Sie sind auf Seite 1von 19

Available online at www.sciencedirect.

com

Geothermics 37 (2008) 347365

Overview of the Wayang Windu geothermal field,


West Java, Indonesia
Ian Bogie a, , Yudi Indra Kusumah b , Merry C. Wisnandary b
aSinclair, Knight Merz Ltd., PO Box 9806, Newmarket, Auckland, New Zealand
b Magma Nusantara Ltd., Wisma Mulia 50th Floor, Jl. Jend. Gatot Subroto no. 42,
Jakarta 12710, Indonesia
Received 25 March 2008; accepted 25 March 2008
Available online 16 May 2008

Abstract
The Wayang Windu geothermal field, West Java, Indonesia, is interpreted to be transitional between
vapour-dominated and liquid-dominated conditions with four coalesced fluid upwelling centres that generally
become younger and more liquid-dominated towards the south. Two of these centres are associated with the
large Gunung Malabar andesite stratovolcano and the other two with the smaller aligned Gunung Wayang and
Gunung Windu andesitic volcanoes to the south. The overall potential resource area is of the order of 40 km2 .
Deep wells encounter a deep liquid reservoir whose top, which ranges from 0 to 400 m above sea level (m
asl) becomes progressively deeper toward the south. As pressure versus elevation conditions are the same
throughout the deep liquid reservoir it is likely to be contiguous. This liquid-dominated reservoir is overlain
by three separate vapour-dominated reservoirs. The northernmost is the largest as it is coalesced over two
separate fluid upwelling centres. Its low gas content, size, prolonged productivity and isobaric for elevation
nature, preclude it from being a parasitic steam zone. Mineralogical relationships demonstrate that this
vapour zone was originally liquid-dominated with a deep water level as high as 1700 m asl. Subsequent boil
off may reflect low recharge rates due to hydrological isolation at depth. To the south, the vapour-dominated
reservoirs decrease in thickness and are characterized by progressively higher pressures, temperatures and
gas contents. These changes suggest that the southernmost vapour-dominated zone is the youngest and that
these zones become increasing older to the north.
2008 Elsevier Ltd. All rights reserved.

Keywords: Vapour dominated; Liquid dominated; Wayang Windu geothermal field; Java; Indonesia

Corresponding author. Tel.: +64 9 913 8900; fax: +64 9 913 8901.
E-mail address: ibogie@skm.co.nz (I. Bogie).

0375-6505/$30.00 2008 Elsevier Ltd. All rights reserved.


doi:10.1016/j.geothermics.2008.03.004
348 I. Bogie et al. / Geothermics 37 (2008) 347365

1. Introductionregional geothermal setting

The Wayang Windu geothermal field is located approximately 35 km south of Bandung, the
provincial capital of West Java, Indonesia (Fig. 1). It is one of a cluster of geothermal fields around
Bandung that also includes Darajat (Hadi et al., 2005), Kamojang (Utami, 2000), Karaha-Telaga
Bodas (Moore et al., 2002, 2004), Papandayan (Wibowo, 2006), Patuha (Layman and Soemarinda,
2003), Tampomas (Wibowo, 2006) and Tangkuban Perahu (Wibowo, 2006).
These fields lie within andesitic, volcanic highlands formed by a concentration of volcanic
centres in this part of the Sunda Arc. The city of Bandung is located in a basin (Dam, 1994) near
the centre of the highlands. That basin does not appear to be a back-arc basin, as it has arc volcanics
on either side, but may owe its origin to flexure from varying rates of subduction roll back along
the Sunda Arc. This arc has formed in response to the subduction of the AustralianIndian Plate
beneath the Eurasian Plate. It has been active since the Cretaceous (Whittaker et al., 2007), but has
undergone changes as increasing amounts of Australian continental crust have become involved
in the collision and it is undergoing roll back. The dominant strike directions of the major faults
in West Java are 40 and 340 (Wibowo, 2006; Fig. 1), forming a conjugate pair of strike-slip
faults, consistent with compression due to near-perpendicular subduction.

2. Geothermal exploration

Initial exploration of Wayang Windu was undertaken by Pertamina (Sudarman et al., 1986). It
included sampling and analysis of thermal springs, DC resistivity (Schlumberger array) traversing,

Fig. 1. The distribution of Quaternary volcanic rocks and high-temperature geothermal fields in West Java, Indonesia.
I. Bogie et al. / Geothermics 37 (2008) 347365 349

Fig. 2. Topographic map of the drilled portion of the Wayang Windu geothermal field; contour interval: 100 m; bold
contour: 2000 m asl. Also shown are the locations of drill pads, well tracks and of sections A and B described in
Figs. 4 and 6.

head-on resistivity profiling, magnetotelluric (MT) gravity and soil geochemistry surveys, as
well as the drilling of temperature gradient holes.
The first deep hole (then called WWD-1, now WWA-1ST, after being side-tracked to the west,
Fig. 2) was drilled by Pertamina just to the west of the saddle between Gunung Wayang1 and

1 The abbreviation G. will be used for Gunung (mountain) from here on.
350 I. Bogie et al. / Geothermics 37 (2008) 347365

G. Windu in 1991 (Budiardjo, 1992; Ganda and Hantono, 1992; Ganda et al., 1992). Well data
showed that a perched steam-heated groundwater aquifer overlies a two-phase vapour-dominated
zone that in turn overlies a neutral-Cl liquid-dominated reservoir. This was the discovery well for
the Wayang Windu field and for transitional liquidvapour type geothermal systems. A 600-m
deep slim hole (MSH-1) drilled by Pertamina in 19931994 on the southern slopes of G. Malabar
also provided indications of the existence of a shallow two-phase zone, overlain by a steam-heated
perched aquifer further to the north.

