., Batu Kapur, Maribaya, Cilayu and Cisolok. The addition of volcanic fluids in Batu Kapur and Maribaya, which are located in the active Quarternary volcanic belt, indicated that these two geothermal systems in fact are volcano-hosted geothermal systems. Those fault-hosted geothermal fields that were located in the old (Neogene) volcanic belt did not experience any addition of volcanic fluids. Shallow magmas heat sources were not indicated in fault-hosted geothermal systems of Cikundul, Pakenjeng, Pacitan and Parangtritis. Thus, deep seated magma heat sources were suggested for those four geothermal systems. Stable isotope enrichments were found in ten of the volcano-hosted geothermal systems, but not in any of the fault-hosted geothermal systems. Stable isotope enrichment due to evaporation was recognized in the Kawah Candradimuka and Kawah Sileri, Kawah Hujan and Candi Gedong Songo geothermal systems. A combination of intensive evaporation and magmatic gases input produced very heavy stable isotopes in the hot acid crater lakes of the Kawah Kamojang, Kawah Sikidang and Kawah Putih geothermal systems. The addition of substantial amounts of Andesitic water to the geothermal fluid was observed in the Candi Songgoriti, Banyuasin and Pablengan geothermal systems Those finding reject the general assumption of a low energy potential of fault- hosted geothermal systems, since although fault-hosted their heat source can be magmatic as seen for several of the fault-hosted geothermal systems on Java. This should give a new perspective for geothermal exploration on Java, where to date, fault-hosted geothermal system were excluded from the geothermal energy development program.

Acknowledgements A part of this study was supported by the Ministry of Energy and Mineral Resources of Indonesia in a form of PhD scholarship grants. Thanks to Seto and Maman for their help in the sampling campaign. TP acknowledges funding from the German Research Foundation (DFG). M. Rowe and two anonymous reviewers are thanked for their fruitful input and comments on the first version of this manuscript.

47

GEOTHERMAL SYSTEMS ON THE ISLAND OF JAVA, INDONESIA

48

BORON ISOTOPE VARIATIONS IN GEOTHERMAL SYSTEMS ON JAVA, INDONESIA

IV. Boron isotope variations in geothermal systems on Java, Indonesia

Modified from:

Budi Joko Purnomo a, Thomas Pichler a, & Chen-Feng You b

a Geochemistry & Hydrogeology, Department of Geosciences, University of Bremen, Germany b Department of Earth Sciences, National Cheng Kung University, Tainan, Taiwan, ROC

Submitted to:

Geochimica et Cosmochimica Acta

49

BORON ISOTOPE VARIATIONS IN GEOTHERMAL SYSTEMS ON JAVA, INDONESIA

Abstract

This paper presents δ11B data for hot springs, hot acid crater lakes, geothermal brines and a steam vent from Java, Indonesia with emphasis to investigate the difference between boron (B) isotope compositions in volcano-hosted geothermal systems and fault-hosted geothermal systems, as well as differences between sulfate crater lakes and acid-chloride crater lakes. The possible seawater input into geothermal systems and the mechanisms of B isotope fractionation were also investigated. The δ11B values of hot springs ranged from -2.4 to +28.7 ‰ and hot acid crater lakes ranged from +0.6 to +34.9 ‰. The δ11B and Cl/B values in waters from the Parangtritis and Krakal geothermal systems indicated seawater input. The δ11B values of acid sulfate crater lakes ranged from +5.5 to +34.9 ‰ and were higher than the δ11B of +0.6 ‰ of the acid chloride crater lake. The heavier δ11B in the acid sulfate crater lakes was caused by a combination of vapor phase addition and further enrichment due to evaporation and B adsorption onto clay minerals. In contrast, the light δ11B of the acid chloride crater lake was a result of continuous magmatic fluid input. The fluids in volcano-hosted geothermal systems had a lighter δ11B signature than those in fault- hosted geothermal systems. The faster ascent and the insignificant input of magmatic fluids in the fault-hosted geothermal systems resulted in a light δ11B signature. In contrast, the input of magmatic fluids and the slower ascent in volcano-hosted geothermal systems were favorable for B isotope fractionation towards heavier values.

Keywords: Java, boron isotope, volcano- and fault-hosted geothermal systems, crater lake, seawater input

50

BORON ISOTOPE VARIATIONS IN GEOTHERMAL SYSTEMS ON JAVA, INDONESIA

IV.1. Introduction Geothermal waters are known to have a large range of δ11B, from -9.3 to +44 ‰ (Aggarwal et al., 2000; Aggarwal et al., 1992; Barth, 1993; Leeman et al., 1990; Musashi et al., 1988; Palmer and Sturchio, 1990; Vengosh et al., 1994b). This range is caused by the δ11B signature of the geothermal host-rock, seawater input, groundwater mixing and B isotope fractionation. Different sources of rocks were identified producing a variety of δ11B in geothermal fluids at the Argentine Puna Plateau (Kasemann et al., 2004). The heavy δ11B of seawater, +39.6 ‰ (Foster et al., 2010), was successfully used to identify seawater components in geothermal waters on Iceland (the Reykjanes and Svartsengi geothermal fields) and in Japan (the Izu- Bonin arc, Kusatsu-Shirane area, and Kagoshima) (Aggarwal and Palmer, 1995; Aggarwal et al., 2000; Kakihana et al., 1987; Millot et al., 2009; Musashi et al., 1988; Nomura et al., 1982; Oi et al., 1993). Shallow groundwater dilution potentially increases the δ11B composition of thermal waters (Palmer and Sturchio, 1990; Yuan et al., 2014). Fractionation of B isotopes in thermal waters occurred due to adsorption/incorporation of B onto clay minerals and iron oxide (Lemarchand et al., 2007; Palmer et al., 1987; Schwarcz et al., 1969; Spivack and Edmond, 1987; Vengosh et al., 1991b), calcite (Hemming and Hanson, 1992; Vengosh et al., 1991a) and evaporite minerals (Agyei and McMullen, 1968; McMullen et al., 1961; Oi et al., 1989; Swihart et al., 1986; Vengosh et al., 1992). These processes enrich the 10B isotope in the solid phases and thus increase the δ11B of thermal waters. Thermal waters which condensated from the vapor phase of a geothermal system are potentially enriched in δ11B as a consequence of 11B fractionation into the vapor phase. The δ11B enrichment during this process was generally considered insignificant (Kanzaki et al., 1979; Leeman et al., 1992; Nomura et al., 1982; Spivack et al., 1990; Yuan et al., 2014), but potentially should not be neglected in vapor- dominated geothermal systems. Java has many hot springs, steam vents, mud pools, hot acid crater lakes and altered grounds. Purnomo and Pichler (2014) divided the geothermal systems into fault-hosted and volcano-hosted geothermal systems, based on their location, either in a volcanic complex or a fault zone. While in fault-hosted geothermal systems deep circulating groundwater is simply heated, volcano-hosted geothermal systems seem to be more complex due to the addition of magmatic fluids to the geothermal water (Alam et al., 2010; Purnomo and Pichler, 2014).

51

BORON ISOTOPE VARIATIONS IN GEOTHERMAL SYSTEMS ON JAVA, INDONESIA

The focus of previous B isotope studies in geothermal systems were generally the mechanisms of B isotopes fractionation, while in this paper B isotope signatures were rather used as tracers to investigate contrasting systems, i.e., (1) volcano- hosted vs. fault-hosted geothermal systems, (2) acid sulfate vs. acid chloride crater lakes and (3) meteoric vs. seawater influenced.

IV.2. Sampling locations Thermal water samples for this study were taken from 21 geothermal systems on Java consist of 15 volcano-hosted and 6 fault-hosted geothermal systems (Fig. 4.1). All fault-hosted geothermal systems are distributed in the Tertiary volcanic belt. The Cikundul and the Parangtritis geothermal systems are hosted in the major fault zones of Cimandiri and Opak, respectively (Effendi et al., 1998; Rahardjo et al., 1995). Other fault-hosted geothermal systems, i.e., Cisolok, Cilayu, Pakenjeng and Krakal, are hosted in minor fault zones (Alzwar et al., 1992; Asikin et al., 1992; Silitonga, 1973; Sujatmiko and Santosa, 1992). Meanwhile, all volcano-hosted geothermal systems on Java are within the Quaternary volcanic complex. The Kamojang, Darajat and Wayang-Windu are located in the Kendang volcanic complex (Rejeki et al., 2005). The Patuha geothermal system is located in the flat volcanic highland of the Patuha volcano (Layman and Soemarinda, 2003). The Sari Ater geothermal system is hosted by the Tangkuban Prahu volcano, Cileungsing by the Tampomas volcano and Segaran by the Lamongan volcano. The Gucci and Baturaden geothermal systems are hosted by the Slamet volcano. Another single volcano, the Lawu volcano is hosting the Lawu geothermal system. The Songgoriti and Padusan geothermal systems are located in the Arjuna-Weilrang volcano complex and the Dieng geothermal system is located in the Dieng caldera.

IV.3. Results - - 11 The temperature, pH, Cl , B, HCO3 , Fe data and δ B values are reported in Table 4.1. Except for δ11B, this data is from Purnomo and Pichler (2014). The physicochemical data and δ11B values of new samples from the Kawah Kreta steam vent (J73), the Krakal hot spring (J74), the Kawah Domas acid sulfate crater lake (J72) and two geothermal brines of well AFT-28 (J70) and PAD-7C (J71) from the Dieng geothermal field are reported separately in Table 4.2.

52

Fig. 4.1. The sampling location of geothermal fields on Java, i.e., (1) Cisolok, (2) Cikundul, (3) Batu Kapur, (4) Tangkuban Parahu, (5)) Tampomas, (6) Patuha, (7) Pangalengan, (8) Darajat, (9) Kamojang, (10) Cipanas, (11) Ciawi, (12) Cilayu, (13) Pakenjeng, (14) Slamet,t, (15) Krakal, (16) Dieng, (17) Kalianget, (18) Parangtritis, (19) Lawu, (20) Arjuna-Welirang and (21) Segaran. Geological structures andd volcanic belts were based on Hamilton (1979), Simandjuntak and Barber (1996), Hoffmann-Rothe et al. (2001) and Soeria-Atmadja et al.l. (1994). Modified from Purnomo and Pichler (2014).

BORON ISOTOPE VARIATIONS IN GEOTHERMAL SYSTEMS ON JAVA, INDONESIA

- - 11 Table 4.1. Temperature, pH, Cl , HCO3 , B and Fe data and δ B of hot springs, hot acid crater lakes and cold springs from Java.

11 Sample Location Geo. Temp. pH Cl HCO3 B Fe δ B ID Type °C mg/L ‰ Hot springs J2 Baturaden (Slamet volc.) V 52 6.9 777.3 722.2 4.3 3.0 5.9 J4 Ciawi V 53 6.7 165.3 976.0 6.8 0.6 -0.7 J7 Cipanas V 48 6.3 119.0 383.1 2.4 0.01 4.6 J10 Kawah Hujan (Kamojang) V 95 4.9 7.2 22.0 4.6 0.1 2.3 J13 Padusan (Arjuna-Welirang volc.) V 48 6.5 246.2 1104.1 5.0 1.7 12 J18 Songgoriti (Arjuna-Welirang volc.) V 46 6.3 1303.5 1378.6 50.6 6.2 7.1 J21 Segaran (Lamongan volc.) V 45 6.5 550.5 1625.0 21.4 0.03 6.7 J24 Banyuasin (Lawu volc.) V 38 6.1 5948.7 835.7 93.2 9.4 8.1 J28 Kondo (Lawu volc.) V 36 6.3 4382.4 1634.8 55.6 1.9 4.5 J30 Ngunut (Lawu volc.) V 33 6.4 904.9 878.4 17.4 0.02 1.3 J34 Gucci (Slamet volc.) V 53 7.5 52.6 556.3 7.2 0.2 7.8 J36 Sari Ater (Tangkuban Parahu volc.) V 47 2.0 822.5

54

Table 4.2. Physicochemical data and δ11B values of the Krakal hot spring, the Kawah Domas crater lake, the Kawah Kreta steam vent and geothermal brines of the Dieng geothermal field.

11 Sample Location Geo. Temp. pH ORP Cond TDS Ca Mg Na K Cl HCO3 SO4 Si Sr Al B Mn Fe δ B ID Type °C mV uS/cm mg/L ‰ Hot springs J74 Krakal, Kebumen F 39 8.2 -86 21470 20170 2082.9

Geothermal brines J70 GW AFT-28 (Dieng) V 150 6.3 -196 59500 64000 893.5

J73 Kawah Kreta V nm nm nm nm nm 2.0

BORON ISOTOPE VARIATIONS IN GEOTHERMAL SYSTEMS ON JAVA, INDONESIA

The B concentrations of hot springs and acid crater lakes had a similar large range, from 2.1 to 93.2 mg/L and from 1.4 to 94.4 mg/L, respectively. The two geothermal brines from the Dieng geothermal field had B contents of 262.5 and 593.6 mg/L. The Kawah Kreta steam vent (J73) had a B content of 3.5 mg/L, which is enriched relative to its Cl- concentration, which was below detection. The B concentration in the acid sulfate crater lakes ranged from 1.4 to 73.3 mg/L, while the Kawah Putih (J51) acid chloride crater lake had a slightly higher value of 94.4 mg/L. The δ11B of the Cangar cold spring (J15) was +11.1 ‰, within the range of the thermal waters of -2.4 to +34.9 ‰, but heavier than the Kawah Kreta steam vent of +3.8 ‰ and the Dieng geothermal brines of 0 to 0.3 ‰. In accordance with the B concentration, the hot springs and hot crater lakes had relatively similar ranges of δ 11B, i.e., -2.4 to +28.7 ‰ and +0.6 to 34.9 ‰, respectively. In contrast to the B content, acid sulfate crater lakes had relatively heavy δ11B values from +5.6 to 34.9 ‰compared to the acid chloride crater lake, which had a value of +0.6 ‰. The fault- hosted and volcano-hosted geothermal systems had a relatively similar large ranges of δ11B, i.e., -2.4 to +28.7 ‰ and -1.0 to +34.9 ‰, respectively. The J74 and J58 fault- hosted hot springs as well as J59 acid sulfate crater lake had a δ11B value close to that of seawater, i.e., +39.6 ‰ (Foster et al., 2010). The δ11B of the thermal waters - - were poorly correlated with their temperature, pH, Cl , B, HCO3 and Fe (Fig. 4.2). Thermal waters with temperatures above 65 °C generally had relatively light δ11B values, close to 0 ‰, while below this temperatures most of the thermal waters were δ11B enriched (Fig. 4.2a). This indicates more significant of B isotope enrichment at low temperatures, as suggested by others, e.g., Palmer et al. (1987) and Aggarwal and Palmer (1995).

IV.4. Discussion IV.4.1. Boron in thermal waters and seawater input Following the procedure of Arnorsson and Andresdottir (1995) the B content of thermal waters on Java was used to identify steam separation for J9, J10, J34, J39, J55, J59 and J64; andesitic host-rock leaching for most of the thermal waters; seawater input for J58; and B adsorption for J2, J36, J38, J47, J49 and J54 (Purnomo and Pichler, 2014). The Krakal (J74) fault-hosted hot spring plotted close to the seawater line, similar to J58, indicating seawater input, while the Kawah Domas acid sulfate crater lake (J72) plotted in the vapor phase separation and the geothermal

56

BORON ISOTOPE VARIATIONS IN GEOTHERMAL SYSTEMS ON JAVA, INDONESIA

brines from Dieng, J70 and J71, plotted close to the andesitic rock leaching line (Fig. 4.3).

