You are on page 1of 32

Available online at www.sciencedirect.

com

Geothermics 37 (2008) 300–331

An overview of the Awibengkok geothermal


system, Indonesia
James Stimac ∗ , Gregg Nordquist, Aquardi Suminar,
Lutfhie Sirad-Azwar
Chevron Geothermal Salak, Ltd., 11th Floor Sentral Senayan I, Jl. Asia Afrika No. 8,
Jakarta 10270, Indonesia
Received 6 October 2007; accepted 1 April 2008

Abstract
The Awibengkok (Salak) geothermal system is a liquid-dominated, fracture-controlled reservoir with
benign chemistry and low-to-moderate non-condensable gas content. The geothermal system is hosted
mainly by andesitic-to-rhyodacitic rocks, and floored by Miocene marine sedimentary rocks cut by igneous
intrusions. The volcanic sequence is capped by an 8400-year-old phreatic explosion breccia, rhyolite fallout
tuff (>8400 years and <40,000 years), rhyolite lavas, domes and related tuffs (≥40–120 ka), and dacite-to-
rhyodate lavas and domes (185–280 ka) that were erupted across the eastern part of the field from NNE-
trending vents controlled by a major fault. More regionally extensive basaltic–andesite to andesite volcanic
centers are mostly between 180 and 1610 ka old.
Surface and subsurface fault patterns, formation image logs and tracer studies indicate strongly anisotropic
permeability aligned with the dominant N to NE fracture trend, dividing the field into a number of subcom-
partments that are locally connected by fractured aquifers and NW- and E-W-trending fractures. Shallow
argillic alteration gives way with increasing depth and temperature to argillic–phyllic and propylitic zones,
with the latter accounting for the bulk of the fluid produced from the geothermal system.
The commercial Awibengkok reservoir is a moderate-to-high temperature (240–312 ◦ C) geothermal
resource with high fracture permeability, moderate porosity (mean = 10.6%) and moderate-to-low matrix
permeability (geometric mean = 0.026 md). The principal deep upflow zone, with fluid temperatures in the
275–312 ◦ C range, is located in the western part of the field. The ascending fluids move up along N- or
NNE-trending structures that breach low-permeability tuff layers in the central and east-central parts of the
field. Fluids in the central part of the reservoir are uniform in composition and temperature, representing the
mixing of upflow and convective reflux. Fluids ascend and flow laterally to the shallow top of the reservoir in
the eastern area near drilling pads Awi 1 and 13 (fluid production temperatures: 240 to 270 ◦ C). The eastern

∗ Corresponding author. Tel.: +62 21 5798 4547; fax: +62 21 573 0981.
E-mail address: jstimac@chevron.com (J. Stimac).

0375-6505/$30.00 © 2008 Published by Elsevier Ltd.


doi:10.1016/j.geothermics.2008.04.004
J. Stimac et al. / Geothermics 37 (2008) 300–331 301

reservoir is shallower and has been more affected by volcanic eruptions during the period of hydrothermal
activity. This has led to steam loss, condensation, and local ingress of shallow steam-heated water that has
resulted in lower temperatures in this area.
Delineation of potential fluid production or injection zones adjoining northern and western margins of
the proven reservoir has been accomplished by integration of new geophysical, geologic and geochemical
studies, followed by drilling. A facies model for small calderas was developed and used during the assessment
of the western area.
© 2008 Published by Elsevier Ltd.

Keywords: Awibengkok; Gunung Salak; West Java; Indonesia; Reservoir geology; Caldera facies model

1. Introduction

The Awibengkok geothermal field, also known as Salak, is located 60 km south of Jakarta
on the island of Java, Indonesia (Fig. 1). The original exploration contract area, including the
current proven field, lies in a highland on the southwestern flank of the Gunung Salak volcano
(2211 m above sea level; m asl), for which the contract area was named. The reservoir is the largest
producer of geothermal power in Indonesia (Ibrahim et al., 2005), and is the fourth largest power
producer from a liquid-dominated field in the world, after Cerro Prieto, Tongonan and Bulalo
(Bertani, 2005). A proven reservoir area of 18 km2 and an installed capacity of 377 MWe yields
a power density of about 20 MWe/km2 despite all injection being done infield.

1.1. Exploration, contractual framework and operational history

The Gunung Salak area was explored and developed for commercial power production by
Union Oil of California (Unocal Geothermal Indonesia; UGI), and has been managed by Chevron
Geothermal Salak since August of 2005 when Chevron acquired Unocal. In February 1982, UGI
entered into a Joint Operation Contract with the Indonesian National Oil Company (Pertamina) for
exploration and development of the geothermal resources of the Gunung Salak Contract Area, and
an Energy Sales Contract with Pertamina and the Indonesian National Power Company (PLN)
for sale of steam to PLN. The contracts allowed UGI to supply steam for up to 495 MWe of
electric power generation. UGI conducted intensive exploration from 1982 to 1984, resulting in
five prospects, with further delineation of the two most prospective areas, Awibengkok and Ratu
(Mosby, 1984; Rohrs, 1986). Fifteen deep wells were drilled from 1983 to 1993, twelve of which
proved a large geothermal resource in the Awibengkok area (Fig. 2). The remaining three wells
delineated a small sub-commercial geothermal system at Ratu.
Commercial power generation at Awibengkok began with an 110 MWe plant consisting of
two 55 MWe units operated by PLN in 1994 (Murray et al., 1995). Production was expanded to
330 MWe in 1998 when a third power plant (Unit 3, also operated by PLN) was installed adjacent
to Units 1 and 2, and an additional 220 MWe plant (Units 4–6) was constructed at a new location
adjacent to the Awi 1 pad (Soeparjadi et al., 1998; Acuña et al., this issue, Fig. 2).
Following the Asian Economic Crisis of 1997–1998, a dispute arose regarding payment for
power that necessitated a reduction in UGI’s investment, and put investigation of optimization and
expansion opportunities on hold. However, new commercial terms were eventually negotiated that
prioritized further expansion of production. As part of this agreement, turbine output was raised to
377 MWe in 2002. Increased revenue also allowed resumption of geologic and geophysical evalu-
ation of the field’s margins, and possible extensions of the reservoir were identified by magnetotel-
302 J. Stimac et al. / Geothermics 37 (2008) 300–331

Fig. 1. Location map. (a) Image of Indonesia with location of study area. (b) Image of West Java with major cities
and volcanic centers. Awibengkok and Darajat contract areas (dashed polygons) and other producing geothermal fields
(circles) are shown.

luric (MT) and time domain electromagnetic (TDEM) surveys. As described below, low-resistivity
anomalies that may represent extensions of the field’s smectite clay cap were identified.
Make-up drilling campaigns in 2002, 2004 and 2006–2007 have maintained the supply of
steam at plant requirements. As of mid-2007, 81 wells had been drilled, with 51 being used
for production and 18 for injection (Fig. 2). Drilling since the second quarter of 2006 has pro-
vided eight infield producers and four outfield delineation wells. For the delineation program,
wells Awi 17 and 18 were drilled to the west of Awi 9 to determine the potential for injection,
and wells Awi 19-2 and 19-3 to the north of Awi 3 to establish production potential. Delin-
eation of new outfield injection areas should eventually allow conversion of some injection areas
within the commercial resource to production. A more complete account of the resource manage-
J. Stimac et al. / Geothermics 37 (2008) 300–331 303

Fig. 2. Map of the Gunung Salak contract area. The Salak contract area (solid black line) is shown along with the
commercial reservoir boundary (dashed black line), as well as the thermal features and the interpreted extent of clay
alteration based on MT-TDEM surveys (dashed grey lines). Wells are numbered based on the sequence of drill sites
established. Exploration wells drilled at Ratu are also shown. Cihideung hot spring lies 4 km north of the map boundary.
Lithology and alteration along cross-section A–A is shown in Figs. 4b and 7, respectively.

ment performance, strategies, and challenges at Awibengkok can be found in Acuña et al. (this
issue).
The geologic overview given below describes the main characteristics of the Awibengkok
geothermal reservoir, and provides the context for a more detailed presentation of reservoir
changes under exploitation and related field management issues, like the one given by Acuña
et al. (this issue). In particular, the Awibengkok field is an example of an exceptionally large
liquid-dominated resource associated with shallow intermediate-to-silicic intrusions in a mature
island arc setting, where detailed geoscientific characterization studies have been conducted.
The commercial resource is spatially associated with andesitic-to-rhyolitic volcanism that has
occurred over the past 330 ka, especially silicic volcanics erupted in the last 280 ka along a major
NNE-trending structure. Management of the field has relied heavily on a detailed conceptual
model that was developed to serve as the basis for numerical simulations, and has evolved as
more information has been gathered.