2.1. Thermal manifestations and surcial hydrothermal alteration

The most intense surficial hydrothermal activity occurs adjacent to the small volcanic centres,
G. Wayang and G. Windu (Fig. 3). Fumaroles, steaming and altered ground, and acidsulfate
springs occur in the Wayang thermal area. The Wayang thermal area lies within a sector collapse,

Fig. 3. Location of geothermal wells, thermal features, volcanic summits, calderas and sector collapses in the Wayang
Windu geothermal field in relation to the base of the conductor.
I. Bogie et al. / Geothermics 37 (2008) 347365 351

with the current peak of G. Wayang representing an eastern remnant of a much larger volcanic
centre, originally located to the west along the axis of alignment of other small volcanic centres.
The matrix of the debris flow from the sector collapse consists of hydrothermal clay, suggesting
slope failure was related to the weakening of the volcanic deposits by hydrothermal alteration or the
result of the increase of pore pressures with heating (Reid, 2004). A radiocarbon date of 7450 110
years was obtained from peat from a small swamp that had developed on top of the debris flow
deposits at drill pad WWD. This would represent a minimum age for the sector collapse itself.
Smaller areas of altered ground with acidsulfate springs and weak fumarolic activity are found
on the southern slopes of G. Malabar, while warm, neutral-bicarbonate-sulfate springs are present
south of G. Malabar and to the south, west and east of the smaller volcanic centres (Fig. 3). The
springs have temperatures ranging from 25 to 66 C and are notable for their lack of Cl (Sudarman
et al., 1986). The Cibolang spring in the south has an elevated B content (16 ppm). As its other
constituents indicate that it discharges steam-heated ground water, the high B content is suggestive
of high-temperature boiling at depth, because B is volatile at high temperatures (Ellis and Mahon,
1977).
The northernmost area of hydrothermal alteration is exposed on the southwest rim of the
Malabar Caldera complex and there is strong alteration in the cirque of the Wayang sector collapse.
Small patches of altered ground are scattered around the area, although the overall extent of the
hydrothermal alteration is only apparent where deep cuts have been made for roads and drill pads
as much of the hydrothermal alteration is covered by a sequence of young unaltered ash beds.
A paleosol between the overlying unaltered ash beds and the underlying hydrothermally altered
deposits in the vicinity of the WWQ pad has yielded a radiocarbon age of 8700 90 years BP.
This represents a minimum age for the underlying alteration. Thus, only the alteration found in
the immediate vicinity of thermal features can clearly be regarded as current.
The northernmost known thermal activity in the field occurs at a break in slope southeast of
Puncak Besar (Figs. 2 and 3). There are no known thermal manifestations on the highest parts and
northern slopes of G. Malabar despite the evidence from the MT surveys for the field to extend
beneath these areas. As the prevailing weather during the rainy season is from the north, a rain
curtain may be obscuring thermal activity where precipitation rates are highest.

3. Field development

The Wayang Windu field was developed by MNL (Magma Nusantara Ltd.) beginning in 1996
as a fast track development that started in the logistically easier areas in the south and east
with a combination of 1500-m deep slim holes (WWC-SH, WWJ-SH, WWL-SH and WWR-SH;
Nurruhliati, 1996; Thaysa, 2003) and deeper production drilling. The existence of a large thermal
anomaly was established, but productive wells were restricted to sites immediately beneath and
southwest of the young volcanic centres.
Wells drilled north of G. Bedil encountered a shallower, two-phase, vapour-dominated reservoir
than the one found in the first well (WWA-1). These wells confirmed the results from slim hole
MSH-1 in the north that indicated a vapour-dominated regime overlies a deep liquid-dominated
reservoir. Thus, a combination of deep and shallow wells was drilled; success was greatest with
the shallow wells. Three dry-steam wells with depths of up to 1700 m, each produced steam
equivalent to over 20 MWe.2 These wells were drilled on the northernmost (at the time) MBD

2 All MWe values given are in terms of the power conversion capacity of the current Wayang Windu power plant (i.e.

1.94 kg/s of steam per MWe).


352 I. Bogie et al. / Geothermics 37 (2008) 347365

pad. In terms of resource potential, an initial 220 MWe (gross) development was planned with a
possible extension to 440 MWe (gross), until the Asian financial crisis of 1997 intervened and the
project was scaled back to 110 MWe (gross).
The initial 110 MWe (gross) development obtained its main steam supply from the northern
two-phase reservoir, with some deep northern production and a combination of shallow and deep
production from wells further to the south, on the WWA pad upon which the Pertamina discovery
well was located. Two-phase fluid transmission pipelines from these drill pads feed a central
separator station with steam passing through a scrubber before entering the dual inlet, 110 MWe
Fuji turbine in the power plant. This unit, which was installed in 1999, is one of the worlds largest
operating geothermal turbines (Murakami et al., 2000). As a result of this turbine installation,
Wayang Windu holds the distinction of being the most rapidly developed geothermal field of its
size. Both condensate and separated brine are reinjected by gravity in the southernmost part of
the known resource.
Unocal Indonesia became a 50% shareholder of the WayangWindu project in 2001. The
poorly productive deep MBD-1 vertical well was side-tracked to intersect the shallow, vapour-
dominated zone and came to be the largest producer in the field at the time. That was followed
by work-overs of other wells on the MBD pad that had reduced production due to the installation
of tie backs (whereby the production casing shoe is cased to the surface reducing the diameter
of the cased section of the well). WWQ-3, another deep well, was also side-tracked but with less
success than MBD-1. Reservoir pressure drawdown has reduced fluid production with time, but
it has now stabilized in wells tapping the vapour-dominated zones, indicating that they are major,
sustainable productive reservoirs rather than limited parasitic ones.
The field was acquired by Star Energy Holdings Pty Ltd. in 2004. A new drilling program began
in August 2006 to supply steam for a second 110 MWe Fuji turbine at the existing power plant.
The first well completed under this program (MBD-5) produces the steam equivalent of 40 MWe
making it at the time it was drilled the largest dry steam well in the world. The whole eight-well
program realized a total of 180 MWe and included make up wells for the existing turbine. The
northern extension of the field is currently being explored with the ultimate goal of obtaining
steam to generate 440 MWe (gross), which on the results of modeling studies of the field (Asrizal
et al., 2006) is eminently achievable.