11 11 11 11 11 Fig. 4.2. a) δ B vs. T, b) δ B vs. pH, c) δ B vs. Cl, d) δ B vs. B, e) δ B vs. HCO3 and f) δ11B vs. Fe diagrams of thermal waters on Java show poor correlations of δ11B - - 11 11 with T, pH, Cl , B, HCO3 and Fe. The δ B vs. T indicates generally lower δ B enrichment at temperatures above 65 °C.

57

BORON ISOTOPE VARIATIONS IN GEOTHERMAL SYSTEMS ON JAVA, INDONESIA

Fig. 4.3. Cl vs. B diagram of thermal waters (modified from Purnomo and Pichler, 2014). Thermal waters underwent seawater input, B depletion, steam phase separation and andesitic rock leaching. Seawater input is indicated for two fault- hosted geothermal systems, J58 and J74, while most thermal waters were resulted by andesitic rock leaching.

The boron isotope composition can be a powerful tool to detect seawater input into geothermal systems, as demonstrated for two geothermal fields on Iceland (the Reykjanes and Svartsengi) and three areas in Japan (the Izu-Bonin arc, Kusatsu- Shirane area, and Kagoshima) (Aggarwal and Palmer, 1995; Aggarwal et al., 2000; Kakihana et al., 1987; Millot et al., 2009; Musashi et al., 1988; Nomura et al., 1982; Oi et al., 1993). A similar approach was used to investigate seawater input in the fault- hosted thermal waters at Krakal (J74) and Parangtritis (J58). These two thermal waters plot close to the mixing line between seawater and groundwater (Fig. 4.4), which proves the presence of seawater. The acid sulfate crater lake of Kawah Sikidang (J59) also had a heavy δ11B of +34.9 ‰. However, this acid sulfate crater lake had a Cl/B ratio five magnitudes lower than seawater (Fig. 4.4), thus seawater

58

BORON ISOTOPE VARIATIONS IN GEOTHERMAL SYSTEMS ON JAVA, INDONESIA

input could be excluded. The fluid source of Kawah Sikidang, which is the deep reservoir of the Dieng geothermal field, J70 and J71, also showed the absence of seawater input. The geothermal brines had a light δ11B of approximately 0 ‰ and Cl/B ratios two magnitudes lower than seawater (Fig. 4.4). If there was seawater input during the vapor phase ascent, the resulting thermal water should have a higher Cl- content corresponding to the vapor/seawater mixing ratio.

Fig. 4.4. δ11B vs. B/Cl diagram of thermal waters, groundwater and geothermal brines. Seawater input is confirmed for two fault-hosted thermal waters, J58 and J74, but not for the Kawah Sikidang (J59) acid sulfate crater lake. The δ11B value of +39.6 ‰ for seawater (Foster et al., 2010) and B/Cl ratio of the Indian Ocean (Purnomo and Pichler, 2014) are used in this diagram.

59

BORON ISOTOPE VARIATIONS IN GEOTHERMAL SYSTEMS ON JAVA, INDONESIA

IV.4.2. The δ11B of acid sulfate and acid chloride crater lakes - Two species of B, B(OH)3 and B(OH)4 exist in natural waters and only the former species is present at a pH of less than 7 (Dickson, 1990; Xiao et al., 2013). - B(OH)3 species produce a higher degree of B isotope fractionation than B(OH)4 , hence B isotope fractionation is stronger at low pH (Palmer et al., 1987). The four acid crater lakes, Kawah Kamojang (J9), Kawah Domas (J72), Kawah Putih (J51) and Kawah Sikidang (J59) had low pH values ranging from 1 to 2.9 and thus were favorable for B isotope fractionation. The B/Cl ratios indicate that, except for Kawah Putih, the three other crater lakes were formed due to condensation of the geothermal vapor phase (Fig. 4.3). The δ11B of vapor phase is generally enriched by up to 4 ‰ relative to the remaining liquid phase (Kanzaki et al., 1979; Leeman et al., 1992; Nomura et al., 1982; Spivack et al., 1990; Yuan et al., 2014). The difference in δ11B between the Dieng geothermal brines and the Kawah Kreta vapor phase was approximately 3.8 ‰ confirming condensation as the main process of fractionation. Assuming steam condensation from a similar vapor phase as Kawah Kreta, the Kawah Domas, Kawah Kamojang and Kawah Sikidang crater lakes were enriched in δ11B by 1.7 ‰, 5.5 ‰ and 31.1 ‰, respectively. The Kawah Kreta steam vent had a B concentration of 3.5 mg/L, which was slightly higher than Kawah Kamojang with 1.4 mg/L and Kawah Domas with 2.7 mg/L. Smith et al. (1987) reported for the Geyser geothermal field (USA) that the water of the vapor trap was approximately 50 times enriched in B concentration compared to the steam phase. Accordingly, an acid crater lake originated from vapor phase condensation is expected to have a higher B concentration than the vapor phase. Therefore, the low B concentration of the Kawah Kamojang and Kawah Domas crater lakes in comparison to the vapor phase should indicate additional processes which lowered the B concentration. This could have been caused either by precipitation of a B-rich mineral phase or adsorption by clay minerals. Both processes reduce the B concentration of the remaining water and thus enrich 11B due to adsorption of 10B by the solid phases (Agyei and McMullen, 1968; McMullen et al., 1961; Oi et al., 1989; Palmer et al., 1987; Schwarcz et al., 1969; Spivack and Edmond, 1987; Swihart et al., 1986; Vengosh et al., 1991b; Vengosh et al., 1992). Palmer and Sturchio (1990) reported more significant B isotope fractionation at low temperature and thus the high temperature of Kawah Domas (85 °C) should allow only little B isotope fractionation, causing the lighter δ11B values compared to Kawah Kamojang, where the temperature was 40 °C. Meanwhile, the heavy δ11B value in

60

BORON ISOTOPE VARIATIONS IN GEOTHERMAL SYSTEMS ON JAVA, INDONESIA

Kawah Sikidang, which was by 31 ‰ heavier than the vapor phase, cannot be explained exclusively by steam phase separation. Other mechanisms must have produced such a B-rich and heavy δ11B signature in this acid sulfate crater lake. Apart from the heavy δ11B, the Kawah Sikidang crater lake had a noticeable B enrichment of 73.3 mg/L. This value is higher than that of the other acid sulfate crater lakes and only comparable to the acid chloride crater lake Kawah Putih, which had a B concentration of 94.4 mg/L. However, Kawah Putih was Cl-rich (8084 mg/L) and its δ11B was +0.6 ‰, whereas Kawah Sikidang in contrast was Cl-poor (14.4 mg/L) with a δ11B of +34.9 ‰. Delmelle and Bernard (1994) explained that the chemical composition of an acid chloride crater lake is the result of condensation and oxidation of magmatic gases, such as, SO2, H2S and HCl, upon contact with oxygenated groundwater followed by water-rock interaction. Fumaroles in Japan (Kanzaki et al., 1979; Nomura et al., 1982) and Vulcano Island, Italy (Leeman et al., 2005), for example, had light δ11B, representing the magmatic fluids. Accordingly, the vapor phase of Kawah Putih should have been derived from magmatic fluids and thus inherited the light δ11B of an andesitic island arc magma of -2.3 to +3.5 ‰ (Palmer, 1991). After discharge at the surface, B isotope enrichment was moreless absent though the water was very acidic (pH = ~1) and relatively cold (T = 32.9 °C), probably was caused by continuous magmatic fluid input. The crater lake had a dimension of approximately 350 x 290 m2 (Sriwana et al., 2000) and the ambient air temperature was 16 °C at the time of sampling, that the fluid was cooled down rapidly. The main mechanism that produces heavy δ11B (+34.9 ‰) of Kawah Sikidang probably was evaporation. Vengosh et al. (1992) reported that during the latest stage of evaporation the δ11B of seawater could be enriched up to 30 ‰ due to incorporation of 10B into evaporate minerals, e.g., halite. Kawah Sikidang had a temperature of 87 °C without outflow to the nearby river, hence excessive evaporation could be expected. The condition is contrary to the Kawah Domas acid sulfate crater lake which had a temperature of 85 °C but the water drained to the adjacent river and thus the effect of evaporation was less significant than at Kawah Sikidang. Evaporation should have lowered the B concentration of Kawah Sikidang due to incorporation of B into evaporites. Therefore, the elevated B content of 73.3 mg/L of the crater lake has to be sustained by constant addition of subsurface B. The reservoir vapor phase probably had an initial B content comparable to the Kawah Kreta steam vent of 3.5 mg/L, however, in the shallow depth the vapor was B enriched due to interaction with

61

BORON ISOTOPE VARIATIONS IN GEOTHERMAL SYSTEMS ON JAVA, INDONESIA

B-rich minerals. This assumption was supported by the elevated B concentration of altered ground surrounding Kawah Sikidang which was up to 25.5 mg/Kg, compared to the less than 0.02 mg/Kg of unaltered ground in Dieng.

IV.4.3. Processes affecting the δ11B value of thermal waters Mixing with groundwater in the shallow subsurface prior to discharge leads to a heavier δ11B signature, because groundwater has generally a heavier δ11B signature than thermal water (Palmer and Sturchio, 1990; Vengosh et al., 1994b). A binary mixing model can be applied to estimate the effect of groundwater mixing for the B concentration and δ11B signature of thermal waters (Vengosh et al., 1991b; Yuan et al., 2014). To set up the model, the groundwater of Cangar (J15) and the geothermal brine from the Dieng geothermal field (J70) were used as end members. Groundwater dilution was evident for eight thermal waters, i.e., J10 (Kawah Hujan), J30(Ngunut), J36 (Sari Ater), J39 (Cibolang), J42 (Pakenjeng), J44 (Cilayu), J60 (Kawah Sileri) and J64 (Bitingan) (Fig. 4.5). However, most of the thermal waters plot above the mixing line and four thermal waters, J4 (Ciawi), J38 (Batu Kapur), J49 (Cisolok) and J54 (Cileungsing), plot below the mixing line (Fig. 4.5), indicating that the B concentration and δ11B signature in those waters is more complex than just groundwater mixing. Those samples plotting above ther mixing line are relatively enriched, which could have been caused by adsorption, phase separation, water-sediment interaction or any combination of these processes (Hemming and Hanson, 1992; Lemarchand et al., 2007; Palmer et al., 1987; Schwarcz et al., 1969; Spivack and Edmond, 1987; Vengosh et al., 1991a; Vengosh et al., 1991b; Vengosh et al., 1992; Xiao et al., 2013; Yuan et al., 2014). The four thermal waters plotting below the mixing line either had initially a light δ11B signature or desorbed B from solid phases. Although J36, J10, J39 and J64 plotted close to the groundwater mixing line (Fig. 4.5), groundwater mixing cannot be the only process responsible for their final δ11B composition. Based on their B/Cl ratios, J36 was relatively B depleted, while J10, J39 and J64 were relatively B enriched (Fig. 4.3). The B concentration in J36 was lowered due to adsorption onto clay minerals, which at the same time increased the δ11B value to +2.4 ‰. Since J36 was of the acid chloride type with a pH of 2, it likely had an initial δ11B value similar to the geothermal brines of approximately 0 ‰. In contrast, the higher B/Cl ratios of J10, J39 and J64 compared to the andesitic rock were a result of steam phase separation (Fig. 4.3). Thermal waters generated by condensation of a vapor phase in

62

BORON ISOTOPE VARIATIONS IN GEOTHERMAL SYSTEMS ON JAVA, INDONESIA

Fig. 4.5. δ11B vs. B diagram of thermal waters and geothermal brines with a theoretical binary mixing line of groundwater (J15) and geothermal brine (J70). Groundwater dilution is dominant for J10, J30, J36, J39, J42, J44, J60 and J64 thermal waters.

a geothermal system generally undergo δ11B enrichment of up to 4 ‰ (Leeman et al., 1992; Spivack et al., 1990). J10 and J64 had a δ11B of +2.3 ‰ and +3.2 ‰, respectively, which were close to the Kawah Kreta steam vent of +3.8 ‰. The slightly lower δ11B of those two thermal waters compared to the steam vent could be due to addition of 10B from carbonate mineral dissolution, because J10 and J64 were under saturated with respect to carbonate minerals (Table 4.3). In the case of J64 desorption of B from iron oxide could have been possible as well, since iron oxides were also undersaturated (Table 4.3). In comparison, J39 was saturated with respect to goethite, carbonate and clay minerals and its temperature of 68.9 °C would dampen B isotope fractionation after discharge. Therefore, the light δ11B of J39 must be derived from its vapor phase leading to the conclusion that B isotope fractionation

63

BORON ISOTOPE VARIATIONS IN GEOTHERMAL SYSTEMS ON JAVA, INDONESIA

Table 4.3. Saturation indices with respect to carbonate, clay and goethite minerals were calculated using PHREEQC (Parkhurst and Appelo, 1999).