1.2. Geologic framework

The Awibengkok geothermal field is located along the axis of the Sunda Volcanic Arc, which
extends from Sumatra to Flores (Hamilton, 1979; Hutchinson, 1989, Fig. 1a). This arc-trench
304 J. Stimac et al. / Geothermics 37 (2008) 300–331

system marks the convergent boundary between the Eurasian Plate to the north and the Indo-
Australian Plate to the south. The orientation of plate convergence for Java is approximately N-S
(McCaffrey, 1991), making subduction nearly perpendicular to the arc front in central Java, but
increasingly oblique towards Sumatra. The convergence rate for Java has been estimated to be
6–7 cm/year (Tregoning et al., 1994). Regional stress indicators (earthquake focal mechanisms,
borehole breakouts, mapped or inferred surface faults and folds) reveal that in western Java the
maximum horizontal stress is directed approximately north parallel to plate convergence (Shemeta,
1994).
The age of the down-going Indo-Australian Plate is on the order of 85 Ma offshore of western
Java. This portion of the Sunda Arc is also thought to mark the furthest extent of terrigenous
clastic material derived from the Himalayan orogen, whereas pelagic sediment of relatively con-
stant thickness is believed to blanket the oceanic plate south of the trench (Hamilton, 1979;
Handley, 2006). The southeast margin of Sundaland, a block of the Eurasian Plate with pre-Tertiary
basement, is thought to lie somewhere between the Sunda Strait and central Java. According to
Hamilton (1979), the southeast limit of Cretaceous continental crust lies near the Salak-Pelabuhan
Ratu (SPR) Fault, and thus lies just southeast of the geothermal area (Fig. 3a). However, recent
isotopic studies by Handley et al. (2006) indicate that pre-Tertiary basement extends at least to
Merapi volcano in central Java. Given this tectonic framework, the prevalence of commercial
geothermal systems in western Java may be related to the presence of older and thicker crust and
the large supply of terrigenous clastic material to the subduction zone. These factors have pro-
moted the development in western Java of more extensive silicic magmatism, shallow intrusion
and formation of related geothermal systems since the Miocene.
The Awibengkok geothermal system is located in a mountainous area ranging from about 950
to 1500 m asl elevation. The highest peaks are the inactive andesitic volcanoes of Gunung Salak,
Gagak, Perbakti and Endut that lie along the main trend of the Sunda Volcanic Arc. These peaks
border the development site on the east, northwest, southeast, and south sides, respectively, and
give way to lower hill country on the north and south (at 600–950 m asl). The Cianten Caldera, a
collapsed andesitic stratocone lies further to the west with a floor at ∼850 to 950 m asl (Fig. 3a).
Western Awibengkok and the Cianten Caldera are drained by rivers that flow to the north via a
gap in the northeastern part of the caldera rim.
The surficial geology of the Awibengkok contract area was mapped in reconnaissance fashion
by UGI geologists and consultants, with heavy reliance on aerial photographs because of the dense
tropical forests and rough terrain. More detailed mapping of the production area and its immediate
margins has been conducted as construction activities allowed better access and exposure (Stimac
and Sugiaman, 2000; Stimac, 2003). Based on unpublished K–Ar and 40 Ar/39 Ar dating, the major
peaks of the Awibengkok area were built from 860 to 180 ka, whereas the ancestral andesitic cone
that forms the rim of the Cianten Caldera to the west was active from about 1610 to 670 ka (Fig. 3a
and b).
Within the Awibengkok production area, andesitic to rhyodacitic tuffs and lavas dated from
185 to 280 ka have partially filled collapse scars in remnants of the Kiaraberes cone and flowed
downslope to the west, southwest, and north. These rocks are overlain by rhyolitic domes, lavas
and related tephra sequences. Both sequences were erupted primarily along a NNE-trending fault
that crosses the eastern portion of the field. This vent trend is similar to the inferred maximum
horizontal stress and the dominant trend of major fractures measured in borehole image logs
discussed below. The age of this rhyolitic volcanism is from ∼120 to 40 ka based on K–Ar and
40 Ar/39 Ar dating. Carbon-14 dating of organic material in lahar deposits that are stratigraphically

higher than this unit confirms it is older than 40,000 years B.P.
J. Stimac et al. / Geothermics 37 (2008) 300–331 305

The uppermost silicic unit is an extensive tephra known as the “Orange Tuff”. The age of
this unit is bracketed between 40,000 and 8400 years B.P. by 14 C dates on underlying lahar and
overlying hydrothermal breccia units (Fig. 3b). This tuff, of presumed rhyolitic composition,
closely mantles topography in most of the Awibengkok development area, attesting to its young

Fig. 3. Surface geology of the Awibengkok area. (a) Geologic map with major rock types, prominent faults, and altered
ground. Reservoir boundary, well locations, and Sarimaya chloride spring shown for reference. Ages shown in thousands
of years (ka). (b) Ages of volcanic deposits based on K–Ar, 40 Ar/39 Ar, and 14 C dating. Current geothermal activity appears
to be localized near volcanic centers less than 410,000 years old.
306 J. Stimac et al. / Geothermics 37 (2008) 300–331

Fig. 3. (Continued ).
J. Stimac et al. / Geothermics 37 (2008) 300–331 307

age. Based on maps of thickness and maximum pumice and lithic size, its vent is between the Awi
1 and 14 drilling pads along the trend of the young eastern vents described above (Stimac and
Sugiaman, 2000); its actual vent, however, is obscured by later hydrothermal eruption deposits
(Fig. 3a).
Hydrothermal breccias that overlie the Orange Tuff are up to 10 m thick south of Kawah
Cibeureum, and 4 m thick near Awi 2 (Fig. 3a and b). Based on maps of the thickness and clast-
size of the hydrothermal breccia, its vent lies approximately in the same area as the inferred vent
of the Orange Tuff. An older phreatic breccia deposit underlying the Orange Tuff near Awi 14
suggests that vigorous hydrothermal activity in this area also preceded deposition of that unit.
Such activity continued through recent times including small historic phreatic eruptions in this
same area. Some fumaroles in the Cibeureum complex have become more active as a result of
commercial exploitation of the geothermal reservoir.
Recent mapping and 40 Ar/39 Ar dating of dacite lavas thought to be erupted after the Cianten
Caldera collapse suggest that the event occurred prior to 318 ± 14 ka; K–Ar ages of andesitic lavas
forming the eroded caldera rim range from about 0.67 to 1.61 Ma. The caldera was subsequently
filled with volcanic and sedimentary rocks, and lahars near the top of this sequence are as young
as 37–40 ka based on 14 C dating. These sediments are overlain by the Lower Brown and Orange
Tuffs (Upper Tuff sequence) present at Awibengkok (Fig. 3b).
Geochemical studies of volcanic rocks from Gunung Salak and associated vents suggest that
two distinct differentiation trends occur due to variations in shallow-level magmatic processes
(Handley, 2006; Handley et al., in preparation). Handley’s study shows that Awibengkok lavas
are dominantly andesitic, have MgO contents mostly <4 wt.%, moderate alkalinity, and divergent
trends in TiO2 and P2 O5 . The isotopic compositions of all samples plot within the broader array
of subduction-related volcanism on Java. However, one group of samples from the main cone of
Gunung Salak has elevated TiO2 , P2 O5 , high field strength elements (Ti, V, Zr, Nb, Hf and Ta)
and heavy rare earth element (Yb, Lu) concentrations compared to most other Javanese volca-
noes. Handley (2006) suggests that these trends result from combined fractional crystallization
and crustal assimilation by mantle-derived arc parent magmas. Handley’s petrologic modeling
suggests that the crustal assimilant has low K/Rb and Ba/Th ratios. No petrogenetic studies have
been done on the dacitic to rhyolitic lavas that have dominated local volcanism for the last 280 ka.

1.3. Surface manifestations of the geothermal system

Fossil and active hydrothermal systems are associated with several of the volcanic centers in the
Gunung Salak contract area (Fig. 2). The most prominent thermal manifestations are fumaroles
and hot springs directly related to the Awibengkok geothermal system and fumaroles of the Ratu
prospect on the upper western flank of the Gunung Salak volcano (Figs. 2 and 3a and Table 1).
Tertiary gold deposits dated at about 2 Ma occur west of Awibengkok at Pongkor, and about 40 km
farther west at Cirotan and Ciawitali (Milesi et al., 1999; Syafrizal et al., 2005).
Thermal features associated with Awibengkok define a classical distribution seen in many
other commercially exploited liquid-dominated geothermal resources (Rohrs, 1986). Fumaroles
and acid sulfate springs lie above the system at elevations >1050 m asl, whereas bicarbonate and
mixed bicarbonate–chloride and chloride springs lie at progressively lower elevations on the flanks
of the proven reservoir (Fig. 2 and Table 1). The Parabakti fumarole, located near Awi 3 (Fig. 3a),
has a gas composition that indicates proximity to a high-temperature geothermal source, whereas
fumaroles in the east and southeast have gas compositions more consistent with lower temperature
geothermal outflows.
308
Table 1
Summary of surface thermal features at the Awibengkok geothermal system, Indonesia
Feature X UTM (m) Y UTM (m) Elev. Type T (◦ C) pH Cl SO4 HCO3 R/Ra NCG (wt.% T Geotherm.
(m asl) in steam) (◦ C)

J. Stimac et al. / Geothermics 37 (2008) 300–331


Ratu 689,924 9,257,300 1400 F 130 – – – – 7.6–7.7 12.6 273a
Parabakti 683,583 9,256,604 1050 F 126–130 – – – – 7.6–8.0 5–7 265a
Cibeureum 685,000 9,253,795 1230 F 94–98 – – – – 7.22–7.68 2.91 250a
Cipamatutan 686,142 9,252,674 1230 F 89–96 – – – – 7.59 2.38 235a
Awi Baratb 683,373 9,255,042 1060 F/HS-B 73 7.96 2.34 5.26 666 7.3–7.4 – –
Sarimaya 683,077 9,259,270 660 HS-Cl 51–64 6.4 3120–3684 0–5 510–516 7.2–7.4 – 166 (Qtz)
Ciherang 681,000 9,252,135 1230 HS-B 53 6.5 1–2 0.5 690–735 6.72 197 (NKC)
Cikarang 683,750 9,258,050 820 HS-B 72–78 6.5–6.8 1–4 2–83 360–1520 6.8 – –
Muhinin 683,750 9,258,550 730 HS-B 41–54 6.1 3–8 1–11 604–852 ND – –
Cikuluwung 684.350 9,260,100 770 HS-B – 6.7 320 3 1130 8.0 – –
Cisketi 679,050 9,255,250 780 HS-B 40–42 6.1–6.7 19.4 <1–6 254–298 6.8–7.0 – –
Cihideung 690,453 9,264,000 HS-B 43–46 6.7 580 7 550 7.43 – –
Ciseupan 686,280 9,251,100 970 HS-B 49 6.5 <1 7.5 544 – – –
Cibodas 682,750 9,252,500 1030 HS-B 60 6.6 4 1 290 – – –
Ratu 689,924 9,257,300 1400 HS-AS 95 1.5–3.5 1.1–29 160–4520 – 7.64 7.7 –

Notes: Elev. (m asl): elevation (m above sea level); T: temperature; NCG: non-condensable gas; T Geotherm.: geothermometric temperature; HS: hot spring; F: fumarole; B:
bicarbonate; Cl: chloride; AS: acid-sulfate pools; Qtz: quartz geothermometer; NKC: Na–K–Ca geothermometer.
a A variety of gas geothermometers were used to estimate fumarole source temperatures.
b Awi Barat has the chemistry of fumarole steam condensed by near-surface water.
J. Stimac et al. / Geothermics 37 (2008) 300–331 309

Bicarbonate springs are found near the flanks of the proven field at 850–730 m asl. The Sarimaya
hot spring has the highest chloride concentration and is located several kilometers to the north
of the Awibengkok geothermal reservoir at about 650 m elevation. It is thought to originate from
one of the main outflow paths of the system. The chloride–bicarbonate Cikuluwung hot spring
is found about 2 km east of Sarimaya at 770 m asl and may represent strongly diluted outflow
from either the Awibengkok or Ratu geothermal systems. The chloride–bicarbonate Cihideung
hot spring, located about 6 km NE of Cikuluwung is thought to discharge strongly diluted outflow
from the Ratu hydrothermal system, based on its location and stable isotope data (Rohrs, 1986).