3.1. Extent of the Wayang Windu geothermal eld

Only the southern part of the western boundary of the Wayang Windu geothermal field is well
defined by drilling (Figs. 2 and 3). Anderson et al. (1999, 2000) found a good relationship between
the contoured elevation of the base of the conductive layer (i.e. the conductor) produced by
hydrothermal smectite, as determined by MT surveys and the temperature distribution in the
drilled portion of the field. Thus, the continuation of elevated portions of the conductor outside
the drilled area can be considered indicative of the northern extent of the field (Fig. 3).
Southwest of G. Windu, productive wells have been drilled on the basis of earlier Schlumberger
and MT ressitivity surveys, whereas later re-interpretation of new MT data indicated a deep base
of the conductor with no doming or ridging. Since there are indications that the southern area is the
youngest part of the system (see below), the position of the conductor is likely to have been dictated
by earlier geothermal activity when the deep liquid reservoir was higher. The conductor in the south
now appears to be too impermeable for the alteration mineralogy to re-equilibrate and allow forma-
tion of a dome or ridge in its base, although the resistivity in the area of productive wells southwest
of G. Windu is slightly higher than where the non-productive WWE-1 and WWA-1ST were drilled.
I. Bogie et al. / Geothermics 37 (2008) 347365 353

Combining new well data with recent MT surveys, which have a greater station density and
have yielded better quality data, the bulk of the field is interpreted to lie beneath G. Malabar, the
andesite stratovolcano now centred at Puncak Besar (Fig. 3). The potential resource in the north
is estimated to be approximately 4 km wide in an EW direction and to extend approximately
14 km to the south beneath a series of aligned, small volcanic centres, where it narrows down to
approximately 2 km across. This gives an overall potential resource area of approximately 40 km2 .
However, since slim hole drilling outside of the resource area encountered elevated temperatures
at depth, the actual geothermal system could be much larger.
The field owes its size to the presence of more than one fluid upwelling centre, as discussed
below. This feature is found in other geothermal fields in Java. Layman et al. (2002) recognize
three centres at Dieng, multiple centres (referred to as cells) are suggested for Karaha-Telega
Bodas (Nemcok et al., 2007) and two geothermal centres are recognized at Awibengkok (Hulen
and Anderson, 1998).

4. Geology of the eld

The stratigraphy of Wayang Windu has been discussed by Bogie and Mackenzie (1998) who
applied volcanic facies models to subdivide the various volcanic units at depth. These units define
a series of overlapping andesitic piles. Their cross section has been extended in Fig. 4, with the
trace of the section shown in Fig. 2. Microdiorite, dolerite and diorite porphyry dykes are found,
but blind drilling and very limited coring have prevented the clear recognition of any major
intrusives. Andesitic lavas, pyroclastic and epiclastic deposits predominate in the volcanic units
with dacite only occurring at G. Gambung. Quartz found in rocks of the other smaller volcanic
centres is xenocrystic, and geochemically these rocks are andesites.
Ash deposits of regional extent occur throughout these volcanic piles. Similar beds are also
found at Patuha (Layman and Soemarinda, 2003) and Awibengkok (Hulen and Anderson, 1998),
where they are referred to as paleosols. At Wayang Windu there are 14 different beds with
thicknesses varying between 5 and 30 m that can be correlated between wells, and numerous
thinner ones with solitary occurrences. These beds have a complex distribution suggesting draping
over the existing topography with erosion in steeper areas and ponding when deposited in valleys.

Fig. 4. Section across the Wayang Windu geothermal field showing well tracks, geological units (extended north and
south from that of Bogie and Mackenzie, 1998) and the top of epidote (section location is given in Fig. 2).
354 I. Bogie et al. / Geothermics 37 (2008) 347365

Scanning electron microscopy and XRD analyses indicates that they originally consisted of very
fine-grained glass shards and titanomagnetite, but the glass has altered to calcium smectite and
in places interlayered smectite-illite. As these clays are usually found at temperatures of less
than 200 C (Anderson et al., 2000), we suggest the beds must have very low permeability, since
measured temperatures at the corresponding depths (>300 C in some instances) are much greater
than the typical stability limit of the clay.
G. Malabar sits on the boundary fault of the Bandung basin (Figs. 1 and 3; Dam, 1994) There
is a multiphase summit caldera complex on G. Malabar. Rocks from the summit of G. Malabar to
the east of the calderas, Puncak Besar a prominent peak south of the caldera complex directly
above the Bandung Basin boundary fault and G. Gambung a parasitic dacite dome to the
southeast all have K-Ar dates of 0.23 Ma and have bulk and trace element chemistries of a
differentiated series (Bogie and Mackenzie, 1998). G. Bedil, the next volcanic centre to the south,
was dated at 0.19 Ma and G. Windu, the southernmost volcanic centre at 0.10 Ma. The 0.49 Ma
date for G. Wayang breaks the trend of having younger centres towards the south. While the
other young volcanic centres are well preserved and samples for dating were obtained from the
youngest part of the volcano, G. Wayang has undergone sector collapse, and it is likely that
the dated sample from that eruptive centre was taken from a much older part of the volcanic
pile.
The geochemistry of the younger volcanic centres is variable, although G. Windu has some trace
element similarities to G. Malabar. As these rocks contain quartz xenocrysts and diorite xenoliths,
the chemical variation may be reflecting the degree of crustal assimilation by an original magma
similar to that producing the Malabar rocks. A more extensive, but less intensive geochemical
study of samples of older units collected in the wells concluded that they are geochemically similar
to the Malabar rocks (Asrizal et al., 2006).
Structurally the field conforms best to regional patterns in the south, with faulting exhibiting
steep dips (>80 ) and strikes of 3040 and 330340 . In the north, along the southern boundary
of the Bandung Basin, further deformation results from movement along a boundary fault. G.
Malabar is actively subsiding into the basin and is deforming the basin fill, as can be seen by
the presence of upthrust Tertiary sediments (Alzwar et al., 1992) as northern foothills to G.
Malabar.
The wells have highly localized structural permeability, with the most permeable geologic
structures following the regional trend of 40 . As these structures have trends similar to regional
faults, then it is likely they are strike-slip faults. Strike-slip faults tend to have lower permeabilities
than normal faults because of shearing and consequent rock comminution. Structures that were
reactivated as the volcanic sequence was deposited will display less shearing at their upper ends
than their extensions into the basement. Thus, the same faults may have lower permeabilities at
depth due to prolonged shearing and comminution of the rock compared to the younger overlying
volcanic pile; at least in competent lithologies.
Bandyopadhay et al. (2006), however, calculated the direction of least principal stress in the
Wayang Windu area as far north as the WWQ pad, utilizing borehole breakouts (the tendency for
drill hole cross sections to become elongated in relationship to the local stress field) as determined
from the caliper measurements obtained from micro-resistivity formation imaging logging. They
found that the calculated stress field did not correspond to the regional orientation, but had the
least principal stress striking at 310 , with an overall normal faulting regime. The NE-striking
faults are thus likely to have been regional strike-slip faults reactivated as more permeable normal
faults due to a change from a regional compressive to a local extensional regime. Further to the
north, extension may be even stronger at the boundary and inside of the Bandung Basin.
I. Bogie et al. / Geothermics 37 (2008) 347365 355