Sample Location Saturation Index ID Calcite Dolomite Strontianite Illite Smectite Goethite J2 Baturaden (Slamet volc.) 0.64 2.74 -0.20 - - 1.56 J4 Ciawi 0.33 2.12 -0.61 - - 1.81 J7 Cipanas -0.81 -0.02 - - - 2.04 J10 Kawah Hujan (Kamojang) -4.27 -7.78 -5.49 2.12 -2.54 1.40 J13 Padusan (Arjuna-Welirang) 0.22 1.81 -0.71 - - 1.50 J18 Songgoriti (Arjuna-Welirang) 0.03 1.45 -0.31 6.96 3.48 0.62 J21 Segaran (Lamongan volc.) 0.15 2.13 -0.33 - - -1.31 J24 Banyuasin (Lawu volc.) -0.28 0.38 -0.08 - - 1.16 J28 Kondo (Lawu volc.) -0.25 0.51 0.37 - - -1.03 J30 Ngunut (Lawu volc.) -0.27 0.33 -0.08 - - 0.77 J34 Gucci (Slamet volc.) 0.73 3.06 -0.21 4.03 4.48 4.79 J74 Krakal, Kebumen 1.14 - 0.39 - - - J38 Batu Kapur -0.14 1.19 -0.26 - - 0.29 J39 Cibolang (Pangalengan) 0.35 1.89 -0.69 1.40 2.29 4.74 J42 Pakenjeng 0.09 - -0.65 - - 3.82 J44 Cilayu 1.32 3.42 0.92 -3.13 1.81 7.01 J47 Kalianget -0.12 1.31 -1.15 - - 0.81 J48 Cikundul -0.14 -0.36 - 3.69 2.02 -0.01 J49 Cisolok 1.18 2.62 0.42 - - 4.49 J54 Cileungsing (Tampomas) 0.54 2.39 0.51 - - 2.01 J58 Parangtritis 0.76 0.64 0.01 - - 4.00 J60 Kawah Sileri (Dieng) -0.80 -0.60 -1.75 3.02 -0.14 -4.36 J61 Pulosari (Dieng) 0.08 1.20 -0.64 0.50 0.80 3.78 J64 Bitingan (Dieng) -1.28 -1.61 -2.07 4.03 -0.98 -4.27

during vapor phase separation must have been insignificant. The isotope fractionation of B is considered unimportant at reservoir temperatures above 400 °C (Leeman et al., 1992; Spivack et al., 1990) or high reservoir pressure (Liebscher et al., 2005). The reservoir temperature of J39 ranged from 221 to 323 °C, similar to J10 and J64 of approximately ranging from 220 to 350 °C (Purnomo and Pichler, 2014; Sudarman et al., 1995), hence temperature cannot be responsible for the light δ11B of the J39 vapor phase. Meanwhile, the reservoir pressure of J39 (Wayang-Windu geothermal field) was recorded up to 84 bar (Mulyadi and Ashat, 2011), higher than J10 (Kamojang geothermal field) and J64 (Dieng geothermal field) of up to 36 and 40 bar, respectively

64

BORON ISOTOPE VARIATIONS IN GEOTHERMAL SYSTEMS ON JAVA, INDONESIA

(Bachrun et al., 1995; Zuhro, 2004). Therefore, the light δ11B of the J39 vapor phase was more likely caused by high reservoir pressure. Twelve thermal waters, J2, J7, J13, J18, J21, J24, J28, J34, J47, J48, J55 and J61, were δ11B enriched, thus plotted above the mixing line (Fig. 4.5). The increasing δ11B in a hot spring where adsorption/coprecipitation of B occurs should coincide with a decrease of the B/Cl ratio. However, as seen in the Cl vs. B diagram only J2 underwent B depletion, while J55 plotted in the steam separation group and J7, J13, J18, J21, J24, J28, J34, J48 as well as J61 plotted close to andesitic rock leaching (Fig. 4.3). Compared to the Kawah Kreta steam vent, J55 was δ11B enriched by 9 ‰, a result of B adsorption onto clay minerals due to its acidic condition (pH = 2.8). This process lowered the B/Cl ratio of J55, however, the thermal water still plotted in the steam separation group due to its initial higher B/Cl ratio, formed by the high temperature geothermal system. Meanwhile, J7, J13, J18, J21, J24, J28, J34, J48 and J61 plot close to the andesitic line, which might be caused by Cl- depletion during ascent or initially higher B/Cl ratios. In a mud volcano, for example, Cl- depletion was considered due to NaCl filtration by clay minerals during clay dehydration (Dahlmann and Lange, 2003; Dia et al., 1999; Hensen et al., 2004; Kastner et al., 1991; You et al., 2004). This mechanism, however, is unlikely in a geothermal system, clay minerals along the fluid pathway should be hydrated instead. Vapor phase separation and reaction with B-rich minerals potentially increase the B/Cl ratio of a thermal water. However, the former could be ruled out because those nine thermal waters had higher Cl- contents compared to the thermal waters generated by steam separation. Therefore, it could be suggested that they underwent B enrichment due to interaction with B-rich minerals in the subsurface. The minerals supplied B into the ascending thermal water and thus increased the B/Cl ratio. Post discharge at the surface, the water underwent B adsorption/coprecipitation that reduced the B/Cl ratio but increased the δ11B. In order to identify the mineral phases that potentially adsorb/desorp B in hot springs, the saturation indices of carbonate, clay and iron oxide minerals were calculated using PHREEQC of Parkhurst and Appelo (1999) (Table 4.3). Lemarchand et al. (2007) reported a higher B isotope fractionation during adsorption onto iron oxide than onto clay and carbonate minerals. The saturation index calculation indicates that B coprecipitation with goethite could be expected for most of the hot springs, but not for J21, J28, J74, J48, J60 and J64 (Table 4.3). In hot springs J21

65

BORON ISOTOPE VARIATIONS IN GEOTHERMAL SYSTEMS ON JAVA, INDONESIA and J74 B likely coprecipitated with calcite, in J48, J60 and J64 B adsorbed onto clay minerals and in J28 it coprecipitated with dolomite and strontianite.

IV.4.4. The δ11B of fault-hosted and volcano-hosted geothermal systems The magmatic fluid input in volcano-hosted geothermal systems, which is absent/minor in fault-hosted geothermal systems, distinguishes their chemical characteristics (Purnomo and Pichler, 2014). The presences of magmatic gases (CO2,

H2S, HCl, SO2) in volcano-hosted geothermal systems produces reactive thermal waters, favorable for clay, carbonate and sulfate mineral precipitation. These minerals then can preferentially adsorb 10B over 11B causing the δ11B of thermal waters to increase. In addition, the longer flow path and thus ascent time of volcano-hosted thermal waters compared to fault-hosted geothermal systems enhances B isotope fractionation. Considering these two factors, volcano-hosted thermal waters should have a heavier δ11B signature than fault-hosted thermal waters. By excluding thermal waters influenced by seawater and acid crate lakes, this assumption could be confirmed. With the exception of J48 (Cikundul), the fault-hosted thermal waters had δ11B values ranging from -0.7 to +0.2 ‰, while volcano-hosted thermal waters were generally heavier and ranged from -2.4 to +12.8 ‰. Figure 4.6 indicates that most of the volcano-hosted thermal waters were δ11B enriched, while for the fault-hosted such indication was only found in J48. Five volcano-hosted thermal waters, J4, J38, J39, 54 and J60, which had light δ11B signatures were saturated with carbonate minerals and goethite (Table 4.3). The addition of 10B to those thermal water due to mineral dissolution can be dismissed and thus sources of those thermal waters should have had a lighter δ11B signature than the geothermal reservoir at Dieng. Meanwhile, the relatively heavy δ11B (+9.3 ‰) and high B value (10.7 mg/L) of the fault-hosted hot spring J48 are considered a result of water-sediment interaction, since this hot spring was located at the bottom of the Cimandiri river. The relatively high B value and a temperature of 51 °C indicate insignificant mixing with river water prior to sampling.

IV.5. Conclusions Hot springs and hot crater lakes on Java had a range of δ11B from -2.4 to +34.9 ‰, relatively similar to other geothermal systems in the World (Barth, 1993). Seawater input was detected for two fault-hosted geothermal systems, Parangtritis and Krakal, and is considered a source of the heavy δ11B. The Kawah Sikidang acid sulfate crater

66

BORON ISOTOPE VARIATIONS IN GEOTHERMAL SYSTEMS ON JAVA, INDONESIA

Fig. 4.6. δ11B vs. B diagram indicates most volcanic-hosted thermal waters were δ11B enriched, while δ11B enrichment for fault-hosted thermal waters were moreless absent.

lake had a δ11B indicating seawater addition to its geothermal source, although the low Cl/B ratio excluded seawater as a source for 11B. Two types of hot acid crater lakes, Cl-poor (Cl- < 15 mg/L) and Cl-rich (Cl- = 8084 mg/L), showed a contrasting δ11B composition. The Cl-poor crater lakes had relatively heavy δ11B values ranging from +5.5 to 34.9 ‰, while the Cl-rich lake had a lighter δ11B of +0.6 ‰, similar to the Dieng geothermal brines. The heavier δ11B values of the acid sulfate crater lakes are a combination of vapor phase separation, evaporation after discharge and preferential adsorption of 10B by clay minerals. The heavy δ11B (+34.9 ‰) of the Kawah Sikidang acid sulfate crater lake was mainly caused by evaporation, while its B content of 73.3 mg/L was due to water-rock interaction with B-rich minerals in the shallow subsurface. Meanwhile, the light δ11B of

67

BORON ISOTOPE VARIATIONS IN GEOTHERMAL SYSTEMS ON JAVA, INDONESIA the Kawah Putih acid chloride crater lake was a result of continuous magmatic gases input. The initial division of the geothermal systems on Java into fault-hosted and volcano-hosted (Purnomo and Pichler, 2014) was confirmed by their different δ11B signatures. In general, fault-hosted systems had lighter δ11B values ranging from -0.7 to +0.2 ‰, while the volcano-hosted systems had heavier δ11B values of above 1 ‰. The comparably faster ascent to the surface and the absence of magmatic fluids in fault-hosted geothermal systems should not facilitate B isotope fractionation, hence the lower δ11B values. In volcano-hosted geothermal systems, the ascent time is relatively longer and the condensation of magmatic gas, as well as the formation of clay, carbonate, sulfate and iron oxide minerals should facilitate B isotope fractionation, hence the higher δ11B values.

Acknowledgments A part of this study was funded by the Ministry of Energy and Mineral Resources of Indonesia through PhD scholarship grants number: 2579 K/69/MEM/2010 for B.J. Purnomo. Thanks to Laura Knigge for the laboratory assistance and to Britta Hinz- Stolle for an editorial review.

68

GEOTHERMAL SYSTEMS ON THE ISLAND OF BALI, INDONESIA

V. Geothermal systems on the island of Bali, Indonesia

Modified from:

Budi Joko Purnomo and Thomas Pichler *

* Geochemistry & Hydrogeology, Department of Geosciences, University of Bremen, Germany

Submitted to:

Journal of Volcanology and Geothermal Research

69

GEOTHERMAL SYSTEMS ON THE ISLAND OF BALI, INDONESIA

Abstract

This paper presents an overview of geothermal systems on the island of Bali, Indonesia by presenting physicochemical data for shallow thermal wells, hot springs, cold springs and volcanic lakes. A total of 4 locations were sampled and classified based on their position in either as a volcano-hosted or a fault-hosted geothermal system. A carbonate host rock for the geothermal reservoirs could not be confirmed, because the characteristic δ18O shift to the right for carbonate reservoir fluids was not - 2+ 2+ 2+ observed. The HCO3 of the thermal waters was well correlated with Ca , Mg , Sr and K+ indicating water-rock interaction in the presence of carbonic acid. The (Ca2+ + 2+ - Mg )/HCO3 molar ratios were approximately 0.4 and K/Mg ratios were typical for interaction with calc-alkaline magmatic rocks. The B/Cl ratios indicated steam phase separation for the Bedugul and Banjar geothermal systems. The heavy δ11B of +22.5 ‰ and a Cl/B ratio of 820 confirmed seawater input into the Banyuwedang geothermal system. Comparison of the Si, Na/K, Na/K/Ca and Na/Li geothermometers with actual reservoir temperature measurements and physicochemical considerations led to the conclusion that the Na/Li thermometer provided the best results for geothermal systems on Bali. Using this thermometer, the following reservoir temperatures were calculated: (1) Penebel (Bedugul) from 235 to 254 °C, (2) Batur 240 °C, (3) Banjar 255 °C and (4) Banyuwedang below 100 °C. The 2H and 18O isotope compositions of the geothermal fluids indicated meteoric water as a source of all thermal waters studied.

Keywords: Bali, carbonate, volcanic, host-rock, seawater input, geothermal system, geothermometer, 2H and 18O isotope

70

GEOTHERMAL SYSTEMS ON THE ISLAND OF BALI, INDONESIA

V.1. Introduction The island of Bali, Indonesia is host to several hot springs and geothermal systems, of which some are of interest for geothermal exploitation. The Bedugul geothermal field, located near Lake Bratan, was identified to cover an area of approximately 8 km2 with an estimated annual electric energy potential of 80 MWe for 30 years (Hochstein et al., 2005; Hochstein and Sudarman, 2008; Mulyadi et al., 2005). However, the development was suspended due to environmental and cultural concerns. In addition to Bedugul, other geothermal prospects on Bali are the Batur, Banyuwedang and Banjar geothermal systems. Bali is dominantly covered by volcanic rocks, overlying the Tertiary carbonate rocks that outcrop in the southern and western part of the island (Hadiwidjojo et al., 1998). The Bedugul geothermal field was reported producing brines dominated by

CO2 of approximately 97 wt. % and the reservoir was assumed to be hosted in carbonate rocks (Mulyadi et al., 2005). To the contrary Geiger (2014) concluded that the reservoir was in volcanic rocks and that the carbonate basement was leached by shallow magma, 2 to 7 km depth, in the Batur and Agung volcanoes based on the thermobarometric study. The dissolution of carbonate rock by magma releases large amounts of CO2 gas due to the breakdown of CaCO3 into CO2 gas and CaO (Allard, 1983; Chadwick et al., 2007; Deegan et al., 2010; Gertisser and Keller, 2003;

Marziano et al., 2007; Marziano et al., 2009). The rich in CO2 volatile magma subsequently ascends to the geothermal reservoir and promotes vaporization, which in turn produces a CO2-rich vapor phase (Lowenstern, 2001). It is possible that geothermal systems on Bali could be hosted by carbonate rocks, but CO2 content alone would be insufficient for that conclusion. The carbonate rocks were determined as the reservoir rock in some volcanic areas in Italy, e.g., Vicano-Cimino and Sabatini- Tolfa (Cinti et al., 2011; Cinti et al., 2014). There the 2+ 2+ - thermal is characterized by a (Ca + Mg )/HCO3 molar ratio of ~1 as a result of calcite and/or dolomite dissolution (Capaccioni et al., 2011; Cinti et al., 2011; Cinti et al., 2014; Gemici and Filiz, 2001; Goff et al., 1981; Grassa et al., 2006; Levet et al., 2002; Pasvanoglu and Chandrasekharam, 2011; Pasvanoglu and Gultekin, 2012). Another characteristic of thermal waters hosted by carbonate rocks could be the relative 18O isotope enrichment due to the heavier δ18O of the carbonate host rock, as indicated at Balikesir and Cesme, Turkey; Lanzarote, Spain; and the Salton Sea, USA (Arana and Panichi, 1974; Craig, 1966; Gemici and Filiz, 2001). Although some carbonate rock associated geothermal systems also showed depletion of 18O isotope

71

GEOTHERMAL SYSTEMS ON THE ISLAND OF BALI, INDONESIA due to exchange with CO2, for instances at Vicano-Cimino, Sabatini, and Sicily, Italy (Cinti et al., 2011; Cinti et al., 2014; Grassa et al., 2006). This paper presents new physicochemical and isotope (18O, 2H and 11B) data for hot springs and shallow thermal wells on Bali with the objective to investigate the host rocks of the geothermal systems. Additionally, boron isotopes were applied to identify seawater input and solute geothermometers to predict the reservoir temperatures.