2. Reservoir characteristics and conceptual model

The conceptual model of the Awibengkok reservoir is based on integration of a wide range of
geoscientific and reservoir engineering studies that began in the 1980s when the field was first
explored (Mosby, 1984; Noor et al., 1992; Murray et al., 1995; Ganefianto and Shemeta, 1996;
Soeparjadi et al., 1998; Stimac and Sugiaman, 2000). In this section we summarize some of the
key data that underpin this model.

2.1. Reservoir stratigraphy and petrophysics

Subsurface lithologies have been inferred from descriptions of drill cuttings from 81 deep
wells, supplemented by spot cores from 13 wells and 1067 m of continuous core from well Awi
1-2 (Anderson, 1996; Hulen and Anderson, 1998). Borehole resistivity and gamma-ray logs also
provided constraints on major rock types and compositions, respectively, and are the only means
of obtaining lithologic data in sections of wells that were drilled with total circulation losses. The
reservoir rocks at Awibengkok are composed mainly of andesitic and lesser basaltic lava, breccia,
tuff and lahar that comprise several long-lived volcanic centers underlying the southwestern
margin of Gunung Salak (Hulen et al., 2000; Stimac and Sugiaman, 2000). Thick rhyodacitic-to-
dacitic ash-flow tuffs and associated domes, breccias and lahars are interspersed with the more
voluminous andesitic to basaltic sequences that date back to Miocene times (Fig. 4).
The stratigraphic section can be divided into four major formations1 thought to represent
discrete episodes in the evolution of the western Java segment of the Sunda Volcanic Arc (Fig. 4).
The oldest rocks are mainly shallow-marine carbonates and epiclastic sediments (mudstones and
sandstones containing abundant volcanic ash and lithic debris). Some of these units appear to
be waterlain tuffs and ash-laden turbidites. These marine sedimentary rocks are very different
from the overlying and interbedded volcanics by virtue of their composition, lower porosity and
permeability, and less-intense hydrothermal alteration. Fossil assemblages in the carbonates are
characteristic of Early-to-Late Miocene shallow-shelf environments. Core and cuttings from Awi
17-1 has shown that these basement rocks have locally been converted to high-grade hornfels
and skarn assemblages in the contact metamorphic aureole of subvolcanic intrusions within the
Cianten Caldera and western Awibengkok reservoir at depths of about 1–3 km. Elsewhere weak
metamorphism associated with shallow intrusion of the Miocence marine sedimentary rocks has
yielded local neoblastic recrystallization in carbonate rocks and enhanced parting or foliation in
mudstones.

1 “Formation” is used to refer to distinctive time bounded rock sequences that have not been formally correlated with

recognized formations on a regional basis.


310 J. Stimac et al. / Geothermics 37 (2008) 300–331

Each major volcanic formation is further subdivided into a lower andesitic section and an
overlying rhyolitic or dacitic section that are inferred to represent distinct or partially overlap-
ping volcanic episodes that became more silicic with time (Fig. 4a). Thinly bedded to massive,
clay-rich ash and lapilli tuffs are present throughout the sequence. Groups of these important
marker beds have been found in some wells, and some have been correlated from well to
well.
The lower volcanic formation, consisting of andesitic to basaltic volcanic rocks, overlies
and is interbedded with the Miocene sedimentary section. These volcanic rocks probably rep-
resent a major episode of calc-alkaline magmatism in the immediate area, and the transition

Fig. 4. (a) Subsurface geology of Awibengkok. (a) Stratigraphic column for the Awibengkok reservoir. (b) Lithologic
cross-section A–A (see Fig. 2) showing the distribution of formations and major faults in the Awigengkok area.
J. Stimac et al. / Geothermics 37 (2008) 300–331 311

Fig. 4. (Continued ).

from marine to subaerial conditions. The lower part of this formation consists mainly of lavas
and breccias (hyaloclastites are common) interbedded with fossiliferous marine carbonates. The
hyaloclastite deposits consist of glassy-to-crystalline lava fragments with abundant spherical vesi-
cles, now completely filled by chlorite and other secondary phases. This sequence is overlain by
andesitic to dacitic lavas, tuffs and breccias that appear to have been deposited in a subaerial
environment.
The second major formation is a widespread sequence of silicic tuffs known as the Rhyodacite
Marker. This unit is interpreted by us to represent an episode of silicic volcanism and caldera
formation that followed the first major episode of andesitic stratovolcano construction in the
area. At least three thick rhyolitic-to-rhyodacitic units are present in most wells, separated by
thinner interbeds of dacitic and andesitic tuffs. These units are recognized as silicic by virtue of
their well-preserved devitrification textures, phenocrysts of chloritized biotite, and large embayed
quartz “eyes”. Together, the rock packages described above comprise the bulk of the western
Awibengkok reservoir.
The middle volcanic formation is another sequence of andesitic-to-dacitic lavas, tuffs, and
lahars and debris flows that represents construction, collapse and erosion of stratovolcanoes
and lava and dome complexes. This subaerial sequence dominates the Awibengkok geothermal
reservoir in the eastern portion of the field. A more detailed knowledge of the stratigraphy and
petrophysical properties of the shallow eastern reservoir was acquired from continuous core taken
in well Awi 1-2 (Hulen and Anderson, 1998). The rock sequence in this well can be divided into
several distinct packages (Hulen et al., 2000). The uppermost one is primarily dacitic in compo-
sition, consisting mainly of lavas, domes and related breccias, tuffs and lahars. The underlying
package is andesitic in composition, consisting dominantly of lavas and tuffs in its lower part, and
312 J. Stimac et al. / Geothermics 37 (2008) 300–331

lahars in its upper part. Coarse-grained epiclastic sediments dominate the upper portions of both
packages, suggesting a decline of volcanism and erosion of edifices, or fault uplift and down-
cutting of adjacent areas. The upper volcanic formation consists of another andesitic sequence
overlain by dacitic to rhyolitic rocks that includes the surface deposits described above.
Intrusive rocks have been identified in many wells including Awi 17, 18, 12, 9, 4, 2 and 1
(Fig. 2). A significant swarm of silicic-to-intermediate composition dikes and sills containing
plagioclase, alkali feldspar, pyroxene, titanite and biotite or hornblende occur at shallow lev-
els in the Cianten Caldera to the west of the proven reservoir and along the eastern caldera
wall.
A thick hypabyssal quartz-dacite porphyry also intrudes the eastern caldera ring fault. A smaller
intrusion in well Awi 1-2 was described as a quartz diorite sill (Hulen and Lutz, 1999; Hulen et al.,
2000). It is generally less altered than the surrounding rocks, due to its low inherent permeability.
An intrusion in Awi 2-1 is present in core taken near the bottom of the well at 2131 m vertical
depth (912 m below sea level). It is a porphyritic silicic rock consisting mainly of feldspar and
quartz. Based on the relatively coarse grain size of this intrusion, it may represent a dike or the
top of a larger granitic body with a texture similar to the microgranite porphyry carapace of the
Geysers Felsite (Hulen et al., 1997). A coarse-grained granitic fragment about 1 m across was
collected from tuffs associated with the upper rhyolite (Fig. 4). Hornblende and biotite separated
from the granite yielded concordant 40 Ar/39 Ar ages of about 113 ka, which are thought to reflect
complete “thermal resetting” and therefore, the age of the eruption.
The intrusion in Awi 4-1 consists of a devitrified-to-glassy rhyolite with sparse phenocrysts
of quartz, plagioclase, pyroxene, hornblende, and biotite. It probably represents a feeder to the
young rhyolitic dome or related lavas (Fig. 4b). The intrusion is weakly altered, suggesting that
it was impermeable to reservoir fluids, or that cooler fluids were moving downward along its
margin.
Relatively coarse-grained holocrystalline textures in cuttings samples from the deepest sections
of Awi 9, 11 and 12 wells are also consistent with the presence of numerous dikes and sills in
this area. Some Miocene carbonate rocks in these wells also show neoblastic textures that are in
agreement with contact metamorphism.
The Awi 1-2 core also clarified the importance of: (1) lahars (Anderson, 1996; Hulen and
Anderson, 1998); (2) the nature of clay-rich layers that had been called “paleosols” (Hulen and
Lutz, 1999; Hulen et al., 2000); and (3) the abundance of hydrothermal breccias (Hulen et al.,
1999) in this part of the Awibengkok reservoir. Facies analysis of the Awi 1-2 core also suggests
that epiclastic deposits such as lahars typically mark the end of each episode (Hulen et al., 2000).
These very coarse-grained, non-bedded and matrix-supported units are relatively easy to identify
in core or image logs, but are much more challenging to be recognized in cuttings. Lahars are
indicated in cuttings by a variety of rock types and chips containing matrix–clast contacts, and in
gamma-ray logs by high variability in count rate.
Petrographic and X-ray diffraction studies of “paleosols” revealed that they are fallout and ash-
flow tuffs that were converted to kaolinite and smectite or smectite-rich mixed layer clay in the
shallow subsurface. Upon deeper burial, these weak formations compacted to such an extent that
they were rendered impermeable and did not re-equilibrate at the higher temperatures to which
they were subjected. Their low strength, high ductility and chemical reactivity with water limit
their ability to maintain open vertical fractures when ruptured and therefore behave as ductile seals
to reservoir fluid circulation. Numerous hairline fractures in these units are sealed by pyrophyllite
that is otherwise rare-to-absent in the alteration assemblage. Clays in these altered tuffs swell on
contact with water and cause severe sloughing problems during drilling. Drilling practices such as
J. Stimac et al. / Geothermics 37 (2008) 300–331 313

rapidly casing off the most problematic formations, selecting bits with high penetration rates and
drilling with aerated polymer-based mud to reduce circulation losses were developed to minimize
their negative impacts on well-drilling costs.
Two prominent tuff layers consisting mainly of fine ash were encountered in Awi 1-2 from 777
to 788 m and from 1142 to 1162 m vertical depths.