Table 1
Alteration mineralogy (possible relict phases shown in italics)
Location Initial alteration Over print

Above conductor Opal, cristobalite, kaolinite, alunite,


natro-alunite and sulfur
Conductor Smectite, illite-smectite, quartz, chlorite, albite, Kaolinite, anhydrite,
calcite, pyrite, heulandite, mordenite, calcite, quartz
clinoptlolite, stilbite, analcime, laumonite
Vapour-dominated reservoirs Quartz, chlorite, calcite, albite, pyrite, Anhydrite, calcite, pyrite
illite-smectite, corrensite, epidote, illite,
chalcedony, wairakite
Amphibole
Pyrophyllite, diaspore, quartz, anhydrite
Deep liquid reservoir Quartz, chlorite, illite, pyrite, wairakite, epidote,
prehnite, adularia, albite, tourmaline
Dickite, pyrophyllite, quartz, woodhouseite,
pyrite
Amphibole, orthoclase, magnetite
Diopside, oligoclase, magnetite

4.1. Hydrothermal alteration at depth

Hydrothermal alteration (Table 1) at depth is most strongly developed in the pyroclastic deposits
with more structurally limited alteration zones in the lava flows. Shallow alteration (above and
locally within the conductor), is marked by the presence of kaolinite, alunite, natroalunite and rare
native sulfur. This alteration is associated with perched steam-heated groundwater aquifers mainly
in the vicinity of updoming in the base of the conductor, pressure profiles from slim holes drilled
to depths of up to 1500 m that did not penetrate into the deep reservoir (Fig. 5) indicate perched
aquifers are widespread and probably continuous. As the warm springs discharge at elevations
similar to those of perched steam-heated aquifers, it is likely that they are fed by them.
Possible condensate aquifers, not connected to the regional ground waters, are indicated by high
porosities measured by using a down hole magnetic resonance imaging logging tool and these are
characterized by strong alteration to kaolinite, calcite, quartz and anhydrite. The conductor itself
is made up mainly of an argillic assemblage dominated by smectite along with near ubiquitous
quartz, chlorite, calcite and pyrite with zeolites, including heulandite, mordenite, clinoptilolite,
stilbite, analcime and laumontite. Kaolinite, calcite, anhydrite and quartz are found within parts
of the conductor occurring as an overprint.
With increasing depth, interlayered illite-smectite rather than smectite is found, until there
is a transition to a propylitic assemblage with its top marked by the presence of corrensite and
epidote. At greater depths, illite becomes the main sheet silicate. Secondary amphibole, orthoclase
and magnetite making up a high-temperature potassic assemblage are encountered at still greater
depths. The formation of secondary amphibole appears to be related to dike emplacement. A
contact metamorphic assemblage of diopside, oligoclase and magnetite has been observed in
WWA-4. In the deep liquid reservoir, wairakite and prehnite, along with epidote, are common as
alteration and vein minerals, with less common adularia.
Other than a generally prograde transition, with illite-smectite and corrensite detected shallow
in the vapour-dominated reservoirs, and prehnite in the lower parts of the deep liquid reservoir,
356 I. Bogie et al. / Geothermics 37 (2008) 347365

Fig. 5. Graph of pre-production pressures in main feed zones of Wayang Windu geothermal wells versus elevation. Note
that the casing in well WWE-1 has a hole allowing the inflow of water from a shallow perched aquifer.

there is no clear difference in the hydrothermal alteration of the liquid- and vapour-dominated
reservoirs. Wairakite, however, is common in the vapour-dominated reservoir, as has also been
noted at Karaha-Telega Bodas (Moore et al., 2002).
Rare occurrences of advanced argillic alteration are found below the conductive cap, both in the
vapour-dominated and the deep liquid reservoirs. Minerals of this assemblage include pyrophyllite,
diaspore, woodhouseite and dickite, with accompanying quartz, anhydrite and pyrite. As these
hydrothermal minerals form under high-temperature, acid conditions and considering that the
I. Bogie et al. / Geothermics 37 (2008) 347365 357

Fig. 6. Section across the Wayang Windu geothermal field showing well tracks, isotherms and the known location of the
tops of the vapour-dominated and deep liquid reservoirs (section location is given in Fig. 2).

reservoir pH is now near neutral, and has temperatures below those at which these minerals are
formed (Reyes et al., 1993), we consider this deep advanced argillic alteration to be relict. It may
possibly reflect the earlier presence of acidic condensed magmatic volatiles, particularly since
woodhouseite (CaAl3 PO4 SO4 (OH)6 ; Stoffregren and Alpers, 1987) has an exclusively magmatic
association (Bogie and Lawless, 2000).
In the northern part of the Wayang Windu geothermal field, hydrothermal epidote is found at
elevations of up to 1330 m above sea level (m asl). The shallowest appearance of this mineral
is above the vapour-dominated reservoir, and all its first occurrences in well samples are above
the deep liquid reservoir (Figs. 4 and 6). As epidote generally forms in geothermal fields at
temperatures above 240 C under near neutral pH conditions in a liquid reservoir (Browne, 1978),
we infer that the water level in the system was previously higher. If the geothermal system is
related to recent volcanism, then it is no older than 0.23 Ma. Thus lowering of the water level
could not be caused by tectonic uplift, which occurred much earlier, during the Pliocene (Alzwar
et al., 1992).
The epidote occurs 400 m below the top of the smectite-bearing argillically altered rocks,
which marks the top of the conductor. If it is assumed temperatures followed boiling point-to-
depth conditions, the original water level would have been at an elevation of 1730 m asl. To the
south beneath the younger volcanic centres, the top of the conductor is found at an elevation of
1400 m asl, and there is a concurrent deepening to the first appearance of epidote (Fig. 4). Thus,
the top of the conductor to the south can be considered to correspond to the original water level.
The top of the epidote zone is very close to the top of the Waringin Unit (Fig. 4) and the
top of the vapour-dominated reservoir (Fig. 6) reflecting, perhaps, a porosity variation. In the
northern part of the field, the Malabar Unit consists mainly of lavas whose average porosity is
1% (Asrizal et al., 2006). The underlying Waringin Unit, consisting mainly of lapilli tuffs, these
have an average porosity of 8% (Asrizal et al., 2006). It is possible that the original porosity
of the pyroclastic deposits was higher prior to alteration, but this would not be the case for the
lava flows. Therefore, the thick sequence of lava flows could have acted as the initial caprock
of the system. The permeability associated with vertical faults that cut the flows may have been
restricted by the presence of the regional ash deposits, which because of their high clay content
would tend to deform plastically rather than act as hosts for vertical conduits.
In the south, the Waringin Unit contains more lava flows in its upper parts. The overlying
Pangalengan Unit, which contains many ash deposits and fine-grained sediments, could have
358 I. Bogie et al. / Geothermics 37 (2008) 347365