V.2. Geological setting Bali is a part of the Sunda-Banda volcanic islands arc, which extends for approximately 4700 km east to west, from the island of Damar to the island of Sumatera. The arc is caused by the convergence of the Indo-Australian and Eurasia plates, with a rate of c.a. 6 to 7 cm/a (Hamilton, 1979; Simandjuntak and Barber, 1996). This process caused volcanisms on Bali since the late Tertiary (Hadiwidjojo et al., 1998; Hamilton, 1979; Van Bemellen, 1949) and produced a vast distribution of volcanic rocks. The Jembrana volcanic complex occupies the western part of the island, the Buyat-Bratan-Batur volcanic complex the central part, and the Agung and Seraya volcanic complexes the eastern parts. Underlying the volcanic rocks are sedimentary rocks of Tertiary age, which are minimally exposed in the east, south and west part of the island (Fig. 5.1) (Hadiwidjojo et al., 1998). The volcanic rocks on Bali are calc-alkaline characterized by a medium concentrations of K (Nicholls and Whitford, 1983; Whitford et al., 1979). Based on the thermobarometric results of clinopyroxene and plagioclase at Batur volcano, Geiger (2014) suggested the existence of a shallow magma in 2 to 4 km depth, associated with a carbonate sedimentary crust. This is corroborated by InSar satellite data indicating shallow magma at 2 to 4 km depth (Chaussard and Amelung, 2012) and by earthquake focal zones at 1.5 to 5 km depth (Hidayati and Sulaeman, 2013). Volcanoes where carbonate assimilation takes place generally have an explosive eruption behavior (Deegan et al., 2010). Explosive eruptions formed two large calderas on Bali, the Batur and the Bratan calderas (Reubi and Nicholls, 2004; Watanabe et al., 2010; Wheller and Varne, 1986). Both calderas are geothermal prospects, due to the presence of thermal water in the shallow subsurface. Although a geothermal reservoir was confirmed beneath the Bratan lake, (Mulyadi et al., 2005), surface features of geothermal systems, such as hot springs are virtually absent in the Bratan caldera. Outside of the caldera to the south, however, several hot springs are present in the Penebel area (Fig. 5.1). On the

72

GEOTHERMAL SYSTEMS ON THE ISLAND OF BALI, INDONESIA northwestern side of the Buyat-Bratan volcano, the Banjar hot spring discharges at the contact between the Buyat-Bratan-Batur volcanic complex and the Tertiary Asah Formation. At the western end of the island, the Banyuwedang hot spring is located in carbonate rocks of the Prapatagung Formation.

Fig. 5.1. Geological map of Bali showing the sampling locations (modified from Hadiwidjojo et al., 1998).

73

GEOTHERMAL SYSTEMS ON THE ISLAND OF BALI, INDONESIA

V.3. Results The physicochemical and stable isotopes (18O, 2H and 11B) data of the water samples from Bali are presented in Table 5.1. The temperatures of thermal waters ranged from 37.2 to 45.2 °C, while the selected cold waters ranged from 24 to 30.1 °C. The thermal waters had slightly acid to neutral pH, while cold springs neutral and lake waters were slightly alkaline. The thermal and cold waters had relatively similar ranges of TDS, Cl-, Na+ and K+ concentrations. Meanwhile the thermal waters had 2+ 2+ - wider ranges of Ca , Mg and HCO3 contents compared to the cold waters. The Ca2+ concentration of the thermal waters ranged from 51.3 to 211.5 mg/L, Mg2+ from - 2+ 51.5 to 243.8 mg/L and HCO3 from 31.7 to 2235 mg/L, while for cold waters the Ca 2+ - and Mg were lower than 100 mg/L and HCO3 ranged from 19.5 to 761.3 mg/L. The Cl- content of thermal waters ranged from 17.3 to 902.1 mg/L and for cold waters from under detection limit to 1025.7 mg/L. The Cl-rich cold waters were found in B12 (Batur Lake) and B8 (Pejarakan), with Cl- contents of 188.6 and 1025.7 mg/L, respectively. 2+ 2+ - Thermal waters of B1, B2, B3, and B4 had Ca , Mg and HCO3 concentrations a magnitude higher than the other thermal waters. However, B4 differed from B1, B2 2- - 2- and B3 due to its higher SO4 and lower Cl concentrations. B4 had a SO4 content of - 2- 111.7 mg/L and a Cl content of 61.2 mg/L, compared to SO4 of below detection limit and Cl- ranged from 377 to 443.9 mg/L in B1, B2 and B3. Thermal waters of B6 and B13 had TDS of <1000 mg/L and Cl- < 20 mg/L, lower than the other thermal waters, which varied between 1430 to 2600 mg/L and from 61.2 to 902.1 mg/L, respectively. The two thermal waters probably were more diluted by shallow groundwater. The thermal water located near the coastline, B7, had the highest TDS and Cl- contents - and the lowest HCO3 concentration. The thermal waters had δ2H and δ18O values ranging from -42.4 to -33.2 ‰ and from -6.8 to -5.6 ‰, respectively. This was probably a result of the range in elevation from close to 0 (seawater) to approximately 1200 m above sea level. The δ2H of cold springs and a shallow well ranged from 36.9 to -30.0 ‰ and δ18O ranged from -6.0 to - 5.4 ‰. In contrast, cold waters from two freshwater lakes, Batur and Bratan, had heavier δ2H, ranging from -16.4 to -14.6 ‰, and δ18O, from –2.3 to -1.7 ‰. The selected thermal waters had a wide range of δ11B compositions, from +1.3 to +22.5 ‰. In accordance to the TDS and Cl- contents, the heaviest δ11B value was found in sample B7, might be due to seawater input.

74

GEOTHERMAL SYSTEMS ON THE ISLAND OF BALI, INDONESIA

V.4. Discussion V.4.1. Geochemistry of thermal waters A geothermal system generally produces three types of hot springs, neutral chloride, acid sulfate and bicarbonate waters, but mixtures between the individual groups are common (Hedenquist, 1990; Hochstein and Browne, 2000; Nicholson, 1993; White, 1957). The discharge composition of thermal springs is controlled by two sets of processes: 1) deep reservoir conditions (deep reservoir = reaction zone immediately above the heat source), and 2) secondary processes during ascent. In the deep reservoir, host rock composition, temperature, direct magmatic contributions and residence time are the controlling factors. During ascent a drop in pressure and temperature can initiate phase separation and mineral precipitation, causing a dramatic change in fluid composition. Mixing with other hydrothermal fluids and/or groundwater is possible at any depth. In near-shore and submarine environments mixing with seawater cannot be ruled out. The chemical composition of a hydrothermal fluid, sampled at the surface, generally contains an imprint of its subsurface history and chemically inert constituents (tracers) provide information about their source, whereas chemically reactive species (geoindicators) record physico-chemical changes (Ellis and Mahon, 1977; Giggenbach, 1991; Nicholson,

1993). Classification of thermal waters on Bali using the Cl-SO4-HCO3 ternary diagram (Chang, 1984; Giggenbach, 1991; Giggenbach, 1997) indicated a - bicarbonate (HCO3 ) type for B1, B2, B3, B4, B6 and B13, a mixing type for B9, B10 and B11, and a neutral chloride (Cl-) type for B7 (Fig. 5.2). Neutral chloride waters are usually thought to represent the deep reservoir fluid, while acid sulfate and bicarbonate waters form by underground absorption of vapors separated from a neutral chloride water into cooler ground water. Whether acid sulfate or bicarbonate waters are formed depends on the gas content of the vapor and redox conditions in the shallow subsurface (Ellis and Mahon, 1977; Giggenbach, 1997; Hedenquist, 1990; Henley and Ellis, 1983).

75

2 18 11 Table 5.1. Sampling locations, physicochemical and stable isotopes ( H, O and B) data of thermal and cold waters on Bali. The ρCO2 was calculated using the computer code PHREEQC (Parkhurst and Appelo, 1999).

18 2 11 ID Location T pH ORP Cond. TDS Ca Mg Na K Cl HCO3 SO4 Si Al Li B Sr Fe δ O δ H δ B Logog

°C mV uS/cm mg/l ‰ ρCCOO2 Hot springs B1 Yeh Panas 1 38.8 6.5 25 3016 2238 122.4 161.6 263.4 50.1 377.0 1466.4

GEOTHERMAL SYSTEMS ON THE ISLAND OF BALI, INDONESIA

Fig. 5.2. Cl-SO4-HCO3 ternary diagram (Giggenbach diagram). Most of the thermal waters were of the bicarbonate type. The samples from Batur (B9, B10 and B11) were a mixing type and B7 a chloride type. The plots of the Pejarakan cold well (B8) and Lake Batur (B12) as mixing water type probably were caused by seawater and thermal water inputs.

The TDS value of 1525 mg/L of B12 was relatively similar to those of the Batur thermal waters (B9, B10 and B11), probably due to major discharge of thermal water into the lake. However, it should be noted that B12 was sampled from a site relatively close to of Batur thermal waters, hence considering the large dimension of the lake of approximately 6.6 km length and 2.5 km width, the sample probably does not represent the general chemistry of the lake water. Meanwhile, the higher Cl- concentration of the shallow well (B8) was likely caused by seawater input due to its location close to sea.

77

GEOTHERMAL SYSTEMS ON THE ISLAND OF BALI, INDONESIA

Solely based on their geological setting geothermal systems on Bali can be divided into two groups, volcano-hosted are Batur (B9, B10 and B11) and Penebel (B1, B2, B3, B4 and B6), while Banyuwedang (B7) and Banjar (B13) are fault-hosted. Following the interpretation of Purnomo and Pichler (2014) only Banyuwedang (B7) was considered a truly fault-hosted geothermal system. This thermal water differs - from the others due to its poor of HCO3 content and thus plots as group C, fault- hosted geothermal system in Figure 5.3. The volcanic-hosted thermal waters are divided into group A, diluted thermal waters, and group B, originated from the margin of the ‘primary neutralization’ zone of Giggenbach (1988). In this zone, the elevated - HCO3 content is produced by separation of CO2 and the subsequent reaction with groundwater. The extent of groundwater mixing governs the Cl- concentration of - 2 thermal waters. The HCO3 content of thermal waters was well correlated (R ~ 0.9) with the Ca2+, Mg2+,Sr2+ and K+ (Fig. 5.4), hence indicated rock dissolution by carbonic acid. The rocks dissolution was triggered by the formation of weak acid, H2CO3, at temperatures below 300 °C due to oxidation of CO2 by groundwater (Bischoff and 2+ 2+ - Rosenbauer, 1996; Giggenbach, 1997; Lowenstern, 2001). The (Ca + Mg )/HCO3 molar ratios of the thermal waters of approximately 0.4 (Fig. 5.5) are lower than the ratios for thermal waters hosted by carbonate rocks of 1, as reported for the geothermal systems of Vicano-Cimino, Sabatini-Tolfa and Sicily, Italy; Cesme, Nevsehir and Terme-Karakurt, Turkey, Bagneres de Bigorre, France; and Jemez springs, USA (Capaccioni et al., 2011; Cinti et al., 2011; Cinti et al., 2014; Gemici and Filiz, 2001; Goff et al., 1981; Grassa et al., 2006; Levet et al., 2002; Pasvanoglu and Chandrasekharam, 2011; Pasvanoglu and Gultekin, 2012). Following this interpretation a carbonate type host rock for the geothermal systems on Bali could be ruled out. The thermal waters are more likely hosted by magmatic rocks of calc- alkaline origin as indicated in the K/Mg vs. Na/K diagram (Fig. 5.6). This type of magma is generally produced by volcanoes located in the middle of a volcanic island arc, in between tholeiitic and high-K calc-alkaline magmatic regions (Whitford et al., 1979).

78

GEOTHERMAL SYSTEMS ON THE ISLAND OF BALI, INDONESIA

Fig. 5.3. HCO3 vs. Cl diagram. Groups A and B were volcano-hosted thermal waters originated from the ‘primary neutralization’ zone, where group B was less diluted by groundwater compared to group A. Group C was fault-hosted thermal water, hence underwent insignificant magmatic fluid input.

The final pH of the thermal water was governed by addition of CO2, hence

H2CO3 formation, and subsequent water-rock interaction. While addition of CO2 lowers the pH, water-rock interaction causes an increase due to the comsuption of H+. The effect of CO2 addition in Penebel group (B1, B2, B3, B4 and B6) and Banjar (B13) dominated water-rock interaction and thus kept a slightly acid pH (Fig. 5.7a).

Meanwhile, the CO2 supply in Batur group (B9, B10 and B11) and thus decreasing pH was neutralized by water-rock interaction. However, the lower ρCO2 and higher pH of Batur group compared to the other thermal waters also could be caused by calcite precipitation, which also releases CO2, indicated by the saturation condition with respect to calcite (Fig. 5.7b).

79

GEOTHERMAL SYSTEMS ON THE ISLAND OF BALI, INDONESIA

Fig. 5.4. a) Ca vs. HCO3, b) Mg vs. HCO3, c) K vs. HCO3 and d) Sr vs. HCO3 - 2+ 2+ + 2+ diagrams. The well correlations of HCO3 with Ca , Mg , K and Sr indicated rock dissolution by carbonic acid.

80

GEOTHERMAL SYSTEMS ON THE ISLAND OF BALI, INDONESIA

2+ 2+ - Fig. 5.5. The (Ca + Mg )/HCO3 molar ratios of thermal waters were approximately 2+ 2+ - 0.4, an indication of non-carbonate host-rocks. For comparison (Ca + Mg )/HCO3 molar ratios of the carbonate-hosted thermal waters plotted using data from Cinti et al. (2014), Capaccioni et al. (2011) and Levet et al. (2002).

81

GEOTHERMAL SYSTEMS ON THE ISLAND OF BALI, INDONESIA

Fig. 5.6. K/Mg vs. Na/K diagram indicates dissolution of calc-alkaline volcanic rocks of Batur volcano. Data of calc-alkaline volcanic rocks of Batur are from Reubi and Nicholls (2004) and high-K volcanic rocks of Muria volcano from Whitford et al. (1979).

Fig. 5.7. a) log ρCO2 vs. pH and b) SI calcite vs. pH diagrams. Thermal waters with high ρCO2 had lower pH, a typical of CO2 fed thermal waters. The water pH probably was also buffered by calcite precipitation.

V.4.2. Phase separation and seawater input The Cl/B ratio of thermal waters can be used to identify water-rock interaction, steam separation and seawater input in the subsurface (Arnorsson and Andresdottir, 1995; Purnomo and Pichler, 2014; Valentino and Stanzione, 2003). The Cl vs. B diagram of thermal waters from Bali indicates reaction with andesitic rock for the Yeh Panas group (B1, B2 and B3) and the Batur group (B9, B10 and B11) of thermal waters. Either B depletion or seawater input is indicated for Banyuwedang (B7) and steam phase separation for Belulang (B4), Angseri (B6) and Banjar (B13) (Fig. 5.8). The last process elevates the B/Cl ratio of the thermal water by scavenging B into the steam phase, thus relatively increasing the Cl content of the residual water (Arnorsson and Andresdottir, 1995; Truesdell et al., 1989). That steam phase separation occurred for B4 and B6 was confirmed due to the existence of two phases, liquid and vapor, zone in the liquid-dominated reservoir of the Bedugul geothermal field (Hochstein et al., 2005).

82

GEOTHERMAL SYSTEMS ON THE ISLAND OF BALI, INDONESIA

The distinct δ11B composition of seawater of +39.61 ‰ (Foster et al., 2010) is heavier than other fluid sources and thus can be used to investigate seawater input or B adsorption onto minerals to explain the B/Cl ratio observed for Banyuwedang (B7). This method was successfully applied for some geothermal systems on Java island, Indonesia (Parangtritis and Krakal), Iceland (the Reykjanes and Svartsengi) and three areas in Japan (the Izu-Bonin arc, Kusatsu-Shirane area, and Kagoshima) (Aggarwal and Palmer, 1995; Aggarwal et al., 2000; Kakihana et al., 1987; Millot et al., 2009; Musashi et al., 1988; Oi et al., 1993; Purnomo et al., 2015). Adsorption of B by minerals also increases the δ11B of the thermal water due to the preferentially fractionation of 10B into the solid phase (Palmer et al., 1987; Schwarcz et al., 1969; Xiao et al., 2013). The magnitude, however, is lower than what can be observed due to seawater input. Banyuwedang (B7) had a δ11B composition of +22.5 ‰ and a Cl/B ratio of 806; therefore, it plots close to the mixing line between seawater and thermal water in the δ11B vs. B/Cl diagram (Fig. 5.9), which would confirm seawater input. The thermal water shift away from the seawater point in the B vs. Cl diagram was caused by groundwater mixing that lowered its B and Cl- concentrations.