2.2. Porosity and permeability trends

Like most volcanic-hosted geothermal systems, permeability at Awibengkok is controlled


mainly by the presence of a highly interconnected network of open to partially sealed frac-
tures. Although of secondary importance to well permeability, the amount of fluid stored within
microfractures and pores is a significant factor in overall reservoir performance and how much
heat can be extracted commercially from the subsurface. To characterize the matrix porosity of
the Awibengkok reservoir, helium porosity measurements were conducted on 87 core samples
from wells distributed throughout the field, but most of the data were obtained from the Awi 1-2
continuous core (Anderson, 1996; Hulen and Anderson, 1998).
Additional estimates of porosity with depth were obtained from density and sonic velocity
logs run in wells where core was available for calibration of the subsurface profile. All these
data support a trend of declining matrix porosity with depth (Fig. 5a), similar to other fields
such as Tiwi (Stimac et al., 2004). An average porosity of about 10.6% was observed in all
core samples. Porosity trends as a function of rock type are similar to other volcanic-hosted
geothermal systems, where dense crystalline rocks such as lava flow and dome interiors and
intrusions have the lowest values (average of 8.6%), and fragmental rocks such as breccias, tuffs
and lahars have higher (average of 13.1%) but more variable porosities that depend on the extent
of mechanical and chemical change the rocks have undergone (Stimac et al., 2004). The trend
of declining porosity with depth is driven primarily by declining porosity in the fragmental rock
types. Porosity data on deep sedimentary basement rocks are sparse, but petrographic examination
of these rocks confirms that they typically lack both open microfractures and significant matrix
porosity, and thereby commonly control the lower limit of the commercial permeability in the
field.
Matrix permeability to air ranges from <0.001 to 5.6 md, with a geometric mean of 0.026 md.
A poor correlation exists between matrix porosity and matrix permeability, in part because per-
meability at the scale of core plugs is mainly related to vuggy, partially sealed microfractures
(Fig. 5a). However, as the matrix porosity of the sample becomes higher, a trend of minimum
permeability for a given porosity is defined by the dashed line in Fig. 5b. In other words, very
few samples from geothermal reservoirs plot below this line because the high volume of pores
geometrically necessitates a degree of connectivity.
Twenty-six core samples from Awi 1-2 were analyzed by mercury injection capillary porosime-
try (MICP) at Chevron Energy Technology Company. Matrix porosity determined by MICP
averages about 1 porosity unit lower than conventional core porosity. The average modal pore
radius was 0.033 ␮m with a range from 0.005 to 0.1 ␮m. The median permeability calculated
from MICP is 0.0038 md as compared to 0.021 md for permeability to helium gas on the same
samples. One reason the MICP permeability is lower than the convectional permeability to air is
that the MICP method does not measure permeability associated with partially sealed fractures,
vugs, microchannels and hairline fractures if they fill at low pressure. An important conclusion
from this data is that given the low-measured matrix permeability and pore throat aperture of these
samples, the rock matrix is more likely to produce steam than liquid water as pressure declines.
314 J. Stimac et al. / Geothermics 37 (2008) 300–331

Fig. 5. Matrix porosity and permeability trends in the Awibengkok reservoir. (a) Matrix porosity to helium gas versus
vertical depth. A best-fit linear regression through the data indicates a decline in porosity with depth. (b) Matrix permeability
to air versus matrix porosity to helium gas. The dashed line marks an increasing minimum in matrix permeability as a
function of increasing matrix porosity.

2.3. Reservoir structure and controls on permeability

Photolineaments observed on satellite images and aerial photographs trend mainly N-to-NE,
NW and E-W in Awibengkok and the surrounding region (Fig. 3a). Mapping of faults and fractures
has proved to be difficult due to heavy tropical vegetation, but some orientations and indicators
of direction of fault slip have been documented (Fig. 6a). Both fault and fracture orientations are
predominantly N-to-NE, with subsidiary NW and E-W trends. The sense of slip on faults based
on rare slickensides varies from purely dip–slip (normal displacement) to strike–slip movement.
The preponderance of N-to-NE and NW-trending structural features that have been mapped
is consistent with the local stress field. The reservoir stress regime has been evaluated based on
well leak-off tests, downhole image and density logs, and apparent active fault offsets (Sugiaman,
2003). The local maximum principal stress is vertical, whereas the maximum horizontal stress
is oriented NNE. The ratio of minimum to maximum horizontal stress shows a 2:1 anisotropy.
Microseismic focal mechanism and fault-offset data also indicate an extensional state of stress,
consistent with the observation of abundant open fractures in the reservoir.
J. Stimac et al. / Geothermics 37 (2008) 300–331 315

Fig. 6. Locations and orientations of surface and subsurface structures in the Awigengkok area. (a) Rose diagrams of
strike of surface faults mapped in the Cianten Caldera and the Awibengkok area (black) and open fracture orientations
interpreted from downhole image logs (dark gray). N refers to the number of open fractures mapped. Abbreviations for
image log type are BHTV: Borehole Televiewer; FMS: Schlumberger Formation Microanalyzer; FMI: Schlumberger
Formation Microimager; XRMI: Halliburton Extended Range Microimager. (b) Map of inferred subsurface faults based
on offset in the Rhyodacite Marker (RDM).
316 J. Stimac et al. / Geothermics 37 (2008) 300–331

The distribution and focal mechanisms based on microseismicity in the eastern part of the
Awibengkok field show NW-trending, right-lateral, strike–slip fault zones that concentrate sec-
ondary NNE-trending active normal faults (Glen Melosh, personal communication, July 2005).
The inferred local stress orientation differs slightly from the regional interpretation. Regional data
suggest that the maximum horizontal stress is towards the north (Shemeta, 1994) as compared to
N20◦ E at Awibengkok (Sugiaman, 2003).

2.3.1. Interpreted reservoir structures


Structural interpretations that combine surface mapping and subsurface offsets on the Rhyo-
dacite Marker and basement rocks serve as the main basis for interpretation of reservoir structure.
A structural map of the field shows a fault pattern that includes roughly N-S-, NE-, E-W-, and
NW-trending structures (Fig. 6b). The wide variety of evidences used to make this interpretation
makes it likely that both active and inactive faults are represented.
The permeability of faults is not known directly, however, it is clear from integration of multiple
datasets (well deliverability, interference tests and pressure trends, fluid geochemistry, tracer return
patterns) that several faults play important roles as flow barriers and/or conduits.
It appears that near-vertical NNE-trending faults tend to partition the system into subcompart-
ments that have slightly different fluid chemistry and temperature. For example, the NNE-trending
Awi fault, which controls the locations of recent volcanic vents, plays an important role in seg-
menting the system, and allowing the descent of shallow fluids into the eastern portion of the
geothermal reservoir. Similarly, the NNE-striking Muara fault is coincident with the eastern mar-
gin of the Cianten Caldera and appears to limit the extent of the commercial permeability and
temperature on the west (Fig. 3a). Subcompartments also show distinct pressure and temperature
histories, especially as measured in shallow steam wells. Steam cap pressures are progressively
lower from the north-central reservoir (Awi 7, 3, 16 pads) toward the Cibeureum fumarole com-
plex to the southeast (Awi 13, 1, 2 pads; see Fig. 2). As noted in Section 1.2, this fumarole complex
is the most likely vent location of the Orange Tuff <40,000 years ago (Fig. 3a).
Image log fracture patterns and return patterns of chemical tracers both indicate preferential
NE-SW fluid flow due to distributed small fractures that are open along this trend, enhanced
permeability along the strike of optimally oriented major faults, or both (Fig. 6a). Initial-state
temperatures and fluid chemical trends suggest that intersections of N to NE- and NW- or E-
W trending faults facilitate fluid migration between adjacent compartments, especially along the
southern and northern margins of the proven reservoir; however, these flowpaths were not validated
by tracer returns. This may be because these faults represent sources of external recharge or, have
relatively lower permeability than N-to-NE structures due to their unfavorable orientation within
the present stress regime.
There is also evidence based on entry distributions in wells that certain stratigraphic units,
such as the Rhyodacite Marker, and their contacts act as fractured aquifers. These units appear to
channel fluids laterally, and link to major fault-controlled pathways. Another prominent cluster
of fluid entries is associated with the contact region between the lower volcanic formation and
the underlying marine sedimentary package. This may be related to fractures developed in or
enhanced at the top of the eroded basement rocks.