formed the original cap in this part the field, and thus might have influenced the distribution of
epidote. However, the plot of depth of the first appearance of epidote follows a similar shape to
the 300 C isotherm, suggesting instead that the initial temperature distribution may outweigh the
influence of porosity variations in the south.
In some places the original conductor has been overprinted by hydrothermal alteration caused
by perched steam-heated aquifers that, where topography allows, extend to higher elevations and
above topographic highs in the conductive layer. However, these aquifers have higher resistivities
(5 m) than the main conductor (2 m), due to the presence of kaolinite, which is more
resistive than smectite, as the predominant clay mineral.
The areas of the conductor that have slightly higher resistivity more clearly define the four fluid
upwelling centres at the Wayang Windu Geothermal Field (i.e. where productive wells have been
drilled; Fig. 3) than the elevation of the base of the conductive layer. Presumably this is because
the areas of higher resistivity in the conductor reflect near present conditions, whereas its base was
defined during an early stage in the development of the system. Similar higher conductivities in the
conductor have been reported at Karaha-Telaga Bodas (Raharjo et al., 2002), and may generally
serve, in combination with the geometry of the base of this conductive layer, as a pre-drilling
indicator of the potential presence of vapour-dominated reservoirs.
North of the wells drilled at Wayang Windu the main conductor is found at up to 1800 m asl.
Here it is possible that the water level may have reached this elevation during the history of the
field. Moreover, the higher resistivities indicative of perched steam-heated aquifers are found at
even higher elevations. Since the current deep liquid level is at much greater depth (400 m asl
in the north) and the alteration mineralogy is that of a near pH-neutral fluid, most of the alteration
in and above the vapour-dominated zones is now above the water level and must be relict. That
is, the alteration, which includes the electrically conductive argillic zone, must have formed early
in the development of the system. This zone, which is characterized by a conductive temperature
profile indicative of low permeability, now constitutes the caprock of the geothermal reservoir.
On the margins of the field, where the base of the conductor deepens, these clay-rich altered rocks
form the lateral hydrological boundaries of the vapour-dominated resource, at least in its upper
parts. This may explain why the vapour-dominated zone is best developed in the north where
there is the steepest drop off in depth of the margin of the conductor.
Hydrothermal alteration in the two-phase, vapour-dominated reservoirs that have formed above
the deep water is only weakly developed. Epidote is partially replaced by calcite, white clay
(possibly kaolinite), pyrite and anhydrite, a further indication of its relict nature. The pervasive
calcite veining present in the rocks may have formed as the system boiled off; but platy textures
typical of boiling (Browne, 1978) are not commonly observed. Alternatively, the calcite could
have formed by descending CO2 -rich condensate. Rarely, platy calcite and chalcedony occur in
epidote-bearing rocks. The deposition of chalcedony at temperatures above 240 C (Bogie et al.,
2003), are suggestive of more intense boiling, which could result from localized pressure drops
during fracturing. In contrast, chalcedony is widely distributed at Karaha-Telaga Bodas (Moore
et al., 2002, 2004), where its presence is interpreted as evidence of rapid boil off.
Fluid inclusion work has been limited by the amount of suitable sample material. Abrenica
(2007) reports homogenization temperatures in primary fluid inclusions in quartz from a vein
at 590 m asl in well MBD-5 between 228 and 255 C, with the mode at 235 C. The current
estimated temperature is 246 C (the downhole logging tool did not reach this depth). Melting
point measurements for these inclusions indicate salinities of 0.531.05 wt% NaCl equivalent
(Abrenica, 2007). These values reflect both the dissolved salt and gas contents of the trapped
fluids. The present-day salinities, calculated on a gas free basis, of the deep liquid reservoir is
I. Bogie et al. / Geothermics 37 (2008) 347365 359

2 wt% NaCl equivalent. Therefore, these inclusion results likely reflect early liquid reservoir
conditions.
Secondary fluid inclusions in the quartz are vapour rich and have higher vapour/liquid ratios in
successive generations, consistent with the presence of increasing vapour-dominated conditions
with time. Homogenization temperatures of the secondary inclusions range from 241 to 334 C,
but as it is unlikely that a single-phase fluid was trapped, they may not reflect temperatures
prevailing at the time.
Primary inclusions in calcite from a sample collected 2 m below the quartz sample from well
MBD-5 (i.e. at 592 m asl) gave higher homogenization temperatures of 264295 C (Abrenica,
2007), with the mode at 275 C. Like the quartz-hosted secondary inclusions, these calcite-hosted
inclusions are vapour-rich and have increasing vapour/liquid ratios in successive generations and
homogenization temperatures may not be reflecting trapping temperatures.
Epidote from 542 m asl in MBA-1 was observed to contain liquid-dominated fluid inclusions
(Abrenica, 2007), although homogenization temperatures could not be obtained. As vapour-
dominated conditions now prevail at this depth, it is inferred that the liquid level was previously
higher and that the epidote is relict.