83

GEOTHERMAL SYSTEMS ON THE ISLAND OF BALI, INDONESIA

Fig. 5.8. Cl vs. B diagram illustrates four processes in the sub surface, i.e., a steam phase separation for B4, B6 and B13; an andesitic rock leaching for B1 to B3 and B9 to B11; and either a B depletion or seawater input for B7.

Fig. 5.9. The plot of Banyuwedang close to the mixing line of seawater-thermal water in the δ11B vs. B/Cl diagram confirms seawater input.

V.4.3. Geothermometry The reservoir temperatures of geothermal systems on Bali were calculated using solute geothermometers, SiO2, Na-K-Ca, Na-K and Na-Li (Table 5.2) (Fournier, 1977; Fournier, 1979; Fournier and Truesdell, 1973; Kharaka and Mariner, 1989). These geothermometers record their temperature-dependent equilibrium reactions, hence application of multiple geothermometers can be used to evaluate secondary processes during thermal water ascent from the reservoir to the surface. These processes include shallow water dilution, conductive cooling, adiabatic cooling, mineral precipitation, water-rock interaction and re-equlibration (Fournier, 1977; Kaasalainen and Stefánsson, 2012). Such an evaluation has been successfully applied, for instances, on Java, Indonesia and Ambitle island, Papua New Guinea (Pichler et al., 1999; Purnomo and Pichler, 2014).

84

GEOTHERMAL SYSTEMS ON THE ISLAND OF BALI, INDONESIA

The silica geothermometer calculated lower reservoir temperatures compared to the Na-K, Na-K-Ca and Na-Li geothermometers, which is common for samples taken at the surface from hot springs, rather than directly from the geothermal reservoir. That geothermometer predicted reservoir temperatures ranging from 44 to 136 °C, while those predicted by the Na/K thermometer ranged from 257 to 773 °C, those predicted by the Na/K/Ca thermometer from 130 to 236 °C and those predicted by the Na/Li thermometer from 162 to 254 °C (Table 5.2). The calculation of silica geothermometry is based on absolute silica content, hence sensitive to boiling, precipitation and dilution (e.g. Nicholson, 1993). This deficiency can be overcome by calculating the silica parent using the silica mixing model of Fournier (1977). However, this method could not be used on Bali because the thermal waters in a given geothermal system had relatively similar temperatures and silica contents, for example the Penebel group (B1, B2, B3, B4 and B6). Meanwhile, the Na/K temperatures are likely overestimations caused by competition of Ca2+, Na+ and K+ during ion exchange (Nicholson, 1993). The use of the Na/Li geothermometer resulted in reservoir temperatures ranging from 235 to 254 °C for the Penebel thermal waters (B1, B2, B3, B4 and B6), which were relatively similar to actual reservoir temperatures at 1800 m below ground of the nearby Bedugul geothermal field of 243 °C (Mulyadi et al., 2005). This indicates the applicability of the Na/Li geothermometer as has been proposed by, e.g., Fouillac and Michard (1981). Based on this, the reservoir temperatures of the other geothermal systems on Bali, with an exception of Banyuwedang (B7), were predicted using Na/Li geothermometer. Therefore, the reservoir temperature of the Batur geothermal system was approximately 240 °C and Banjar was 255 °C. However, due to the input of seawater in Banyuwedang (B7), there a geothermometer based on Na+ content is unreliable. The silica geothermometer predicted a temperature of 44 °C for B7, similar to the discharge temperature and hence a likely underestimation. Therefore, the reservoir temperature of B7 could not be reliably calculated, but probably is lower than 100 °C, a temperature similar to most of the fault-hosted geothermal system on Java (Purnomo and Pichler, 2014). Calcite precipitation during thermal water ascent was predicted by comparing the result of Na/Li and Na/K/Ca geothemometers. Precipitation of calcite during thermal water ascent reduces the Ca2+ concentration and thus should result in lower calculated Na/K/Ca temperatures. Calculated temperatures ranged from 209 to 236 °C for the Penebel group of thermal waters and thus were similar to those calculated

85

GEOTHERMAL SYSTEMS ON THE ISLAND OF BALI, INDONESIA the Na/Li geothermometer, which would indicate insignificant calcite precipitation during fluid ascent. In contrast, the difference between 170 °C (Na/K/Ca) and 240 °C (Na/Li) for the Batur group of thermal waters indicates that calcite precipitated during ascent. This is corroborated by the lower ρCO2 and higher pH compared to the Penebel thermal waters (Fig. 5.7).

Table 5.2. Calculated reservoir temperatures using solute geothermometers. Location Sample Geothermometers (°C) ID Silica Na/K/Ca Na/K Na/Li (Truesdell, (Fournier and (Fournier, (Kharaka et 1975) Truesdell, 1973) 1979) al., 1982) B1 122 209 568 244 B2 122 209 567 242 Penebel B3 126 211 566 235 B4 120 227 688 254 B6 136 236 733 254 Banyuwedang B7 44 130 257 190 B9 108 172 395 236 Batur B10 122 171 396 240 B11 108 171 394 239 Banjar B13 120 205 602 255

V.4.4. Oxygen and hydrogen isotope considerations The deuterium and oxygen isotopic composition of thermal waters has been successfully applied to investigate the fluid origin, i.e., meteoric, marine or magmatic, mixing and physicochemical processes, such as, water-rock interaction and water-

CO2 isotope exchange (Arnason, 1977; Cinti et al., 2014; Craig et al., 1956; Gemici and Filiz, 2001; Giggenbach et al., 1983; Pichler, 2005; Purnomo and Pichler, 2014). The δ2H and δ18O composition of water is generally a good indicator of its origin. Hydrothermal fluids normally plot to the right of the LMWL due to exchange of 18O during water-rock interaction (Craig, 1966) or due to subsurface mixing with an andesitic water, as defined by (Giggenbach, 1992). In the δ2H vs. δ18O diagram all thermal waters, except the two lake waters, plot close to the local meteoric water line (LMWL) and weighted mean annual value for precipitation in the region, indicating

86

GEOTHERMAL SYSTEMS ON THE ISLAND OF BALI, INDONESIA local rainwater as the ultimate fluid source (Fig. 5.10). With the exception of sample B10 the thermal waters from Bali did not shift to the right of the LMWL indicating neither water-rock isotope exchange or mixing with an andesitic water. Neither process, however, can be completely ruled out because an initial 18O-shift to the right may have been later reversed by a subsequent isotope exchange between CO2 and

H2O. Such isotope shifts were observed in several CO2-rich aquifers and hydrothermal waters (Chiodini et al., 2000; Cinti et al., 2011; Cinti et al., 2014; Grassa et al., 2006; Vuataz and Goff, 1986). The Tirta Husada hot spring (B10) from the Batur group plots slightly right shifted from Toya Devasya (B11), potentially signaling some water-rock interaction. However, the TDS of B10 was relatively similar to B11 and B9 (Fig. 5.11). Although shallow magma is present at the Batur volcano (Geiger, 2014), an isotope isotope enrichment by magmatic input was not observed. The absence of a pronounced δ18O shift to heavier values also points towards a non-carbonate reservoir. Due to water-rock interaction in carbonate reservoirs heavier δ18O values are generally observed at such locations (Pasvanoglu and Chandrasekharam, 2011).

Fig. 5.10. δ2H vs. δ18O diagram with the local meteoric water line (LMWL) from Wandowo et al. (2001), GMWL from Craig (1961) and the mean annual value for

87

GEOTHERMAL SYSTEMS ON THE ISLAND OF BALI, INDONESIA precipitation in Jakarta from IAEA (2008). The thermal waters were of meteoric water origin. B10 is slightly horizontal right shifted from B11, probably was caused by water- rock interaction. The heavy water isotope of Lake Bratan and Lake Batur were caused by evaporation.

Fig. 5.11. TDS vs. δ18O diagram shows relatively similar TDS for the Batur thermal waters, B9, B10 and B11, hence indicating an insignificant 18O enrichment due to water-rock interaction.

The elevated δ2H and δ18O compositions of Lake Batur (B12) and Lake Bratan (B15), which were studied in detail by Varekamp and Kreulen (2000), were a result of evaporation in a cold lake. B12 was slightly enriched in δ18O compared to B15, likely due to the significant input of thermal waters into the lake that elevated its temperature. A lake with a higher temperature would produce a heavier δ18O due to evaporation (Gonfiantini, 1986; Varekamp and Kreulen, 2000).

V.5. Conclusions Two types of geothermal systems are present on Bali, the fault-hosted for Banyuwedang geothermal system and the volcano-hosted Penebel, Batur and Banjar geothermal systems. Contrary to assumptions by others we conclude that, although Bali is underlain by a carbonate basement, the geothermal reservoir rocks are of calc- alkaline magmatic origin. Steam phase separation occurred in Penebel and Banjar geothermal systems, while seawater input was confirmed for the fault-hosted

88

GEOTHERMAL SYSTEMS ON THE ISLAND OF BALI, INDONESIA geothermal system of Banyuwedang. Comparison of several geothermometers with actual reservoir temperature measurements and physicochemical considerations led to the conclusion that the Na/Li thermometer provided the best results for geothermal systems on Bali. Using this thermometer, the following reservoir temperatures were calculated: (1) Penebel (Bedugul) from 235 to 254 °C, (2) Batur 240 °C, (3) Banjar 255 °C and (4) Banyuwedang below 100 °C. The deuterium and oxygen isotopic composition indicated the thermal water were of meteoric water origin without significant of water-rock interaction and/or mixing with an andesitic magmatic fluid.

Acknowledgments B.J. Purnomo likes to thank the Ministry of Energy and Mineral Resources of Indonesia for the PhD scholarship grants number: 2579 K/69/MEM/2010. Thanks to Chen-Feng You for the boron isotope measurement, to Laura Knigge for the laboratory assistance, to Ketut Suardana for the help during fieldwork and to Britta Hinz-Stolle for an editorial review.

89

GEOTHERMAL SYSTEMS ON THE ISLAND OF BALI, INDONESIA

90

CONCLUSIONS AND OUTLOOK

VI. Conclusions and outlook

VI.1. Conclusions Geothermal systems in the Sunda volcanic island arc are mostly a volcano- hosted system. In this geological setting, the geothermal systems are considered mainly influenced by Quaternary volcanic activities. However, a closer look at the geothermal systems on Java showed that a fault-hosted geothermal system could exist exclusively, without any influence from the Quaternary magma. Specific conclusions coming from this study are:

1. The geothermal systems on Java could be classified into two types, volcano- hosted and fault-hosted. However, the classification could not be solely based on the location, either in a volcano or fault, magmatic fluids input has to be taken into account. Geothermal systems hosted by fault distributed in the Quaternary volcanic belt were supplied by magmatic fluids, thus could not be classified as a fault- hosted geothermal system.

2. The presence of magmatic fluids input in the volcano-hosted geothermal system - 2 18 elevated the HCO3 , Mg/Na ratios and stable isotope ( H and O). These characteristics were not identified in the fault-hosted geothermal systems.

3. In the fault-hosted geothermal systems deep circulated groundwater is conductive heated. Quaternary magma fluids input are absent due to the relatively impermeable Tertiary volcanic rocks. However, the Quaternary magma could be the heat source for the fault-hosted geothermal system, as indicated in the Cilayu and Cisolok geothermal systems.

4. Another heat source for the fault-hosted geothermal systems located in the Tertiary volcanic belt was considered a deep-seated magma, like has been identified elsewhere, e.g., the Alpine fault, New Zealand (Allis and Shi, 1995; Shi et al., 1996). The plates subduction in the south of Java has uplifted and thinned the crust of the sourthern part of the island, hence deep circulated groundwater can extract a deep-seated magma.

5. The boron isotope composition of thermal waters further distinguished volcano- hosted and fault-hosted geothermal systems. The fast ascent and absence of

91

CONCLUSIONS AND OUTLOOK

magmatic fluids input kept a light δ11B value of the fault-hosted geothermal systems and conversely, slow ascent and magmatic fluids input, which is favorable for minerals precipitation, promoted δ11B enrichment in the volcano-hosted geothermal systems.

6. The contrasting processes between acid chloride and acid sulfate crater lakes produced different B isotope signatures. The former had light δ11B values representing direct magmatic fluids supply from the subsurface. In contrast, the latter were δ11B enriched by vapor phase separation in the subsurface, followed by evaporation and B adsorption into clay minerals after discharge.

7. The study of geothermal systems on Bali identified two types of geothermal systems, a fault-hosted for Banyuwedang and a volcano-hosted for Penebel 2+ 2+ - (Bedugul), Batur and Banjar. The thermal waters had (Ca + Mg )/HCO3 ratios of approximately 0.4 and stable (2H and 18O) isotope typical of meteoric water origin, hence ruled out a carbonate host rock type. The host rocks are volcanic of calc- alkaline magma, as indicated by the K/Mg. Vapor phase separation was identified occurring in Penebel (Bedugul) and Banjar geothermal systems. The fault-hosted geothermal system of Banyuwedang is influenced by seawater input.

VI.2. Outlook This study reveals a complex geothermal systems in the Sunda volcanic island arc. The fault-hosted geothermal systems are not necessarily influenced by the Quaternary volcanic activities. Since every location has unique geological characteristics, the methods used in this study could be applied in other volcanic islands arcs, e.g., Japan, Philippines and New Zealand. The studies will benefit for better understanding the geothermal systems in such a geological setting, hence the model can be established. The absence of Quaternary magmatic fluid should be 34 2- 13 corroborated in the next studies by incorporating sulfur ( S of SO4 ) and carbon ( C of CO2) isotopes to trace the origin. Sulfur in a geothermal water can be sourced from

1) seawater, 2) magmatic gases (H2S and SO2) and 3) oxidation of sulfide minerals 13 (Pichler, 2005). C isotope of CO2 was applied to trace the contribution of mantle degassing and rock leaching, e.g., in Tutum Bay, Papua New Guinea by Pichler (2005) and the Sabatini volcanic area, Italy by Cinti et al. (2011). The geochemical study can be supported by details geological mapping and geophysical measurement to configure the subsurface condition in the boundary of the Quaternary volcanic belt

92

CONCLUSIONS AND OUTLOOK

and Tertiary volcanic belt. Although the Tertiary volcanic rocks are relatively impermeable, the presence of fractures potentially circulating fluids. Concerning on the boron isotope signature of the two contrast thermal crater lakes, acid sulfate and acid chloride, this study shows an overview of their different boron isotope signatures and provides the possible explanations. More samples representing both acid crater lakes from different location would corroborate this finding. The distinct characteristic of the two crater lakes, the Kawah putih, rich in B and light in δ11B values, and Kawah Sikidang, rich in B and heavy in δ11B values, deserve for a more detail study. A time series of sampling should be done to understand the fluctuation of δ11B isotope value. The behavior of B isotope after discharge can be determined by performing experiments of B isotope fractionation during evaporation and clay adsorption for low pH water with high and low Cl- concentrations, representing acid sulfate and acid choride, respectively. To date the effect of pH on B isotope fractionation during water-clay interaction is only known for higher pH (> 5), e.g., Yingkai and Lan (2001). The study on Bali provides evidences of a non carbonate reservoir type, thus the rich of CO2 in the geothermal brines likely was resulted by CO2-rich magma volatile due to a carbonate assimilation, as was proposed by Geiger (2014). 13C isotope of CO2 has to be incorporated in the upcoming studies. Access to the geothermal brines of the Bedugul geothermal field could be used to examine its contrasting physicochemical properties compared to the geothermal brines from Java (Bogie et al., 2008; Mulyadi et al., 2005; Purnomo and Pichler, 2014). The indication of significant thermal water supply into the Lake Batur needs to be confirmed by a systematic sampling. The sampling also can be applied to map the physicochemical characteristic distribution, leading to an identification of thermal water sources. The significant thermal water input might be originated from underwater thermal water discharge, hence a survey of underwater hot spot is needed.