2.4. Hydrothermal alteration and paragenesis

Hydrothermal alteration has produced argillic, propylitic, and phyllic alteration zones; no
advanced argillic alteration has been identified within the high-temperature reservoir (Fig. 7).
J. Stimac et al. / Geothermics 37 (2008) 300–331 317

Fig. 7. Cross-section A-A (see Fig. 2) showing the distribution of alteration mineral assemblages. Faults may play a role
in localizing changes in the elevation of the reservoir top. HT-clay: hydrothermal clay (rich in smectite and associated
with pyrite).

The system is capped by a zone of intense argillic alteration that is dominated by smectite,
with accessory pyrite, hematite, calcite, anhydrite and zeolites that formed at temperatures less
than 180 ◦ C (e.g. Henley and Ellis, 1983). The depth interval corresponding to this alteration
typically has resistivities below 10 -m and a conductive temperature profile indicating very low
permeability. It has been successfully mapped by methylene blue and X-ray diffraction analyses
of cuttings in selected wells (Gunderson et al., 2000).
Hydrothermal alteration shows a transition from argillic to propylitic assemblages with increas-
ing depth. This transition zone can span up to 300 m in thickness, and is dominated by mixed
layer smectite–illite, chlorite, calcite, pyrite, titanite and quartz. In well Awi 1-2, the bottom of
this transitional assemblage is marked by a 20-m thick andesitic fallout tuff that is massively
altered to mixed layer illite–smectite and chlorite–smectite as described above; Hulen and Lutz
(1999) classified this assemblage as argillic–phyllic. Because of its high smectite content, the tuff
appears to have been altered at much lower temperature than measured today. This clay-altered
tuff unit forms an effective seal to the underlying propylitically altered reservoir section. In gen-
eral, the transition zone has a conductive-to-weakly convective temperature gradient indicating
low-to-moderate permeability.
The propylitic alteration zone is dominated by epidote, illite, quartz, and chlorite, but also
contains albite, adularia, calcite, wairakite, pyrite, anhydrite, and titanite (Hulen et al., 2000).
This alteration assemblage corresponds to reservoir temperatures between 240 and 270 ◦ C (Henley
and Ellis, 1983). Rare occurrences of garnet, prehnite, biotite, and amphibole occur at deep levels
318 J. Stimac et al. / Geothermics 37 (2008) 300–331

indicating higher temperatures of deposition near the igneous intrusions (>280 ◦ C for amphibole;
300 ◦ C for biotite; Henley and Ellis, 1983). Phyllic alteration consisting mainly of quartz, sericite,
and pyrite with lesser anhydrite, chlorite, adularia, and epidote is common in the more silicic
rocks (dacites and rhyolites). This alteration assemblage corresponds to reservoir temperatures of
260–290 ◦ C in other fields (Henley and Ellis, 1983).
Paragenetic relationships in Awi 1-2 indicate that the earliest secondary minerals formed were
calcite and anhydrite, probably precipitated from formation or steam-heated waters in response
to heating (Moore and Norman, 1999). Early veins of distinctive Fe-rich calcite ± hematite
are cut by later calcite and epidote veins in Awi 18-1, but in general this early assemblage
appears to be restricted to wells in the shallow steam cap or on the reservoir margins. Most
fractures throughout the deeper high-temperature reservoir were initially lined with acicular epi-
dote ± pyrite, followed by various combinations of quartz, calcite, adularia, pyrite, and sericite
with local traces of actinolite, prehnite and grossular garnet. Hulen et al. (2000) documented pre-
cipitation of calcite–wairakite veins in newly created fractures and remaining voids, followed by
quartz-epidote-sulfide veins and vug-linings. Another generation of wairakite was then deposited,
followed by quartz-epidote veins, and finally the coarse, blocky, rhombic calcite.
This rather complicated paragenesis suggests multiple cycles of fracturing and sealing in the
shallow eastern reservoir, possibly related to its proximity to young fault-mediated intrusion and
volcanism as described in Section 1.2. Wells near the margins of the commercial reservoir (e.g.
Awi 18) commonly have late vein fillings dominated by calcite, anhydrite, or zeolite that signal
ingress of cooler fluids, and sealing of the reservoir margin. Sealed calcite veins are also locally
abundant in the Miocene sedimentary basement in all deep wells, some of which are cut by
stylolites and therefore probably predate recent hydrothermal activity.
Fluid-inclusion evidence from well Awi 1-2 indicates that some of the early calcite was precip-
itated at about 270 ◦ C from highly saline (16–18 equivalent wt.% NaCl) fluids, possibly magmatic
in origin and derived from the quartz diorite sill (Moore and Norman, 1999). Later inclusions
in quartz, epidote and wairakite have lower salinities, up to 3.5 equivalent wt.% NaCl, more
typical of the modern hydrothermal fluid regime; they exhibit a range of temperatures and gas
ratios indicating mixing of geothermal and steam-heated meteoric waters. Hydrothermal breccias
are typically cemented by various combinations of quartz, wairakite, epidote, chlorite, pyrite,
and other minor phases (Hulen et al., 1999). Wairakite is particularly abundant in these breccia
cements and appears to be largely responsible for sealing their permeability.

2.5. Reservoir temperature and fluid geochemistry

Awibengkok is a liquid-dominated geothermal system with benign fluid chemistry. At initial


conditions, fluid salinity was approximately 1.3 wt.% and non-condensable gas content (NCG)
was <0.4 wt.% except for the shallowest zones (Molling and Rohrs, 1996; Stimac and Sugiaman,
2000).
Prior to exploitation, the geothermal system was on a pressure gradient that would support
a liquid column to an elevation of about 800 m asl, but it was underpressured with respect to a
liquid column from the ground surface of eastern wells (Awi 1, 13, 16, and 4) at 1300 m asl;
Acuña et al. (this issue) show the initial-state pressure gradient of the geothermal reservoir (see
their Fig. 2). Reservoir temperatures were in the single-phase liquid field throughout most of the
reservoir but reached two-phase conditions at about 560 m asl in the Awi 1 pad area. This resulted
in the development of a relatively gas-rich steam cap in the shallowest part of the reservoir once
exploitation commenced.
J. Stimac et al. / Geothermics 37 (2008) 300–331 319

Fig. 8. Sector map of the Awibengkok geothermal reservoir. The field can be divided into four sectors with distinctive
temperatures, fluid chemical signatures, and tracer return patterns. Faults and fractures shown in Fig. 6 play a major role
in determining sector boundaries. Arrows indicate the general direction of fluid flow under natural-state (pre-exploitation)
conditions.

Subtle variations in fluid chemistry allow the field to be subdivided into four distinct sectors
or “cells” that are likely bounded by faults. There is clear evidence that dikes and sills have
intruded into some fault zones during recent episodes of volcanism and may have either sealed
existing permeability or created new pathways near the boundaries of these cells. The four sectors
have been termed the Western, Central, Eastern and Far Eastern Cells (Fig. 8). In addition, some
wells located on the periphery of the proven reservoir (e.g. Awi 12, Awi 10-2OH) have distinct
chemical signatures of “reservoir margin” fluid that is generally more variable in Cl and higher in
Mg. Some chemical characteristics of the fluids produced from these sectors and the field margins
were summarized in Stimac and Sugiaman (2000).
The Western Cell, which includes the Awi 9 wells, had the highest initial measured temperatures
(290–312 ◦ C), while NaKCa and quartz geothermometers yielded temperatures up to 316 and
280 ◦ C, respectively. The initial chloride concentration was about 6200 ppm, whereas the gas
composition was depleted in H2 , CH4 , N2 , and Ar and enriched in CO2 and H2 S compared to
other reservoir regions, possibly reflecting a stronger magmatic influence. The high temperatures
of this area indicate it is a locus of deep fluid upflow.
The Central Cell is characterized by wells in the area of drilling pads Awi 7, 8, 10 and
11, where initial chloride concentrations ranged between 6500 and 6900 ppm. Initial measured
temperatures ranged from 270 to 280 ◦ C, whereas NaKCa and quartz geothermometers yielded
temperatures of 260 and 280 ◦ C, respectively. Compared to the Western Cell, this cell had higher
320 J. Stimac et al. / Geothermics 37 (2008) 300–331

chloride and lower gas contents. Furthermore, chemical modeling suggests that the Central Cell
gas was initially enriched in H2 S and NH3 and depleted in H2 relative to the other cells. All
these features suggest that the Central Cell has experienced gas loss as a result of long-term
fumarolic emissions and associated boiling. Deep outflow to the north of the Awi 3 pad is sup-
ported by an extensive low-resistivity anomaly and the distribution of surface thermal features
(Fig. 3a).
The Eastern Cell, comprising the shallowest portion of the Awibengkok reservoir includes most
of the wells on the Awi 1, 2, 13, and 16 pads (Fig. 8). This portion of the field was distinguished
by its initial 250 to 260 ◦ C temperatures, except in the far south (Awi 15 wells) where a deeper
reservoir top with initial temperatures from 260 to 277 ◦ C prevailed. The Eastern Cell was initially
characterized by dilution (Cl from 5100 to 6400 ppm) and higher NCG concentrations relative to
the more deeply sourced Central and Western Cells (Stimac and Sugiaman, 2000).
A pattern of dilution, which runs NNE from the Awi 16 to Awi 13 location, suggests in the
Eastern Cell meteoric or steam-heated water is descending along the same structural zone that
localized the most recent NNE-trending volcanic and hydrothermal activity. A mass-balance
calculation reveals that this lower-temperature fluid contains relatively higher Mg concentrations
and significant amounts of SO4 2− , HCO3 − and NH3 . The higher NCG in this cell is a feature
associated with the shallower reservoir top. During the evolution of the reservoir, gas accumulated
in this shallow portion of the system as the result of boiling and condensation. A productive steam
cap formed in this area in response to increased mass withdrawals beginning in late 1997. The
NCG concentration is markedly higher and more variable in the steam cap than in the liquid
reservoir, varying from about 0.5 to 10 wt.% in individual wells. Wells dominated by shallow
fluid entries near the location of cooler downflows are highest in NCG due to enhanced steam
condensation and gas accumulation.
The Far Eastern Cell is defined by wells drilled east of the Awi 1 pad, which cross the prominent
NNE-trending Awi and Cibeureum Faults. These wells (Awi 14, 5, 1-2RD, 1-7, 1-8 and 1-9)
have higher temperatures in their deeper portions than adjacent Eastern Cell (i.e. up to 270 ◦ C).
However, like other field margins this cell shows some influence of deep marginal air-saturated
groundwater. Awi 5 and Awi 14 wells have relatively high initial concentrations of chloride and
lower geothermometer temperatures indicating that boiled fluid flows out of the geothermal system
to the SE.