5. Reservoir characteristics and geochemistry

Wayang Windu has a deep, hot, neutral pH, liquid reservoir that, in the area drilled, is overlain
by a perched vapour-dominated two-phase reservoir (Fig. 6). Throughout the field within the deep
liquid reservoir, pressures and temperatures versus elevation are similar (Table 2 and Fig. 5). In the
north its top is at 400 m asl, and it deepens towards the south where the top of the liquid is found
near sea level elevation. Since the reservoir has almost the same vertical pressure distribution
throughout, it can be considered to be a contiguous body, and since it is near pressure equilibrium
there is little fluid flow. It is under-pressured with respect to groundwater hydrostatic pressure and
there is geochemical evidence for only limited recharge (Suminar et al., 2003).
The deep liquid reservoir is possibly recharged from west of G. Bedil and G. Wayang and
southwest of G. Windu. In these areas, the decrease in the elevation of the base of the conductive
layer is less steep, and thus the early argillic alteration that may have created hydrological barriers
on these sides of the vapour-dominated reservoir is more limited. In the areas where recharge
may be occurring the shallowest well feed zones from the deep liquid reservoir are cooler than
elsewhere in that reservoir, but without actual temperature inversions deeper in the well. This
indicates some limited ingress of meteoric waters in the shallow zones, rather than indicating
outflows in these areas.
Two-phase vapour-dominated reservoirs overlie the deep liquid-dominated regime. The largest
in the north appears to be coalesced over two fluid upwelling centres (associated with Puncak Besar
and G. Gambung), while the two further south (associated with G. Wayang and G. Windu) appear
to be separate, giving three vapour-dominated reservoirs over four fluid upwelling centres. The
characteristics of the vapour-dominated zones change progressively towards the south: pressures,
temperatures and gas contents increase, they are found at greater depths, and their thicknesses
decrease.
The degree of communication between the vapour-dominated reservoirs is uncertain. However,
if they were all hydraulically connected, the drop in fluid pressure for any given elevation towards
the north (i.e. the oldest part of the system if the ages of the spatially related volcanic centres
are also temporally related to the fluid upwelling centres) may point to a northerly flow from
the inferred youngest part of the system in the south. The fact that the field can be divided up
360
Table 2
Geochemical and pressuretemperature properties of Wayang Windu reservoir areas prior to production
Area Deep liquid reservoir Two-phase reservoirs

I. Bogie et al. / Geothermics 37 (2008) 347365


Reservoir Cla NaKCa NCGc NCGc Measured Measured pressurea Elevation
(ppm) temperatureb ( C) (wt%) (wt%) temperaturea ( C) (bar) (m asl)

Puncak Besar d d d d d d ?1120


Gambung 12,00 295300 0.30.6 0.62.6 250260 3545 4001100
13,000

Wayang 12,000 295308 0.50.6 24.5 255267 5055 200700


13,000

Windu 6000 285300 3.5 10 260290 8085 80400


8000

a Prior to production.
b NaKCa geothermometer of Fournier and Truesdell (1973).
c Non-condensable gases in weight percent.
d Well MBB-1 drilled in this area produces 20 MWe on initial discharge. It did not penetrate into the deep liquid reservoir and fully stabilized gas, temperature and pressure

data are not yet available.


I. Bogie et al. / Geothermics 37 (2008) 347365 361

geochemically into a minimum of three areas (Table 2), and that high temperatures are maintained
at depth and are actually highest beneath G. Wayang rather than in the Windu area, argue against a
single heat source in the south. Separation of these three vapour-dominated reservoirs may reflect
the distribution of impermeable, regional ash deposits that restrict vertical permeability and the
distribution of the more porous pyroclastic-rich deposits that host the two-phase zones.
Interpretation of the MT surveys indicate that the Wayang Windu Geothermal Field may
extend approximately 7 km to the north of MBE-2 and that this well is located south of two
domed structures in the base of the conductor (Fig. 3), with MBE-2 and the major producing
wells on the MBA and MBD drill pads associated with the southern dome. In the north, a fourth
area associated with the northernmost domed structure in the base of the conductor has now been
drilled and found to be productive [well MBB-1 (see Figs. 24) produces 20 MWe].
The four possible fluid upwelling centres areas are spatially associated with four eruptive
centres (Puncak Besar, G. Gambung, G. Wayang and G. Windu, going from north to south;
Fig. 2), whose age decreases generally towards the south. When considering these four particular
areas, one finds that the deep liquid reservoir in the G. Windu area has geochemical characteristics
similar to that of geothermal reservoirs associated with andesitic stratovolcanoes elsewhere, for
example Tongonan in the Philippines (Lovelock et al., 1982). The slightly lower temperatures at
depth in the G. Windu area when compared to those measured further north suggests that any
concentration of solutes by boiling (as is indicated by the downhole pressure profiles) is being
offset by some mixing with meteoric waters.
The waters of the deep reservoirs in the G. Gambung3 and G. Wayang areas have similar Cl/B
ratios than those in the G. Windu area, but are much more saline and have lower gas contents
suggestive of boiling at high temperatures, with very limited mixing with ground waters (a process
that requires continuous heat recharge to maintain reservoir temperatures). Parts of the system
where the deep liquid reservoir is more saline would therefore have to be older than those at G.
Windu in order to provide time for this extent of boiling off to occur, which is consistent with
the age of the heat centres (i.e. generally getting younger towards the south). The variation in gas
content between the areas (Table 2) may reflect a lower gas flux from the older centres, and may
be responsible for the deepening of two-phase conditions as the increase in gas content towards
the south increases the depth of first boiling.

6. Discussion

As pressure-versus-elevation in the deep Wayang Windu liquid reservoir is the same through-
out the drilled area and the deep reservoir is under-pressured with respect to local ground waters,
the systems hydrology cannot be interpreted in terms of the liquid upflows and outflows typ-
ical of most geothermal systems associated with andesitic stratovolcanoes. The presence of a
deep neutral-Cl reservoir, the low degree of under-pressure in the deep liquid reservoir, pre-
exploitation pressures and temperatures above that of the maximum enthalpy of steam in the
vapour-dominated reservoirs and the limited depth range where vapour-dominated conditions
prevail at Wayang Windu also mean that the field cannot be strictly compared to the solely
vapour-dominated systems of Darajat or Kamojang. Wayang Windu must therefore be regarded
as a new type of geothermal field, transitional between liquid dominated and vapour dominated.