93

CONCLUSIONS AND OUTLOOK

94

ACKNOWLEDGEMENTS

Acknowledgements

This PhD was made possible by funding from the Ministry of Energy and Mineral Resources of Republic Indonesia and University of Bremen.

I would like to first thank my supervisor, Prof. Dr. Thomas Pichler for the opportunity to do PhD, something beyond my imagination before you invited me. I learned the passion of doing fieldwork as a geoscientist from the experience working along with you during sampling campaign across Java in 2012. Words will not be enough to express my gratitude for your advices, supports and critical discussions during my 4 years doing PhD.

I thank to Prof. Dr. Peter LaFemina for the agreement to review and examine my PhD thesis. Thanks to Prof. Dr. Chen-Feng You for the boron isotope analysis. The fourth chapter of this thesis would not be possible without your contributions.

Special thank to Laura Knigge, Christine Albert and Christian Breuer for the help during laboratory analyses, also Britta Hinz-Stolle for your much effort and time on the editorial review of my manuscripts. Wisanukorn Poonchai is thanked for your German to English translation during our PHREEQC summer course. I also thank to Dr. Kay Hamer and Dr. Lars-Eric Heimburger for the fruitful advices, discussions and inputs. Pak Ketut, Maman and Seto are thanked for the assistance during sampling campaign on Java and Bali.

My fellow PhD students in the Geochemistry and Hydrogeology group, Wu Debo, Qi Pengfei and Ali Mozafarri, for the support, sharing, discussion and encouragement. Special for my office mates, Debo and Pengfei, thanks for always keep a warm atmosphere in our beloved office.

I thank to my fellow ‘Cator’-ers (Indonesian PhD students in Bremen): Riza Setiawan, Mba Asty, Mba Hesty, Teh Rima, Kang Ayi, Condro, Yusup, Haliq, Mba Enab, Mba Meity, and Rovi. Our togetherness kept my motivation and spirit on finishing this study.

95

ACKNOWLEDGEMENTS

Finally, my beloved wife, Novi, and daughter, Inca, deserve very special thank. I dedicate this for you both. Thanks to my big family, my father and mother (in heaven), my 9 brothers and sisters, Mas Kentut, Mba Pon, Mba Iten, Mba Sri, Mas Tris, Mba Nik, Mas Cipto, Mba Yani and Ulin. I could not achieve this level without all of you. My parents in law are thanked for taking care my small family during I stay in Germany.

96

REFERENCES

References

Abrenica, A.B., Harijoko, A., Kusumah, Y.I. and Boogie, I., 2010. Characteristics of hydrothermal alteration in part of the northern vapor-dominated reservoir of the Wayang Windu geothermal field, West Java. Proceedings World Geothermal Congress, Bali, Indonesia. Aggarwal, J.K. and Palmer, M.L., 1995. Fractionation of boron isotopes in Icelandic hydrothermal systems. Proceedings 17th NZ Geothermal Workshop: 67-70. Aggarwal, J.K., Palmer, M.R. and Ragnarsdottir, K.V., 1992. Boron isotopic composition of Icelandic hydrothermal systems. In: Y.K. Kharaka and A.S. Maest (Editors), 7th International Symposium on Water-Rock Interaction. AA Balkema, Park City, Utah, USA, pp. 893-895. Aggarwal, J.K., Palmer, M.R., Bullen, T.D., Ragnarsdottir, K.V. and Arnorsson, S., 2000. The boron isotope systematics of Icelandic geothermal waters: 1. Meteoric water charged systems. Geochimica et Cosmochimica Acta, 64: 580-585. Agyei, E.K. and McMullen, C.C., 1968. A study of the isotopic abundance of boron from various sources. Canadian Journal of Earth Sciences, 5: 921-927. Alam, M.A., Sanchez, P. and Parada, M.A., 2010. Interplay of volcanism and structural control in defining the geothermal system(s) along the Liquine-Ofqui fault zone, in the South-Central Chile. GRC Transactions, 34: 747-750. Allard, P., 1983. The origin of hydrogen, carbon, sulphur, nitrogen and rare gases in volcanic exhalations: evidence from isotope geochemistry. In: H. Tazieff and J. Sabroux (Editors), Forecasting Volcanic Events. Elsevier, New York, pp. 337-386. Allen, E.T. and Day, A.L., 1927. Steam wells and other thermal activity at the Geysers, California. Carnegie Institution of Washington publication, 378: 1-106. Allis, R.G. and Shi, Y., 1995. New insights to temperature and pressure beneath the central Southern Alps, New Zealand. New Zealand Journal of Geology and Geophysics, 38: 585-592. Alzwar, M., Akbar, N., Bachri, S., 1992. Geological map of the Garut and Pamengpeuk sheet, Java. Geological Research and Development Center, Bandung. Arana, V. and Panichi, C., 1974. Isotopic composition of steam samples from Lanzarote, Canary Islands. Geothermics, 3: 142-145.

97

REFERENCES

Arehart, G.B., Coolbaugh, M.F. and Poulson, S.R., 2003. Evidence for a magmatic source of heat for the Steamboat springs geothermal system using trace elements and gas geochemistry. GRC Transactions, 27: 269-274. Arnason, B., 1977. Hydrothermal systems in Iceland traced by Deuterium. Geothermics, 5: 125-151. Arnorsson, S. and Andresdottir, A., 1995. Processes controlling the distribution of boron and chlorine in natural waters in Iceland. Geochimica et Cosmochimica Acta, 59: 4125-4146. Arnorsson, S., 1970. Underground temperatures in hydrothermal areas in Iceland as deduced from the silica content of the thermal water. Geothermics Special Issue, 2: 536-541. Arnorsson, S., 1975. Application of the silica geothermometer in low temperature hydrothermal areas in Iceland. American Journal of Science, 275: 763. Asikin, S., Handoyo, A., Busono, H., Gafoer, S., 1992. Geological map of the Kebumen sheet, Java. Geological Research and Development Center of Indonesia, Bandung. Bachrun, Z.I., Soeroso, Suwana, A., 1995, 1995. Twelve years exploitation history of well Dieng-2, Dieng geothermal field, Indonesia, Proceedings of the World Geothermal Congress 1995, Florence, Italy, pp. 1769-1772. Bahar, I. and Girod, M., 1983. Controle structural du volcanisme Indonesien (, Java-Bali): application et critique de la methode de Nakamura. Bulletin de la Societe Geologique de France, 7: 609-614. Barth, S., 1993. Boron isotope variations in nature: a synthesis. Geologische Rundschau, 82: 640-651. Benderitter, Y. and Cormy, G., 1990. Possible approach to geothermal research and relative cost estimate. In: M.H. Dickson and M. Fanelli (Editors), Small geothermal resources. UNITAR/UNDP Centre for Small Energy Resources, Rome, Italy, pp. 61-71.

Bischoff, J.L. and Rosenbauer, R.J., 1996. The alteration of rhyolite in CO2 charged

water at 200 and 350°C: The unreactivity of CO2 at higher temperature. Geochimica et Cosmochimica Acta, 60: 3859-3867. Bogie, I., Kusumah, Y.I. and Wisnandary, M.C., 2008. Overview of the Wayang Windu geothermal field, West Java, Indonesia. Geothermics, 37: 347-365. Byerlee, J.D., 1978. Friction of Rocks. Pure and Applied Geophysics, 116: 615-626.

98

REFERENCES

Capaccioni, B., Vaselli, O., Tassi, F., Santo, A.P. and Huertas, A.D., 2011. Hydrogeochemistry of the thermal waters from the Sciacca Geothermal Field (Sicily, southern Italy). Journal of Hydrology, 396: 292–301. Catanzaro, E.J., Champion, C.E., Garner, E.L., Marinenko, G., Sappenfield, K.M. and Shield, W.R., 1970. Boric acid: Isotopic and assay standard reference materials. US National Bureau Standards Special Publication. Chadwick, J.P., Troll, V.R., Ginibre, C., Morgan, D., Gertisser, R., Waight, T.E. and Davidson, J.P., 2007. Carbonate assimilation at Merapi Volcano, Java, Indonesia: Insights from crystal isotope stratigraphy. Journal of Petrology, 48: 1793-1812. Chang, C.L., 1984. Triangular diagrams for predication of aquifer chemistry. Geothermal Resources Council Transactions, 8: 373-376. Chaussard, E. and Amelung, F., 2012. Precursory inflation of shallow magma reservoirs at west Sunda volcanoes detected by InSAR. Geophysical Research Letters, 39. Chiodini, G., Allard, P., Caliro, S. and Parello, F., 2000. 18O exchange between steam and carbon dioxide in volcanic and hydrothermal gases: Implications for the source of water. Geochimica et Cosmochimica Acta, 64: 2479–2488. Cinti, D., Procesi, M., Tassi, F., Montegrossi, G., Sciarra, A., Vaselli, O. and Quattrocchi, F., 2011. Fluid geochemistry and geothermometry in the western sector of the Sabatini Volcanic District and the Tolfa Mountains (Central Italy). Chemical Geology, 284: 160-181. Cinti, D., Tassi, F., Procesi, M., Bonini, M., Capecchiacci, F., Voltattorni, N., Vaselli, O. and Quattrocchi, F., 2014. Fluid geochemistry and geothermometry in the unexploited geothermal field of the Vicano–Cimino Volcanic District (Central Italy). Chemical Geology, 371: 96-114. Clark, I.D., Fritz, P., Michel, F.A. and Souther, J.G., 1982. Isotope hydrogeology and geothermometry of the Mount Meager geothermal area. Canadian Journal of Earth Sciences, 19: 1454-1473. Clements, B., Hall, R., Smyth, H.R. and Cottam, M.A., 2009. Thrusting of a volcanic arc: a new structural model for Java. Petroleum Geoscience, 15: 159–174. Cottam, M., Hall, R., Cross, L., Clements, B. and Spakman, W., 2010. Neogene subduction beneath Java, Indonesia: Slab tearing and changes in magmatism, EGU General Assembly 2010, Vienna. Craig, H., 1961. Isotopic variations in meteoric waters. Science, 133: 1702-1703.

99

REFERENCES

Craig, H., 1963. The isotopic geochemistry of water and carbon in geothermal areas. In: E. Tongiorgi (Editor), Nuclear geology on geothermal areas. Consiglio Nazional delle Ricerche, Pisa, pp. 17-53. Craig, H., 1966. Isotopic composition and origin of the Red Sea and Salton Sea geothermal brines. Science, 154: 1544-1548. Craig, H., Boato, G. and White, D.E., 1956. Isotopic geochemistry of thermal waters. National Academic Siences. National Research Council Publication 400: 29-38. Curray, J.R., Shor, G.G., Raitt, R.W. and Henry, M., 1977. Seismic refraction and reflection studies of crustal structure of the eastern Sunda and Western Banda Arcs. Journal of Geophysical Research, 82: 2479-2489. Dahlmann, A., Lange, G.J., 2003. Fluid–sediment interactions at Eastern Mediterranean mud volcanoes: a stable isotope study from ODP Leg 160. Earth and Planetary Science Letters 212, 377-391. D'Amore, F. and Bolognesi, L., 1994. Isotopic evidence for a magmatic contribution to fluids of the geothermal systems of Larderello, Italy, and the Geysers, California. Geothermics, 23: 21-32. Deegan, F.M., Troll, V.R., Freda, C., Misiti, V., Chadwick, J.P., McLeod, C.L. and Davidson, J.P., 2010. Magma-carbonate interaction processes and associated

CO2 release at Merapi Volcano, Indonesia: Insights from experimental petrology. Journal of Petrology, 51: 1027-1051. Delmelle, P. and Bernard, A., 1994. Geochemistry, mineralogy, and chemical modeling of the acid crater lake of Kawah Ijen Volcano, Indonesia. Geochimica et Cosmochimica Acta, 58: 2445-2460. Delmelle, P., Bernard, A., Kusakabe, M., Fischer, T.P. and Takano, B., 2000. Geochemistry of the magmatic-hydrothermal system of Kawah Ijen volcano, East Java, Indonesia. Journal of Volcanology and Geothermal Research, 97: 31-53. Dia, A.N., Rouelle, M.C., Boulegue, J., Comeau, P., 1999. Trinidad mud volcanoes: where do the expelled fluids come from? Geochimica et Cosmochimica Acta 63, 1023–1038. Dickson, A.G., 1990. Thermodynamics of dissociation of boric acid in synthetic seawater from 273.15 to 318.15 oK. Deep-sea Research 37, 755-766. Effendi, A.C., Kusnama, Hermanto, B., 1998. Geological map of the Bogor sheet, Java, 2 ed. Geological Research and Development Center of Indonesia, Bandung. Ellis, A.J. and Mahon, W.A.J., 1977. Chemistry and geothermal systems. Academic Press, New York, 392 pp.