2.6. Geophysical characteristics of the reservoir and its possible extensions

Geophysical surveys and monitoring data have played an important role in the development
and refinement of the conceptual model used during the exploration, development and production
of the Awibengkok field. During the initial exploration program, detailed comparisons of the MT
resistivities with lithology and clay alteration showed that the low resistivities (<10 -m) corre-
sponded with the smectite-rich clay cap. The reservoir section with higher-temperature propylitic
alteration typically had resistivities above 20 -m (Cumming and Mulyadi, 1984). This interpre-
tation led to exploration drilling west of the thermal areas and helped define the most likely extent
of the Awibengkok geothermal system. More recent MT and TDEM surveys, especially around
the margins of the proven production area, have identified potential reservoir extensions to the
west and north described in more detail below.
Monitoring of microseismicity has been carried out since production began in 1994. This was
initially done with drum recorders, then with a five-station analog telemetry array (April 1995), and
currently with a 10-station digital array (since April 2005). Detailed and focused microseismicity
J. Stimac et al. / Geothermics 37 (2008) 300–331 321

monitoring over geothermal systems has proven effective in defining field margins, determining
the state of stress, and revealing permeability trends (e.g. Stark, 1992; Kato et al., 2000).

2.6.1. Resistivity interpretation


The correlation of low resistivities with smectite and smectite-rich mixed layer clay alteration
at Awibengkok is characteristically observed over volcanic-hosted geothermal systems. Based
on these observations a conceptual model was developed that was subsequently applied to the
interpretation of resistivity data from Awibengkok and other prospect areas (Cumming et al.,
2000). Fig. 9 shows the basic elements of the model and illustrates the correspondence of low
resistivities with the smectite-rich cap. Resolution of the base and top of the low-resistivity section
can be used to map the clay cap and help to infer the location of a commercial geothermal system.
The resistivity structure at Awibengkok, showing the base of the <10 -m resistivity layer, is
illustrated in Fig. 10. Resistivity models at each MT station are based on one-dimensional (1D)
models using the transverse electric (TE) mode (electric field parallel to strike). The map was
produced by determining the elevation of the base of the <10 -m layer from 1D models for each
MT station and then contouring the data. The TE mode has been shown to generally give a better
1D model of the subsurface resistivity structure than the invariant of the transverse magnetic
(TM-electric field perpendicular to strike) and TE modes (Anderson et al., 2000).
There are three important features evident in Fig. 10: (1) correlation of the shallowing of the
base with the location of the proven production area (200 m contour); (2) extension of the relatively
shallow base northwest of Awi 9 into the Cianten Caldera; and (3) extension northward of the Awi
3 area. The margins of these extensions are characterized by a rapid deepening of the base of the
low-resistivity layer. Prolongations of the relatively shallow conductor to the NW and N from the
production area define possible extensions of geothermal conditions. Two cross-sections along
the axes of these shallow conductors are shown in Fig. 11a and b, and compare the distribution
of the low-resistivity anomaly, temperatures, and the clay cap as defined by methylene blue data
(Gunderson et al., 2000).
Along the northern extension, the distribution of the low-resistivity layer is well correlated
with the 175 ◦ C isotherm and the clay-rich zone defined by the methylene blue data (Fig. 11a).
The base of this layer is consistent with the temperature range where smectite and smectite-rich
mixed layer clays are stable, indicating that over proven portions of the field the elevation of the
base of the low-resistivity layer corresponds to the distribution and base of the argillic cap. The
layer bows upward beneath the production area and its base is shallowest beneath Awi 2 and 13
where high temperatures are the shallowest. South of Awi 15 the low-resistivity layer thickens and
the base becomes markedly deeper, consistent with the lower shallow temperatures and the deeper
top of the commercial permeability defined by drilling. This area has the expected geophysical,
thermal and permeability signature of the margin of a productive geothermal reservoir (Fig. 9).
North of Awi 3 the base of the layer is relatively flat until north of the Sarimaya chloride spring,
consistent with geothermal outflow extending that far.
A cross-section along the northwest extension shows a similarly good correlation with the
175 ◦ C isotherm and the clay-rich zone defined by the methylene blue data (Fig. 11b). Near Awi
14, the methylene blue and the low-resistivity match, but temperatures are lower at shallow levels
than at Awi 2. Southeast of Awi 14, the conductive layer thickens and the high temperatures
deepen as expected at a reservoir margin. The lower temperatures near Awi 14 relative to the base
of the low-resistivity layer and clay cap indicate the temperature may have been higher at this
depth in the past. This is common near a reservoir margin where cooling has occurred at shallow
levels without significant re-equilibration of the clay minerals.
322 J. Stimac et al. / Geothermics 37 (2008) 300–331

Fig. 9. Conceptual distribution of temperatures, smectite-rich alteration and electrical resistivities for a typical volcanic
geothermal system. The geometry of the base and top of the low-resistivity layer can provide a good estimate for the
location of the geothermal reservoir. Arrows show the flow of hot fluids through the system.
J. Stimac et al. / Geothermics 37 (2008) 300–331 323

Fig. 10. Elevation of the base of the low-resistivity anomaly in relation to the proven Awibengkok reservoir area is
given in m above sea level (negative values are below sea level). Extensions of the anomaly occur north of Awi 3 and
west-northwest of Awi 9. The location of Sarimaya chloride spring is indicated. Also shown are the locations of the
cross-sections discussed in Fig. 11.

Northwest of Awi 9, the base of the low-resistivity layer shallows to depths that are similar to
those observed in the production area (Fig. 11b), and it is relatively flat across the Cianten Caldera.
Prior to exploratory drilling, it was considered that the conductive layer in this area could be due
either to an extension of the Awibengkok reservoir, a separate hydrothermal system (active or
extinct), or indicate the presence of smectite-rich post-caldera lake sediments. Low-temperature
bicarbonate springs along this NW-SE profile provide evidence that the low resistivities are at
least in part due to a source of hydrothermal fluids. Post-drilling assessment of the anomaly is
discussed in Section 3.

2.6.2. Microseismic data interpretation


At Awibengkok there is a strong spatial correlation between the microseismic activity and the
areas used for waste geothermal brine and condensate injection (Fig. 12). Most microseismic
events are magnitude M2 or less, and are too small to be felt at the surface. During the more
than 13 years of field exploitation, there have been no events large enough to cause any surface
disturbance or damage. Seismicity rates tend to be highest when an injection well is first put on
line but then decrease with time. For new injectors there have been cases where clusters of 60 or
more detectable microseismic events occurred shortly after injection started, often within a single
day. Currently field wide detection rates average about 20–30 local events per month. These rates
324 J. Stimac et al. / Geothermics 37 (2008) 300–331

Fig. 11. Resistivity profiles of the Awibengkok area highlighting the <10 -m layer defined by 1D MT models (shaded
blue) temperature isotherms and methylene blue contours (in g/ml). (a) N-S resistivity cross-section showing the thickening
of the low-resistivity layer south of Awi 15 and deepening of the high-temperature isotherms is consistent with a reservoir
margin. A similar thickening of the low-resistivity layer is observed north of the Sarimaya chloride spring. (b) Pre-drilling
NW-SE resistivity. The thickening of the low-resistivity layer SE of Awi 14 and deepening of the high-temperature
isotherms is consistent with a reservoir margin. The isotherms do not show clear evidence for a margin NW near the Awi
9 pad. (c) Post-drilling NW-SE cross-section with observed temperatures (in red) and methylene blue data (in g/ml) from
new wells Awi 17 and 18.

are typical for other fields monitored by Chevron in Indonesia and the Philippines, as well as in
other geothermal field throughout the world (e.g. Majer et al., 2007).
An east-west cross-section illustrating the characteristics of the observed microseismicity has
been drawn through the western (Awi 9) and eastern (Awi 14 and 15) injection wells (Fig. 12b)
and shows that the events have distinct depth distribution in these two areas. In the east, the
base of the seismicity correlates with the depth of low-permeability marine sedimentary rocks,
whereas in the west the seismicity extends well into the continuous marine sedimentary rocks. The
deep western microseismic events occur below the area of the field with the highest measured
temperatures, a distinct fluid chemistry with a stronger magmatic influence, and petrographic
J. Stimac et al. / Geothermics 37 (2008) 300–331 325

Fig. 12. Microseismicity recorded in the Awibengkok area during the April 2005–April 2007 period. Most of the seismic
events can be correlated with injection into wells Awi 9, 12, 14 and 15. A distinct deepening of the microseismicity
beneath the Awi 9 area is coincident with the highest temperatures and interpreted upflow for the field.

evidence of intrusive rocks and contact metamorphism. All these features have led us to interpret
the Western Cell as a zone of deep fluid convection and upflow related to cooling plutons (Fig. 8).
Deep microseismicity in this area indicates pressure connection between the injection zones at
2–3 km depth and deeper zones at 5 km, most likely along faults and intrusive margins. This
326 J. Stimac et al. / Geothermics 37 (2008) 300–331

insight has important implications for heat exchange and transport in the system and has been
incorporated into the conceptual model of the hydrology and movement of injectate in the reservoir
(Acuña et al., this issue).