3 Fluid chemistry data on the deep liquid reservoir is lacking for the northernmost wells as they were completed above

it. Thus, all the deep liquid geochemistry data come from well MBE-2.
362 I. Bogie et al. / Geothermics 37 (2008) 347365

This transition is most advanced in the northern parts of the field where the drop in the deep water
table has been largest. Towards the south, the vapour-dominated zones are thinner and deeper
and make up proportionality less of the resource. This relationship implies a series of steps in the
transition from liquid- to vapour-dominated conditions represented in the geothermal field.
The Wayang Windu system is largely sealed off from surrounding and overlying ground water
aquifers, but still receives deep heat and fluid recharge from possibly four fluid upwelling centres
that are inferred to be progressively younger to the south. Vapour-dominated reservoirs have devel-
oped over the fluid upwelling centres, with a vapour-dominated reservoir in the north coalesced
over the fluid upwelling centres associated with the Puncak Besar and G. Gambung volcanic
centres. Two possibly separate vapour-dominated reservoirs further south, lie above separate
fluid upwelling centres associated with G. Wayang and G. Windu. Allis (2000) concluded that
stock-sized intrusives cannot provide sufficient heat to maintain a long-lived hydrothermal system
like the one at Wayang Windu. Thus, it is probable that the field is underlain by a major multi-
phase intrusion, which is the ultimate heat source. The diorite xenoliths observed in the younger
volcanics may be from the older parts of this intrusion.
The locations of the four fluid upwelling centres could overlie shallow apophyses of the larger
intrusive body, which have been fractured by a combination of secondary boiling and thermal
contraction, to provide conduits that channel steam and gas flow from a much larger intrusive
source at depth. If the ages of the eruptive centres can be related to the heat centres, it would
explain how the geothermal system has been active for possibly 0.23 Ma.
Early pulses of acidic, magmatic condensates produced from the intrusive apophyses may be
responsible for the formation of the rare advanced argillic alteration observed deep in the Wayang
Windu field. The deep liquid reservoir now has a neutral pH; thus there is no evidence for the
presence of acidic fluids produced by the condensation of magmatic volatiles reaching shallow
levels of the system.
The deep magmatic degassing is now heating and partially recharging a large exploitable
reservoir. As this deep recharge is composed of water vapour and non-condensable gases, it does
not require a large mass, only enough to maintain the large volume of the system under boiling
point versus depth curve conditions, and allow the upper part of the original water-dominated
reservoir to boil off. Whether or not there is sufficient heat flow to dry out the entire reservoir is
another problem, as the volume of the deep contiguous deep reservoir at Wayang Windu may be
proportionately larger than that of geothermal systems associated with a single or closely spaced
multiple intrusive heat conduits, as may be the case at Kamojang and Darajat.
The Wayang Windu field is part of a cluster of Indonesian high-temperature geothermal fields,
which include the vapour-dominated reservoirs of Darajat and Kamojang, and may represent
the end point of the liquid-to-vapour transition. Vapour capped reservoirs at Patuha (Layman and
Soemarinda, 2003) and Karaha-Telaga Bodas (Moore et al., 2002, 2004) exhibit magmatic vapour
cores as defined by Reyes et al. (1993), and may represent an earlier stage of transition than found
in the Windu part of the Wayang Windu Geothermal Field; the Wayang and Gambung-Puncak
Besar areas being successively further advanced in that transition.
A common feature of some of these Indonesian geothermal fields is that they underlie sector
collapses; this is most strongly the case at Darajat, where it includes most of the field. This may
possibly also be the situation at Kamojang, where a partially circular collapse feature is reported,
although this has been previously interpreted to be a caldera (Healy and Mahon, 1982). Moore et
al. (2002, 2004) consider that the sector collapse of G. Galunggung Volcano triggered the drying
out of the Karaha-Telaga Bodas system. The collapse of G. Wayang may have also contributed
to the boiling off of at least that part of the Wayang Windu geothermal system. The radiocarbon
I. Bogie et al. / Geothermics 37 (2008) 347365 363

dating (discussed in Section 2.1) indicating that this sector collapse took place after the formation
of the shallow hydrothermal alteration further north, supports this notion. Other areas of major
slope failure that may also be interpreted as sector collapses are found on the western side of G.
Windu, north of G. Gambang, west of Puncak Besar and north of G. Malabar (Fig. 3). However,
that to the north of G. Gambung and that to the west of Puncak Besar may be more closely related
to slippage on the Bandung Basin Boundary Fault. Hydrothermal alteration and activity have not
been reported from these areas and it is more difficult to relate these to changes in the hydrology
of the system.
This concentration of vapour-dominated and transitional resources associated with andesite
stratovolcanoes in the Bandung area is yet to be satisfactorily explained given their apparent
dearth in the rest of Java, or elsewhere in the world. Allis (2000) suggested that the perpendic-
ular subduction beneath Java produced compressive deformation in the upper plate restricting
the recharge from depth, but this applies all along Java, not just to the Bandung region. This
compression is enhanced by the subduction of the Roo Rise (Whittaker et al., 2007), a thickened
section of oceanic crust of the down-going Australian Plate. However, there is a break in the Roo
Rise in the Australian Plate immediately adjacent to the Bandung area, which is marked by a
large bulk-sound velocity anomaly of the down going slab (Gorbatov and Kennett, 2003), with a
gravity high directly above it (Newcomb and McCann, 1987).
Subduction roll back (Whittaker et al., 2007) is taking place all along the Javan part of the
Sunda Arc because of the subduction of the old (95135 Ma) slab of the Australian Plate. It is
likely that this rollback is locally accentuated by the thinner and denser oceanic crust adjacent to
Bandung, and that this produces local extensions in the overlying plate, which may be responsible
for the large number of volcanic centres and associated geothermal fields in the Bandung area. The
roll back may also be responsible for the occurrence of multiple heat centres in Javan geothermal
fields, as the location of the upper crustal plate and the deep magma source will vary with time
as rollback proceeds.
The fields around Bandung are in terrains of sufficient elevation for there to be room for
steam reservoirs to form above the general level of deep meteoric recharge. This recharge may
be limited by the poor permeability at depth as the faults in the basement had an initial strike-
slip movement that reduced fault permeability due to prolonged shearing. The subsequent fault
movement within the volcanic deposits was normal (at least at Wayang Windu) resulting in higher
rock mass permeabilities, thus favoring the formation of highly productive vapour-dominated
reservoirs.