100

REFERENCES

Ellis, A.J. and Wilson, S.H., 1960. The geochemistry of alkali metal ions in the Wairakei hydrothermal system. New Zealand Journal of Geology and Geophysics, 3: 593- 617. Ellis, A.J., 1979. Chemical geothermometry in geothermal systems. Geothermics, 25: 219-226. Faulds, J., Coolbaugh, M., Bouchot, V., Moeck, I. and Oğuz, K., 2010. Characterizing Structural Controls of Geothermal Reservoirs in the Great Basin, USA, and Western Turkey: Developing Successful Exploration Strategies in Extended Terranes, Proceedings World Geothermal Congress, Bali, Indonesia. Foster, G.L., Pogge von Strandmann, P.A.E. and Rae, J.W.B., 2010. Boron and magnesium isotopic composition of seawater. Geochemistry, Geophysics, Geosystems, 11. Fouillac, C. and Michard, G., 1981. Sodium/lithium ratio in water applied to geothermometry of geothermal reservoirs. Geothermics, 10: 55-70. Fournier, R.O. and Rowe, J.J., 1962. The solubility of cristobalite along the three phase curve, gas plus liquid plus cristobalite. American Mineralogist, 47: 897-902. Fournier, R.O. and Rowe, J.J., 1966. Estimation of underground temperatures from the silica content of water from hot springs and wet steam wells. American Journal of Science, 264: 685-697. Fournier, R.O. and Truesdell, A.H., 1973. An empirical Na-K-Ca geothermometer for natural waters. Geochimica et Cosmochimica Acta, 37: 1255-1275. Fournier, R.O., 1977. Chemical geothermometers and mixing models for geothermal systems. Geothermics, 5: 41-50. Fournier, R.O., 1979. A revised equation for the Na/K geothermometer. Geothermal Resources Council Transactions, 3: 221-224. Fournier, R.O., 1989. Lectures on geochemical interpretation of hydrothermal waters. UNU Geothermal Training Programme, Reykjavik, Iceland, Report 10: 73pp. Geiger, H., 2014. Characterising the magma supply system of Agung and Batur volcanoes on Bali, Indonesia, Uppsala University, Uppsala, Sweden, 118 pp. Gemici, U. and Filiz, S., 2001. Hydrochemistry of the Cesme geothermal area in Western Turkey. Journal of Volcanology and Geothermal Research, 110: 171- 187. Gerardo, J.Y., Nuti, S., D'Amore, F., Seastres JR, J.S. and Gonfiantini, R., 1993. Isotopic evidence for magmatic and meteoric water recharge and the processes

101

REFERENCES

affecting reservoir fluids in the Palinpinon geothermal systems, Philippines. Geothermics, 22: 521-533. Gertisser, R. and Keller, J., 2003. Trace element and Sr, Nd, Pb and O isotope variations in medium-K and high-K volcanic rocks from Merapi volcano, central Java, Indonesia: Evidence for the involvement of subducted sediments in Sunda arc magma genesis. Journal of Petrology, 44: 457-489. Giggenbach, W.F. and Stewart, M.K., 1982. Processes controlling the isotopic composition of steam and water discharges from steam vents and stem-heated pools in geothermal areas. Geothermics, 11: 71-80. Giggenbach, W.F., 1978. The isotopic composition of waters from the El Tatio geothermal field, northern Chile. Geochimica et Cosmochimica Acta, 42: 979-988. Giggenbach, W.F., 1984. Mass transfer in hydrothermal systems - A conceptual approach. Geochimica et Cosmochimica Acta, 48: 2693-2711. Giggenbach, W.F., 1988. Geothermal solute equilibria. Derivation of Na-K-Mg-Ca geoindicators. Geochimica et Cosmochimica Acta, 52: 2749-2765. Giggenbach, W.F., 1991. Chemical techniques in geothermal exploration. In: F. D'Amore (Editor), Application of geochemistry in geothermal reservoir development. UNITAR/UNDP, Rome, pp. 252-270. Giggenbach, W.F., 1992. Isotopic shifts in waters from geothermal and volcanic systems along convergent plate boundaries and their origin. Earth and Planetary Science Letters, 113: 495-510. Giggenbach, W.F., 1997. The origin and evolution of fluids in magmatic-hydrothermal systems. In: H.L. Barnes (Editor), Geochemistry of hydrothermal ore deposits. Wiley, pp. 737-796. Giggenbach, W.F., Gonfiantini, R., Jangi, B.L. and Truesdell, A.H., 1983. Isotopic and chemical composition of Parbati valley geothermal discharges, NW-Himalaya, India. Geothermics, 12: 199-222. Goff, F. and Janik, C.J., 2000. Geothermal Systems. In: H. Sigurdsson (Editor), Encyclopedia of volcanoes. Academic Press, pp. 817-834. Goff, F.E., Grigsby, C.O., Trujillo, P.E., Counce, D. and Kron, A., 1981. Geology, water geochemistry and geothermal potential of the Jemez springs are, Canon de San Diego, New Mexico. Journal of Volcanology and Geothermal Research, 10: 227- 244

102

REFERENCES

Gonfiantini, R., 1986. Environmental isotopes in lake studies. In: Frits, P., Fontes, J.C.H. (Eds.), Handbook of Environmental Isotope Geochemistry. Elsevier, Amsterdam: 113–168. Grassa, F., Capasso, G., Favara, R. and Inguaggiato, S., 2006. Chemical and isotopic composition of waters and dissolved gases in some thermal springs of Sicily and adjacent volcanic islands, Italy. Pure and Applied Geophysics, 163: 781–807. HACH, 2007. Alkalinity Titrator manual. Hadi, J., 1997. A hydrothermal alteration study to assess temperature and overprinting from the S-1 exploration well, north Darajat field, Indonesia. Diploma Project Report, Geothermal Institute, The University of Auckland. Hadiwidjojo, M.M.P., Samodra, H. and Amin, T.C., 1998. Geological map of the Bali sheet, Nusa Tenggara. Geological Research and Development Center, Bandung. Hall, R., 2002. geological and plate tectonic evolution of SE Asia and the SW Pacific: computer-based reconstructions, model and animations. Journal of Asian Earth Sciences, 20: 353-431. Hall, R., 2009. Indonesia, Geology. In: R. Gillespie and D. Clague (Editors), Encyclopedia of Islands. University of California Press, Berkeley, California, pp. 454-460. Hall, R., Clements, B., Smyth, H.R. and Cottam, M.A., 2007. A new interpretation of Java's structure. IPA-31st Annual Convention Proceedings. Hamilton, W., 1979. Tectonics of the Indonesian region. USGS Professional Paper, 1078. Harder, H., 1970. Boron content of sediments as a tool in facies analysis. Sedimentary Geology, 4: 153-175. Hedenquist, J.W., 1990. The thermal and geochemical structure of the Broadlands- Ohaaki geothermal system, New Zealand. Geothermics, 19: 151-185. Hemming, N.G. and Hanson, G.M., 1992. Boron isotopic composition and concentration in modern marine carbonates. Geochimica et Cosmochimica Acta, 56: 537-543. Henley, R.W. and Ellis, A.J., 1983. Geothermal Systems Ancient and Modern: a geochemical review. Earth-Science Reviews, 19: 1-50. Hensen, C., Wallmann, K., Schmidt, M., Ranero, C.R., Suess, E., 2004. Fluid expulsion related to mud extrusion off Costa Rica—a window to the subducting slab. Geology 32, 201–204.

103

REFERENCES

Hershey, J.R., Fernandez, M., Milne, P.J. and Millero, F.J., 1986. The ionization of boric acid in NaCl, Na-Ca-Cl, Na-Mg--Cl solutions at 25°C. Geochimica et Cosmochimica Acta, 50: 143-148. Hidayati, S. and Sulaeman, C., 2013. Magma supply system at Batur volcano inferred from volcanotectonic earthquakes and their focal mechanism. Indonesian Journal of Geology, 8: 97-105. Hochstein, M.P. and Browne, P.R.L., 2000. Surface manifestations of geothermal systems with volcanic heat sources. In: H. Sigurdsson (Editor), Encyclopedia of volcanoes. Academic Press, pp. 835-855. Hochstein, M.P. and Sudarman, S., 2008. History of geothermal exploration in Indonesia from 1970 to 2000. Geothermics, 37: 220–266. Hochstein, M.P., Mulyadi, R. and Joenos, E.J., 2005. The Bedugul Geothermal Field, Bali (Indonesia). IGA News, 59: 12-13. Hoffmann-Rothe, A., Ritter, O. and Haak, V., 2001. Magnetotelluric and geomagnetic modelling reveals zones of very high electrical conductvity in the upper crust of Central Java. Physics of the Earth and Planetary Interiors, 124: 131-151. IAEA, 2008. Atlas of isotope hydrology Asia and the Pasific. IAEA, Vienna, Austria. Ishikawa, T. and Nakamura, E., 1992. Boron isotope systematics of island arc volcanics: implications for subducted components in the wedge mantle, 29th International Geological Congress IGC, Kyoto, Japan, pp. 182. Italiano, F., Bonfanti, P., Pizzino, L. and Quattrocchi, F., 2010. Geochemistry of fluids discharged over the seismic area of the Southern Apennines (Calabria region, Southern Italy): Implications for Fluid-Fault relationships. Applied Geochemistry, 25(4): 540-554. Kaasalainen, H. and Stefánsson, A., 2012. The chemistry of trace elements in surface geothermal waters and steam, Iceland. Chemical Geology, 330-331: 60-85. Kakihana, H., Ossaka, T., Oi, T., Musashi, M., Okamoto, M. and Nomura, M., 1987. Boron isotopic ratios of some hot spring waters in the Kusatsu-shirane area, Japan. Geochemical Journal, 21: 133-137. Kanzaki, T., Yoshida, M., Nomura, M., Kakihana, H., Ozawa, T., 1979. Boron isotopic composition of fumarolic condensates and sassolites from Satsuma Iwo-jima, Japan Geochimica et Cosmochimica Acta 43, 1859-1863. Kasemann, S.A., Meixner, A., Erzinger, J., Viramonte, J.G., Alonso, R.N., Franz, G., 2004. Boron isotope composition of geothermal fluids and borate minerals from

104

REFERENCES

salar deposits (central Andes/NW Argentina). Journal of South American Earth Sciences 16, 685–697. Kastner, M., Elderfield, H., Martin, J.B., 1991. Fluids in convergent margins: What do we know about their composition, origin, role in diagenesis and importance for oceanic chemical fluxes? Philosophical Transactions: Physical Sciences and Engineering 335, 243-259. Kharaka, Y.K. and Mariner, R.H., 1989. Chemical geothermometers and their application to formation waters from sedimentary basins. In: N.D. Naeser and T.H. McCulloch (Editors), Thermal history of sedimentary basins: methods and case histories. Springer-Verlag, New York, pp. 99-117. Kharaka, Y.K., Lico, M.S. and Law, L.M., 1982. Chemical geothermometers applied to formation waters, Gulf of Mexico and California basins. American Association of Petroleum Geologists Bulletin, 66: 588. Koga, A., 1970. Geochemistry of the waters discharged from drillholes in Otake and Hatchobaru areas (Japan). Geothermics, 2: 1422-1425. Kuhn, M., 2004. Reactive flow modeling of hydrothermal systems. Springer-Verlag, Berlin Heidelberg, 261 pp. Lachenbruch, A.H. and Sass, J.H., 1980. Heat flow and energetics of the San Andreas Fault zone. Journal of Geophysical Research, 85: 6185-6222. Layman, E.B. and Soemarinda, S., 2003. The Patuha vapor-dominated resources West Java, Indonesia. Proceedings 28th Workshop Geothermal Reservoir Engineering, Stanford University, California. Leeman, W.P., Tonarini, S., Pennisi, M., Ferrara, G., 2005. Boron isotopic variations in fumarolic condensates and thermal waters from Vulcano Island, Italy: Implications for evolution of volcanic fluids. Geochimica et Cosmochimica Acta 69, 143-163. Leeman, W.P., Vocke, R.D. and McKibben, M.A., 1990. Boron isotope studies of geothermal fluids Eos, 71: 1686-1687. Leeman, W.P., Vocke, R.D. and McKibben, M.A., 1992. Boron isotopic fractionation between coexisting vapor and liquid in natural geothermal systems. In: Y.K. Kharaka and A.S. Maest (Editors), Water-Rock Interaction. Balkema, pp. 1007- 1010. Lemarchand, D., Schott, J. and Gaillardet, J., 2007. How surface complexes impact boron isotope fractionation: evidence from Fe and Mn oxides sorption experiments. Earth and Planetary Science Letters, 260: 277-296.

105

REFERENCES

Levet, S., Toutain, J.P., Munoz, M., Berger, G., Negrel, P., Jendrzejewski, N., Agrinier, P. and Sortino, F., 2002. Geochemistry of the Bagne`res-de-Bigorre thermal waters from the North Pyrenean Zone (France). Geofluids, 2: 25–40. Liebscher, A., Meixner, A., Romer, R.L., Heinrich, W., 2005. Liquid-vapor fractionation of boron and boron isotopes: Experimental calibration at 400 °C/23 MPa to 450 °C/42 MPa. Geochimica et Cosmochimica Acta 69, 5693-5704. Lowenstern, J.B., 2001. Carbon dioxide in magmas and implications for hydrothermal systems. Mineralium Deposita, 36: 490-502. Macpherson, C.G. and Hall, R., 1999. Tectonic controls of geochemical evolution in arc magmatism of SE Asia, 4th PACRIM Congress, Australian Institute of Mining and Metallurgy, pp. 359–368. Macpherson, C.G. and Hall, R., 2002. Timing and tectonic controls on magmatism and ore generation in an evolving orogen: Evidence from Southeast Asia and the western Pacific. In: D.J. Blundell (Editor), The Timing and Location of Major Ore Deposits in an Evolving Orogen. Geological Society, London, pp. 49–67. Majumdar, N., Mukherjee, A.L. and Majumdar, R.K., 2009. Mixing hydrology and chemical equlibria in Bakreswar geothermal area, Eastern India. Journal of Volcanology and Geothermal Research, 183: 201-212. Marziano, G.I., Gaillard, F. and Pichavant, M., 2007. Limestone assimilation and the

origin of CO2 emissions at the Alban Hills (Central Italy): Constraints from experimental petrology. Journal of Volcanology and Geothermal Research, 166: 91–105. Marziano, G.I., Gaillard, F., Scaillet, B., Pichavant, M. and Chiodini, G., 2009. Role of

non-mantle CO2 in the dynamics of volcano degassing: The Mount Vesuvius example. Geology, 37: 319-322. McCarthy, K.T., Pichler, T. and Price, R.E., 2005. Geochemistry of Champagne Hot Springs shallow hydrothermal vent field and associated sediments, Dominica, Lesser Antilles. Chemical Geology, 224(1-3): 55-68. McMullen, C.C., Cragg, C.B. and Thode, H.G., 1961. Absolute ratios of 11B/10B in Searles Lake borax Geochimica et Cosmochimica Acta, 23: 147-149. Millot, R., Asmundsson, R., Négrel, P., Sanjuan, B. and Bullen, T.D., 2009. Multi- isotopic (H, O, C, S, Li, B, Si, Sr, Nd) approach for geothermal fluid characterization in Iceland, Goldschmidt Conference 2009, Davos, Switzerland.

106

REFERENCES

Minissale, A.A. and Duchi, V., 1988. Geothermometry on fluids circulating in a carbonate reservoir in north-central Italy. Journal of Volcanology and Geothermal Research, 35: 237-252. Müller, R.D., Gaina, C. and Clarke, S., 2000. Seafloor spreading around Australia In: J.J. Veevers (Editor), Billion-Year Earth History of Australia and Neighbours in Gondwanaland. GEMOC Press, Sydney, pp. 18-28. Mulyadi, A., Joenoes, E.J. and Widiasari, N.M., 2005. Bedugul geothermal prospect and developments Proceedings World Geothermal Congress 2005, Antalya, Turkey. Mulyadi, Ashat, A., 2011. Reservoir modeling of the northern vapor-dominated two- phase zone of the Wayang Windu geothermal field, Java, Indonesia, Proceedings Thirty-Sixth Workshop on Geothermal Reservoir Engineering, Stanford University, Stanford, California. Musashi, M., Nomura, M., Okamoto, M., Ossaka, T., Oi, T. and Kakihana, H., 1988. Regional variation in the boron isotopic composition of hot spring waters from central Japan. Geochemical Journal, 22: 205-214. Musashi, M., Oi, T., Ossaka, T. and Kakihana, H., 1991. Natural boron isotope fractionation between hot spring water and rock in direct contact. Isotopenpraxis Isotopes in Environmental and Health Studies, 27: 163-166. Nicholls, I.A. and Whitford, D.J., 1983. Potassium-rich volcanic rocks of the Muriah Complex, Java, Indonesia: Products of multiple magma sources? Journal of Volcanology and Geothermal Research, 18: 337-359. Nicholson, K., 1993. Geothermal fluids: chemistry and exploration techniques. Springer Verlag, 266 pp. Nomura, M., Kanzaki, T., Ozawa, T., Okamoto, M. and Kakihana, H., 1982. Boron isotopic composition of fumarolic condensates from some volcanoes in Japanese island arcs. Geochimica et Cosmochimica Acta, 46: 2403-2406. Ohba, T., Hirabayashi, J. and Nogami, K., 2000. D/H and 18O/16O ratios of water in the crater lake at Kusatsu-Shirane volcano, Japan. Journal of Volcanology and Geothermal Research, 97: 329-346. Oi, T., Nomura, M., Musashi, M., Ossaka, T., Okamoto, M. and Kakihana, H., 1989. Boron isotopic compositions of some boron minerals. Geochimica et Cosmochimica Acta, 53: 3189-3195. Oi, T., Ogawa, J. and Ossaka, T., 1993. Boron isotopic compositions of Shimogamo hot springs, Izu, Japan. Geochemical Journal, 27: 147-154.