2.7. Field conceptual model

Fluid-flow patterns in the Awibengkok geothermal system have been inferred by analysis of
regional and local geological and geophysical data, trends of initial-state well temperatures and
pressures, resistivity and rock alteration patterns, fault and fracture orientations and densities, well
pressure, enthalpy and fluid chemistry histories during production, microseismicity distribution,
and patterns of tracer returns and well interference.
The hot fluid upflow at Awibengkok appears to be controlled by deep intrusion along the
general E-W trend of the Sunda Volcanic Arc. At shallow levels of the geothermal system the
hot fluid upflow is localized along N- to NE-trending faults where subvolcanic intrusions (small
stocks, dikes, sills) may be found closest to the surface. Initial-state well temperature and pressure
data, summarized by Acuña et al. (this issue), show a deep reservoir top in the Western and Central
Cells and a shallow reservoir top in the Eastern Cell (Fig. 8). Based on well temperatures and fluid
geochemistry, it is inferred that deep upflow occurs in the Western Cell and along the western
margin of the Central Cell. Ascending fluids are locally confined by low-permeability clay-rich
tuffs on the west, but move up along one or more geologic structures that breach these barriers
in the central and east-central parts of the Central Cell. Fluids ascend to the shallowest portion of
reservoir in the Eastern Cell and boil under the Cibeureum fumarole complex. Part of the boiled
fluid flows at shallow depths towards Awi 5, a portion moves down into the Awi 14 area, and
some small fraction may ultimately re-circulate to the base of the Central Cell. Another prominent
outflow occurs to the north of Awi 3. Influx of steam-heated meteoric water occurs in the Eastern
Cell, as indicated by dilution of the fluids in that area. Minor influx of, or mixing with, deep
marginal air-saturated meteoric also occurs in wells located near the boundaries of the system.

3. Pre-drilling and post-drilling models of the Cianten Caldera

As described in Section 2.7, the shallow low-resistivity anomaly over the proven Awibengkok
reservoir extends northwest of the proven field into the Cianten Caldera (Fig. 3a). Here we describe
how a geologic model for the area was developed prior to drilling, and how it was modified to
incorporate well results.

3.1. Pre-drilling model for the Cianten Caldera

Currently waste geothermal brine from western production is injected at Awi 9, but in recent
years this has begun to negatively impact production and new injection options are being explored
(see Acuña et al., this issue). Two wells were recently drilled northwest of the proven reservoir
to test whether there was sufficient permeability for injection. The wells were sited based mainly
on the geophysical anomaly described above (Fig. 11c). Specific structural targets reachable
from these locations were selected based on mapping of the surface geology and a caldera model
developed from comparison of the Cianten Caldera with other calderas of similar size and tectonic
setting (Fig. 13).
Geologic mapping and geochronology indicate that the collapse of the Cianten Caldera (Fig. 3a)
occurred less than 670,000 years ago, and has since been partially filled by post-caldera lavas,
J. Stimac et al. / Geothermics 37 (2008) 300–331 327

Fig. 13. Facies model for the Cianten Caldera based on other calderas of similar size and tectonic setting. This model
focuses on the volcanic and sedimentary products found in such collapse calderas, not on the structural details of the
caldera system. Note that the topographic rim of the caldera migrates outward from the location of the original ring
fracture with time due to mass wasting and erosion of the caldera walls. It is assumed a lake forms shortly after caldera
formation and persists until the lake level overtops the lowest spot on the caldera wall. A lake may be re-established in
part or all of the caldera if drainage is blocked.

sedimentary rocks and tuffs from diverse sources. Dense vegetation and young (<40,000 years
old) caldera-fill deposits obscure the presence of alteration zones and faults that might otherwise
be exposed within this relatively old caldera. The topographic rim of the caldera is composed
primarily of andesite lava, with locally abundant clay-altered dacite autobreccia, lava and related
sediments likely defining the ring fracture zone. The topographic rim is poorly defined on its
eastern side, where it is cut by faults and covered by lava flows and related sediments that form
the western flank of Gunung Gagak, a relatively young (360,000–410,000 years old) andesitic to
dacitic stratocone with a prominent sector collapse to the north.
The Cisaketi River and its tributaries currently drain both the Cianten Caldera and the north-
western portion of the Salak geothermal field. This river flows through a narrow gap between
the northern caldera margin and Gunung Gagak, which appears to be controlled by one or more
NE-striking faults, and has probably existed in its present form for a relatively short time (i.e.
<40,000 years). One of these structures, the Muara Fault (Fig. 3a), was mapped with a 20◦ trend,
and an oblique-normal sense of displacement.
All fractures cutting the youngest units show prominent NNE to NE (0–50◦ ) and NW
(330–340◦ ) trends. The less prevalent NW and E-W fractures are more common in older lava
flows, but one prominent E-W fault near the caldera margin cuts the youngest rocks. Northeast
and NNE-striking structures probably provide the best targets for permeability, although such
faults may also serve to partition deep fluid circulation, and act as barriers to flow in an E-W to
328 J. Stimac et al. / Geothermics 37 (2008) 300–331

NW direction. The less common NW- and E-W-trending fractures may still be active and provide
some cross-linkage for hydrothermal fluid circulation.
A facies model based on a number of well-documented caldera examples was developed in
order to predict what types of rocks were expected in the Cianten Caldera (Fig. 13). It was
thought that sedimentary rocks and post-caldera lavas and domes would be present to a depth
of 200–600 m, followed by a similar thickness of ash-flow tuff and collapse breccias. It was
recognized that this “layer-cake” stratigraphy may be tilted or disrupted by post-caldera faulting
and doming. The amount of post-caldera volcanism was also unknown due to the nearly complete
veneer of surficial sedimentary and distal tuffaceous rocks. The floor of the caldera was likely to
be composed primarily of andesitic lavas, cut by intrusive rocks to a depth of 1–1.4 km. Assuming
no major regional faults had offset the basement, the depth to Miocene sedimentary rocks would
likely be greater in the caldera than in the adjacent Awibengkok area (i.e. more than 2 km), and
therefore may not be reached by drilling. However, a high in the regional gravity situated astride
the southeastern caldera rim indicated a possible shallowing of basement or intrusion in that
area.
Given the above model, it was considered possible that the low-resistivity layer observed in MT
and TDEM surveys was related to fine-grained, clay-rich sediments deposited in the caldera basin,
or in a crater lake. These types of rocks are rich in smectite, either because of their primary grain
size (clay-sized particles derived mainly from volcanic glass and altered feldspar), or because
of subsequent hydrothermal alteration. In either case, the finer-grained sediments might control
the top of any post-caldera hydrothermal circulation system. These rock types might also present
a hazard to drilling as they are likely to contain swelling clays that render boreholes prone to
collapse.
Dacite-to-rhyodacite outcrops on the eastern margin of the caldera were thought to be post-
caldera lavas/domes that might represent the surface manifestation of ring fault intrusions. Such
an intrusive mass may be, in part, responsible for the observed gravity high.

3.2. Initial drilling results and model revision

Drilling in the caldera confirmed many aspects of the geologic and geophysical models, but
also led to some significant model revisions. Methylene blue and X-ray diffraction data from
cuttings have confirmed that the low-resistivity anomaly present in the caldera is related to
smectite alteration that is underlain by progressively higher-temperature alteration assemblages
typical of the adjacent Awibengkok geothermal reservoir. However, subsurface temperatures in
the caldera are markedly lower than is implied by the alteration, indicating that the area has cooled
in response to declining hydrothermal activity (Fig. 11c). Late-stage veins filled with calcite and
zeolite are common, suggesting that cooler fluids have invaded the area and sealed many of the
fractures.
Miocene sedimentary rocks are present at shallower depths than anticipated based on the caldera
model. These rocks are locally metamorphosed to high-grade calc–silicate skarn and hornfels.
Intrusive rocks are fine-to-medium grained diorite and granodiorite and quartz-dacite porphyry.
The porphyry appears to be filling the ring fault of the caldera and may represent a more extensive
ring dike and sill complex (Fig. 13).
The injection potential of one well is being evaluated through injection of condensate at high
pressure (Yoshioka et al., 2008). While the initial results indicate low permeability, other injection
wells at Awibengkok have shown improved injection rates with time, especially if hydraulic or
acid stimulation is applied (Jorge Acuña, personal communication, December 2007).
J. Stimac et al. / Geothermics 37 (2008) 300–331 329

4. Conclusions

The Awibengkok field, operated since 1994, is the largest geothermal power producer in
Indonesia. The well field taps an extensive liquid-dominated hydrothermal system with benign
chemistry and initial temperatures in the 240–312 ◦ C range. Hydrothermal circulation is hosted
primarily by late-Tertiary to Pleistocene andesitic to rhyolitic volcanic rocks with propylitic-to-
phyllic alteration. A typical smectite-rich cap overlies progressively higher-temperature alteration
assemblages that host the geothermal reservoir.
The commercial reservoir permeability is controlled by an interconnected network of fractures
affecting much lower permeability rocks. Fluid flow occurs primarily along the strike of N- to NE-
trending faults and fractures that partition the reservoir into several sectors with distinct physical
and chemical characteristics. These sectors are more weakly linked by NW and E-W structures
and stratigraphically controlled aquifers. Reservoir porosity averages about 10% and shows a
declining trend with depth due to compaction of fragmental deposits and progressively more
intense hydrothermal alteration.
Upflow of high-temperature fluid occurs primarily on the western side of the field where
drilling has revealed a deep reservoir top (below sea level) with underlying dioritic-to-granodioritic
intrusions. These intrusions reached shallow levels on a series of N-to-NE trending faults on the
south flank of Gunung Gagak, a relatively young andesitic stratovolcano, and the eastern margin
of the Cianten Caldera.
A much shallower reservoir top (∼500 m asl) on the eastern side of the field is associated with
silicic intrusion and young dacitic-to-rhyolitic volcanism that occurred along the NNE-trending
Awi Fault. The largest fumarole field in the area and the inferred vent of the most recent eruption
(less than 40,000 years old) also lies along this fault trend. Because of exploitation, a steam cap has
formed in this portion of the Awibengkok reservoir, with minor leakage of steam-heated waters
along permeable faults.
Areas of low resistivity to the north and west of the geothermal field are being evaluated as possi-
ble reservoir extensions. Drilling results from the Cianten Caldera indicate subcommercial temper-
atures, abundant intrusive rocks, and relatively shallow basement rocks. Northeast-trending faults
and a ring dike intrusion appear to restrict fluid circulation from the proven reservoir to the caldera.