107

REFERENCES

Palmer, M.R. and Sturchio, N.C., 1990. The boron isotope systematics of the Yellowstone National Park (Wyoming) hydrothermal system: a reconnaissance. Geochimica et Cosmochimica Acta, 54: 2811-2815. Palmer, M.R., 1991. Boron-isotope systematics of Halmahera arc (Indonesia) lavas: Evidence for involvement of the subducted slab. Geology, 19: 215-217. Palmer, M.R., Spivack, A.J. and Edmond, J.M., 1987. Temperature and pH controls over isotopic fractionation during adsorption of boron on marine clays. Geochimica et Cosmochimica Acta, 51: 2319-2323. Palmer, M.R. and Sturchio, N.C., 1990. The boron isotope systematics of the Yellowstone National Park (Wyoming) hydrothermal system: a reconnaissance. Geochimica et Cosmochimica Acta 54, 2811-2815. Parkhurst, D.L. and Appelo, C.A.J., 1999. User´s guide to PHREEQC (Version 2)—a computer program for speciation, batch-reaction, one-dimensional transport, and inverse geochemical calculations, United States Geological Survey. Pasvanoglu, S. and Chandrasekharam, D., 2011. Hydrogeochemical and isotopic study of thermal and mineralized waters from the Nevşehir (Kozakli) area, Central Turkey. Journal of Volcanology and Geothermal Research, 202: 241-250. Pasvanoglu, S. and Gultekin, F., 2012. Hydrogeochemical study of the Terme and Karakurt thermal and mineralized waters from Kirsehir Area, central Turkey. Environmental Earth Sciences, 66: 169–182. Pichler, T., 2005. Stable and radiogenic isotopes as tracers for the origin, mixing and subsurface history of fluids in submarine shallow-water hydrothermal systems. Journal of Volcanology and Geothermal Research, 139: 211-226. Pichler, T., Veizer, J. and Hall, G.E.M., 1999. The chemical composition of shallow- water hydrothermal fluids in Tutum Bay, Ambitle island, Papua New Guinea and their effect on ambient seawater. Marine Chemistry, 64: 229-252. Prasetio, R., Abidin, Z. and Yulizar, Y., 2010. Isotope and gas geochemistry of Dieng geothermal field, Indonesia. Proceedings World Geothermal Congress, Bali, Indonesia. Purnomo, B.J. and Pichler, T., 2014. Geothermal systems on the island of Java, Indonesia. Journal of Volcanology and Geothermal Research, 285: 47–59. Purnomo, B.J., Pichler, T. and You, C.F., 2015. Boron isotope variations in geothermal systems on Java, Indonesia. in preparation.

108

REFERENCES

Rahardjo, W., Sukandarrumidi, Rosidi, H.M.D., 1995. Geological map of the Yogyakarta sheet, Java. Geological Research and Development Center of Indonesia, Bandung. Rejeki, S., Hadi, J., Suhayati, I., 2005. Porosity study for detail reservoir characterization in Darajat geothermal field, West Java, Indonesia, World Geothermal Congress 2005, Antalya, Turkey. Reubi, O. and Nicholls, I.A., 2004. Magmatic evolution at Batur volcanic field, Bali, Indonesia: petrological evidence for polybaric fractional crystallization and implications for caldera-forming eruptions. Journal of Volcanology and Geothermal Research, 138: 345–369. Rowe Jr, G.L., 1994. Oxygen, hydrogen, and sulfur isotope systematics of the crater lake system of Poas Volcano, Costa Rica. Geochemical Journal, 28: 263-287. Saemundsson, K., 2009. Geothermal systems in global perspective, Short Course IV on Exploration for Geothermal Resources. UNU-GTP, Lake Naivasha, Kenya. Samodra, H., Gafoer, S. and Tjokrosapoetro, S., 1992. Geological map of the Pacitan quadrangle, Java. Geological Research and Development Centre of Indonesia, Bandung. Sarsito, D.A., Meilano, I., Abidin, H.Z. and Kato, T., 2011. Five years geodetic GPS observation in the West Java of Java Island. Japan Geoscience Union Meeting. Satyana, A.H., 2007. Central Java, Indonesia - a 'terra incognita' in petroleum exploration: new considerations on the tectonic evolution and petroleum implications. Proceedings Indonesian Petroleum Association. Schellart, W.P., Lister, G.S. and Toy, V.G., 2006. A Late Cretaceous and Cenozoic reconstruction of the Southwest Pacific region: Tectonics controlled by subduction and slab rollback processes. Earth-Science Reviews, 76: 191-233. Schwarcz, H.P., Agyei, E.K. and McMullen, C.C., 1969. Boron isotopic fractionation during clay adsorption from seawater. Earth and Planetary Science Letters, 6: 1- 5. Setijadji, L.D., 2010. Segmented volcanic arc and its association with geothermal fields in Java island, Indonesia. Proceedings World Geothermal Congress. Shi, Y., Allis, R.G. and Davey, F., 1996. Thermal Modeling of the Southern Alps, New Zealand. Pure and Applied Geophysics, 146: 469-501. Silitonga, P.H., 1973. Geological map of the bandung quadrangle, Java. Geological Research and Development Centre of Indonesia, Bandung.

109

REFERENCES

Simandjuntak, T.O. and Barber, A.J., 1996. Contrasting tectonic styles in the Neogene orogenic belts of Indonesia. Geological Society, London, Special Publications, 106: 185-201. Situmorang, B., Siswoyo, Thajib, E. and Paltrinieri, F., 1976. Wrench fault tectonics and aspects of hydrocarbon accumulation in Java, 5th Annual Convention of Indonesian Petroleum Association (IPA), pp. 53-66. Smith, C.L., Ficklin, W.H., Thompson, J.M., 1987. Concentrations of arsenic, antimony, and boron in steam and steam condensate at the Geysers, California. Journal of Volcanology and Geothermal Research 32, 329-341. Smyth, H.R., Hall, R. and Nichols, G.J., 2008. Cenozoic volcanic arc history of East Java, Indonesia: The stratigraphic record of eruptions on an active continental margin. In: A.E. Draut, P.D. Clift and D.W. Scholl (Editors), Formation and Applications of the Sedimentary Record in Arc Collision Zones. Geological Society of America, pp. 199–222. Soeria-Atmadja, R., Maury, R.C., Bellon, H., Pringgoprawiro, H., Polve, M. and Priadi, B., 1994. Tertiary magmatic belts in Java. Journal of Southeast Asian Earth Sciences, 9: 13-27. Spivack, A.J. and Edmond, J.M., 1987. Boron isotope exchange between seawater and oceanic crust. Geochimica et Cosmochimica Acta, 51: 1033-1042. Spivack, A.J., Berndt, M.E. and Seyfried Jr., W.E., 1990. Boron isotope fractionation during supercritical phase separation. Geochimica et Cosmochimica Acta, 54: 2337-2339. Sriwana, T., van Bergen, M.J., Varekamp, J.C., Sumarti, S., Takano, B., van Os, B.J.H. and Leng, M.J., 2000. Geochemistry of the acid Kawah Putih lake, Patuha Volcano, West Java, Indonesia. Journal of Volcanology and Geothermal Research, 97: 77-104. Sudarman, S., Boedihardi, M., Pudyastuti, K. and Bardan, 1995. Kamojang geothermal field: 10 year operation experience. Proceedings of the World Geothermal Congress, Florence, Italy, 3: 1773-1777. Sujatmiko and Santosa, S., 1992. Geological map of the Leuwidamar quadrangle, Java. Geological Research and Development Centre of Indonesia, Bandung. Swihart, G.H., Moore, P.B. and Callis, E.L., 1986. Boron isotopic composition of marine and non-marine evaporite borates. Geochimica et Cosmochimica Acta, 50: 1297- 1301.

110

REFERENCES

Taran, Y., Rouwet, D., Inguaggiato, S. and Aiuppa, A., 2008. Major and trace element geochemistry of neutral and acidic thermal springs at El Chichón volcano, Mexico. Journal of Volcanology and Geothermal Research, 178(2): 224-236. Taran, Y.A., Pokrovsky, B.G. and Esikov, A.D., 1989. Deuterium and oxygen-18 in fumarolic steam and amphiboles from some Kamchatka volcanoes: 'andesitic waters'. Doklady Akademii Nauk SSSR, 304: 440-443. Trompetter, W.J., Reyes, A.G., Vickridge, I.C. and Markwitz, A., 1999. Lithium and boron distributions in geological samples. Nuclear Instruments and Methods in Physics Research, 158: 568-574. Truesdell, A.H. and Fournier, R.O., 1977. Procedure for estimating the temperature of a hot water component in a mixed water using a plot of dissolved silica vs enthalpy. USGS Journal of Research, 5: 49-52. Truesdell, A.H., 1975. Summary of Section III. Geochemical techniques in exploration, 2nd UN Symposium on the development and use of geothermal resources, San Fransisco. Truesdell, A.H., Haizlip, J.R., Armannsson, H. and D’Amore, F., 1989. Origin and transport of chloride in superheated geothermal steam. Geothermics, 18: 295 – 304. Valentino, G.M. and Stanzione, D., 2003. Source processes of the thermal waters from the Phlegraean Fields (Naples, Italy) by means of the study of selected minor and trace elements distribution. Chemical Geology, 194(4): 245-274. Van Bemellen, R.W., 1949. The Geology of Indonesia. Martinus Nijhoff, the Hague, vol. 1A: 732 p. Varekamp, J.C. and Kreulen, R., 2000. The stable isotope geochemistry of volcanic lakes, with examples from Indonesia. Journal of Volcanology and Geothermal Research, 97: 309-327. Vengosh, A., Heumann, K.G., Juraske, S. and Kasher, R., 1994a. Boron Isotope Application for Tracing Sources of Contamination in Groundwater. Environment Science & Technology, 28: 1968-1974 Vengosh, A., Kolodny, Y., Starinsky, A., Chivas, A.R. and McCulloch, M.T., 1991a. Coprecipitation and isotopic fractionation of boron in modem biogenic carbonates. Geochimica et Cosmochimica Acta, 55: 2901-2910. Vengosh, A., Starinsky, A., Kolodny, Y. and Chivas, A.R., 1991b. Boron-isotope geochemistry as a tracer for the evolution of brines and associated hot springs from the Dead Sea, Israel. Geochimica et Cosmochimica Acta, 55: 1689-1695.

111

REFERENCES

Vengosh, A., Starinsky, A., Kolodny, Y. and Chivas, A.R., 1994b. Boron isotope geochemistry of thermal springs from the Northern Rift Valley, Israel. Journal of Hydrology, 162: 155-169. Vengosh, A., Starinsky, A., Kolodny, Y., Chivas, A.R. and Raab, M., 1992. Boron isotope variations during fractional evaporation of sea water: New constraints on the marine vs. nonmarine debate. Geology, 20: 799-802. Vuataz, F.D. and Goff, F., 1986. Isotope geochemistry of thermal and nonthermal waters in the Valles Caldera, Jemez Mountains, Northern New Mexico. Journal of Geophysical Research, 91: 1835–1853. Wandowo, Abidin, Z., Alip and Djiono, 2001. Daerah resapan air tanah Cekungan Jakarta. Risalah Pertemuan Ilmiah Penelitian dan Pengambangan Aplikasi Isotop dan Radiasi. Wang, B.S., You, C.F., Huang, K.F., Wu, S.F., Aggarwal, S.K., Chung, C.H. and Lin, P.Y., 2010. Direct separation of boron from Na- and Ca-rich matrices by sublimation for stable isotope measurement by MC-ICP-MS. Talanta, 82: 1378- 1384. Watanabe, K., Yamanaka, T., Harijoko, A., Saitra, C. and Warmada, I.W., 2010. Caldera activities in north Bali, Indonesia. Journal of Southeast Asian Applied Geology, 2: 283–290. Wheller, G.E. and Varne, R., 1986. Genesis of dacitic magmatism at Batur Volcano, Bali, Indonesia: Implications for the origins of stratovolcano calderas. Journal of Volcanology and Geothermal Research, 28: 363-378. Wheller, G.E., Varne, R., Foden, J.D. and Abbot, M.J., 1987. Geochemistry of Quaternary volcanism in the Sunda-, Indonesia, and three-component genesis of island-arc basaltic magmas. Journal of Volcanology and Geothermal Research, 32: 137-160. White, D.E., 1957. Thermal waters of volcanic origin. Bulletin of The Geological Society of America, 68: 1637-1658. White, D.E., Muffler, L.P.J. and Truesdell, A.H., 1971. Vapour-dominated hydrothermal systems compared with hot-water systems. Economic Geology, 66: 75-97. Whitford, D.J., Nicholls, I.A. and Taylor, S.R., 1979. Spatial variations in the geochemistry of quaternary lavas across the Sunda Arc in Java and Bali. Contributions to Mineralogy and Petrology, 70: 341-356.

112

REFERENCES

Wisian, K.W., Blackwell, D.D. and Richards, M., 1999. Heat flow in the western United States and extensional geothermal systems. Proceedings 24th Workshop on Geothermal Reservoir Engineering Stanford University. Xiao, J., Xiao, Y.-K., Jin, Z.-D., He, M.-Y. and Liu, C.-Q., 2013. Boron isotope variations and its geochemical application in nature. Australian Journal of Earth Sciences, 60: 431-447. Yingkai, X. and Lan, W., 2001. The effect of pH and temperature on the isotopic fractionation of boron between saline brine and sediments. Chemical Geology 171, 253-261. You, C.-F., Gieskes, J.M., Lee, T., Yui, T.-F., Chen, H.-W., 2004. Geochemistry of mud volcano fluids in the Taiwan accretionary prism. Applied Geochemistry 19, 695- 707. Yuan, J., Guo, Q. and Wang, Y., 2014. Geochemical behaviors of boron and its isotopes in aqueous environment of the Yangbajing and Yangyi geothermal fields, Tibet, China. Journal of Geochemical Exploration. Zuhro, A.A., 2004. Numerical modelling of the Kamojang geothermal systems, Indonesia, Geothermal training programme, Reykjavík, Iceland, pp. 475-492.

113

REFERENCES

114