Acknowledgements

We are indebted to Chevron and PLN for permitting us to publish this work, and are especially
grateful to Bayong Chandra and Barry Andrews of Chevron IBU for granting internal approval.
We would also like to acknowledge the largely unpublished technical contributions made at
Awibengkok by Unocal and Chevron geoscientists including Mark Mosby, Dave Rohrs, Bill
Cumming, Rich Gunderson, and Phil Molling. The patience of Dudi Wikanda, who produced
most of the figures, is also appreciated. Finally we thank Rick Allis and Joe Moore for their
helpful reviews and the editors for their constructive comments and meticulous attention to detail.

References

Acuña, J.A., Stimac, J.A., Sirad-Azwar, L., Passiki, G.R., this issue. Reservoir Management at Awibengkok Geothermal
Field, West Java, Indonesia. Geothermics.
Anderson, E., Crosby, D., Ussher, G., 2000. Bulls-eye-simple resistivity imaging to reliably locate the geothermal reservoir.
In: Proceedings of the World Geothermal Congress 2000. Japan, pp. 909–914.
330 J. Stimac et al. / Geothermics 37 (2008) 300–331

Anderson, T.D., 1996. Analysis of Continuous Core Samples from the Awi 1-2 Corehole, Awibengkok Geothermal Field
West Java, Indonesia. Unpublished Unocal report, 37 pp.
Bertani, R., 2005. World geothermal generation in the period 2001–2005. Geothermics 34, 651–690.
Cumming, W.B., Mulyadi, 1984. Gunung Salak Contract Area Geophysical Summary. Unpublished Unocal report,
48 pp.
Cumming, W.B., Nordquist, G.A., Astra, D., 2000. Geophysical Exploration for Geothermal Resources: An Application
of Combined MT-TDEM, vol. 19. Society of Exploration Geophysics, Extended Abstracts. pp. 1071–1074.
Ganefianto, N., Shemeta, J., 1996. Development strategy for the Awibengkok geothermal field, west Java, Indonesia.
Proceedings Indonesian Petroleum Association, 453–462.
Gunderson, R.P., Cumming, W.B., Astra, D., Harvey, C., 2000. Analysis of smectite clays in geothermal drill cuttings by
the methylene blue method: for well site geothermometry and resistivity sounding correlation. In: Proceedings of the
World Geothermal Congress, Japan, pp. 1175–1181.
Hamilton, W., 1979. Tectonics of the Indonesian Region. U.S. Geological Survey Professional Paper 1078, 345 pp.
Handley, H.K., 2006. Geochemical and Sr-Nd-Hf-O isotopic constraints on volcanic petrogenesis at the Sunda arc,
Indonesia. Unpublished Ph.D. Thesis. Department of Earth Sciences, Durham University, Durham, UK, April, 286 pp.
Handley, H.K., Davidson, J.P., Macpherson, C.G., Gertisser, R., 2006. Along-arc heterogeneity in crustal architecture
and subduction input at the Sunda arc, Indonesia. 16th Annual V.M. Goldschmidt Conference, Melbourne, Australia.
Geochimica Cosmochimica Acta 70 (Suppl. 1 (18)), A227.
Handley, H.K., Davidson, J.P., Macpherson, C.G., Stimac, J.A., in preparation. Untangling differentiation in arc lavas:
constraints from unusual minor and trace element variations at Salak Volcano, Indonesia. Chemical Geology.
Henley, R.W., Ellis, A.J., 1983. Geothermal systems, ancient and modern. Earth Science Reviews 19, 1–50.
Hulen, J., Anderson, T.D., 1998. The Awibengkok, Indonesia, Geothermal Research Project. In: Proceedings 23rd
Workshop on Geothermal Reservoir Engineering. Stanford University, pp. 256–263.
Hulen, J., Lutz, S., 1999. Alteration mineralogy and zoning in Corehole AWI 1-2, Awibengkok Geothermal Field, West
Java, Indonesia. Geothermal Resources Council Transactions 23, 19–23.
Hulen, J., Goff, F., Woldegabriel, G., 1999. Hydrothermal Breccias in the AWI 1-2 Corehole, Awibengkok Geothermal
Field, West Java, Indonesia. Geothermal Resources Council Transactions 23, 13–17.
Hulen, J.B., Heizler, M.T., Stimac, J.A., Moore, J.N., Quick, J.C., 1997. New constraints on the timing of magmatism,
volcanism, and the onset of vapor-dominated conditions at The Geysers steam field, California. In: Proceedings of the
22nd Workshop on Geothermal Reservoir Engineering. Stanford University, pp. 75–82.
Hulen, J.B., Stimac, J.A., Sugiaman, F., 2000. The Awibengkok core research program. Part II. Stratigraphy, volcanic
facies, and hydrothermal alteration. In: Proceedings of the World Geothermal Congress, Japan, pp. 1271–1276.
Hutchinson, C.S., 1989. Geological Evolution of South-East Asia. Claredon Press, Oxford Science Publications, Oxford,
UK, 368 pp.
Ibrahim, R.F., Fauzi, A., Suryadarma, 2005. The progress of geothermal energy resources activities in Indonesia. In:
Proceedings of the World Geothermal Congress 2005. Antalya, Turkey, paper 0142, 7 pp.
Kato, O., Okabe, T., Shigehara, S., Doi, N., Tosha, T., Koide, K., 2000. Permeable fractures and in-situ stress at the
Kakkonda Geothermal field, Japan. In: Proceedings of the World Geothermal Congress, Japan, pp. 1337–1342.
Majer, E.L., Baria, R., Stark, M., Oates, S., Bommer, J., Smith, B., Asanuma, H., 2007. Induced seismicity associated
with enhanced geothermal systems. Geothermics 36, 185–222.
McCaffrey, R., 1991. Slip vector and stretching of the Sumatran fore arc. Geology 19, 881–884.
Milesi, J.P., Marcoux, E., Sitorus, T., Simandjuntak, M., Leroy, J., Bailly, L., 1999. Pongkor (West Java) Indonesia: a
Pliocene Supergene-enriched Epithermal Au–Ag–(Mn) deposit. Mineralium Deposita 34, 131–149.
Molling, P., Rohrs, D., 1996. Awibengkok reservoir baseline geochemistry. Unpublished Unocal report, 106 pp.
Moore, J., Norman, D., 1999. The thermal and chemical evolution of the hydrothermal minerals in Awibengkok
1-2, Awibengkok Geothermal Field, West Java, Indonesia. Geothermal Resources Council Transactions 23,
25–29.
Mosby, M.D., 1984. A geological review of the Awibengkok geothermal field, Java, Indonesia. Unpublished Unocal
report, 30 pp.
Murray, L.E., Rohrs, D.T., Rossknecht, T.G., Aryawijaya, R., Pudyastuti, K., 1995. Resource evaluation and development
strategy Awibengkok field. In: Proceedings of the World Geothermal Congress, Florence, Italy, pp. 1525–1529.
Noor, A.J., Rossknect, T.G., Ginting, A., 1992. An overview of the Awibengkok geothermal field. Proceedings Indonesian
Geothermal Association, 597–604.
Rohrs, D., 1986. Fluid Geochemistry of the Gunung Salak Contract Area. Unpublished Unocal report, 52 pp.
Shemeta, J.E., 1994. Regional tectonic stress orientation in the vicinity of the Awibengkok geothermal field, Java,
Indonesia. Unpublished Unocal report ASPT 94-1M, 19 pp.
J. Stimac et al. / Geothermics 37 (2008) 300–331 331

Soeparjadi, R., Horton, G.D., Bradley, E., Wendt, P.E., 1998. A review of the Gunung Salak geothermal expansion project.
In: Proceedings of the 20th New Zealand Geothermal Workshop. Auckland, pp. 153–158.
Stark, M.A., 1992. Microearthquakes—a tool to track injected water in The Geysers reservoir. In: Stone, C. (Ed.),
Monograph on The Geysers Geothermal Field. Geothermal Resources Council Special Report 17. pp. 111–117.
Stimac, J., 2003. Surface geology of the Awibengkok geothermal development area, Java, Indonesia. Unpublished Unocal
report, 38 pp.
Stimac, J.A., Sugiaman, F., 2000. The AWI1-2 Core Research Program. Part I. Geologic overview of the Awibengkok
Geothermal Field, Indonesia. In: Proceedings of the World Geothermal Congress, pp. 2221–2226.
Stimac, J.A., Powell, T.H., Golla, G., 2004. Porosity and permeability of the Tiwi Geothermal Field. Philippines based
on continuous and spot core measurements. Geothermics 33, 87–107.
Sugiaman, F., 2003. State of Stress and Wellbore Stability in Awibengkok Field. Unpublished Unocal report, 20 pp.
Syafrizal, Imai, A., Motomura, Y., Watanabe, K., 2005. Characteristics of gold mineralization at the Ciurug vein, Pongkor
Gold–Silver Deposit, West Java, Indonesia. Resource Geology 55, 225–238.
Tregoning, P., Brunner, F.K., Bock, Y., Puntodewo, S.S.O., McCaffrey, R., Genrich, J.F., Calais, E., Rais, J., Subarya,
C., 1994. First geodetic measurement of convergence across the Java Trench. Geophysical Research Letters 21,
2135–2138.
Yoshioka, K., Izgec, B., Pasikki, R., 2008. Optimization of geothermal well stimulation design using a geomechanical
reservoir simulator. In: Proceedings of the 33rd Workshop on Geothermal Reservoir Engineering. Stanford University,
28–30 January, 9 pp.

You might also like