Acknowledgements

The authors wish to gratefully acknowledge the permission and support of the management of
PT Star Energy to publish this paper and the help of Manfred Hochstein in doing so. Constructive
reviews by Rick Allis and Dick Henley are also acknowledged, along with the editorial assistance
of Joe Moore, Marcelo Lippmann, Sabodh K. Garg and Greg Bignell. Thanks go to Mariano
Gutierrez of Sinclair Knight Merz Ltd. and Iis Dian Indriani for their assistance with the figures.

References

Abrenica, A.B., 2007. Hydrothermal alteration and fluid inclusion studies in the northern Wayang Windu geothermal
field, West Java, Indonesia. MSc Thesis. Geological Engineering Department, Faculty of Engineering, Gadjah Mada
University, Yogyakarta, Indonesia, 151 pp.
364 I. Bogie et al. / Geothermics 37 (2008) 347365

Allis, R., 2000. Insights on the formation of vapour-dominated systems. In: Proceedings of the World Geothermal Congress
2000, Kyushu-Tohoku, Japan, pp. 24892496.
Alzwar, M., Akbar, N., Bachri, S., 1992. Geology of the Garut and Pameungpeuk Quadrangle, Jawa. Republic of Indonesia,
Department of Mines and Energy, Directorate General of Geology and Mineral Resources, Geological Research and
Development Centre. Geological Map.
Anderson, E., Crosby, D., Ussher, G., 1999. As plain as the nose on your face: geothermal systems revealed by deep
resistivity. In: Proceedings of the Twenty-first New Zealand Geothermal Workshop, University of Auckland, pp.
107112.
Anderson, E., Crosby, D., Ussher, G., 2000. Bulls-eye!simple resistivity imaging to reliably locate the geothermal
resource. In: Proceedings of the World Geothermal Congress 2000, Kyushu-Tohoku, Japan, pp. 909914.
Asrizal, M., Hadi, J., Bahar, A., Sihombing, J.M., 2006. Uncertainty quantification by using stochastic approach in pore
volume calculation, Wayang Windu Geothermal Field, W. Java Indonesia. In: Proceedings of the Thirty-first Workshop
on Geothermal Reservoir Engineering, Stanford University, Stanford, CA, USA, pp. 235242.
Bandyopadhay, I., Juandi, D., Baroek, M., Hadi, J., Pramono, B., 2006. In situ stress analysis in deviated wells using
inversion methoda case study from volcanic pyroclastics of Wayang Windu Field, Java Indonesia. In: Proceedings
of the Jakarta 2006 International Geosciences Conference and Exhibition, p. 11.
Bogie, I., Lawless, J.V., 2000. Application of mineral deposit concepts to geothermal exploration. In: Proceedings of the
World Geothermal Congress 2000, Kyushu-Tohoku, Japan, pp. 10031009.
Bogie, I., Mackenzie, K.M., 1998. The application of a volcanic facies model to an andesitic stratovolcano hosted geother-
mal system at Wayang Windu, Java Indonesia. In: Proceedings of the Twentieth New Zealand Geothermal Workshop,
University of Auckland, pp. 265270.
Bogie, I., White, P.W., Lawless, J.V., 2003. Chalcedony within low-sulphidation epithermal gold ore as an indicator
of decompressive boiling. In: Proceedings of the AusIMM 2002 Annual Conference, Auckland, New Zealand, pp.
173178.
Browne, P.R.L., 1978. Hydrothermal alteration in active geothermal fields. Ann. Rev. Earth Planet Sci. 6, 229250.
Budiardjo, B., 1992. Petrographic study of cores and cuttings from drillhole WWD-1. Geothermal Diploma Project Report
92.04. Engineering Library, University of Auckland, New Zealand, 43 pp.
Dam, M.A.C., 1994. The late quaternary evolution of the Bandung Basin, West-Java, Indonesia. Dissertation, Free
University of Amsterdam, 252 pp.
Ellis, A.J., Mahon, W.A.J., 1977. Chemistry and Geothermal Systems. Academic Press, New York, NY, USA, 392 pp.
Fournier, R.O., Truesdell, A.H., 1973. An empirical NaKCa geothermometer for natural waters. Geochim. et Cos-
mochim. Acta 37, 12551275.
Ganda, S., Hantono, D., 1992. Alteration mineralogy of the Wayang Windu geothermal field, West Java Indonesia.
In: Kharaka, Y.K., Maest, A.S. (Eds.), Water-Rock Interaction. Balkema, Rotterdam, The Netherlands, pp. 1405
1409.
Ganda, S., Hantono, D., Sunaryo, D., 1992. Alteration mineralogy of the Wayang Windu geothermal field, West Java
Indonesia. In: Proceedings of the Twenty-first Annual Scientific Meeting of the Indonesian Association of Geologists,
Yogyakarta, pp. 309314.
Gorbatov, A., Kennett, B.L.N., 2003. Joint bulk-sound and shear tomography for Western Pacific subduction zones. Earth
Planet. Sci. Lett. 210, 527543.
Hadi, J., Harrison, C., Keller, J., Rejeki, S., 2005. Overview of Darajat reservoir characterization; a volcanic hosted
reservoir. In: Proceedings of the World Geothermal Congress 2005, Antalya, Turkey, p. 11.
Healy, J., Mahon, W.A.J., 1982. Kawah Kamojang geothermal field, West Java Indonesia. In: Proceedings of the Pacific
Geothermal Conference Incorporating the 4th New Zealand Geothermal Workshop, vol. 2, University of Auckland,
pp. 313319.
Hochstein, M.P., Sudarman, S., 2008. History of geothermal exploration in Indonesia from 1970 to 2000. Geothermics
(this issue).
Hulen, J.B., Anderson, T.D., 1998. The Awibengkok Indonesia, geothermal research project. In: Proceedings of the
Twenty-third Workshop on Geothermal Reservoir Engineering, Stanford University, Stanford, CA, USA, pp. 256
263.
Layman, E.B., Agus, I., Warsa, S., 2002. The Dieng geothermal resource, Central Java. Indonesia. Geothermal Resources
Council Transactions 26, 573580.
Layman, E.B., Soemarinda, S., 2003. The Patuha vapour-dominated resource, West Java Indonesia. In: Proceedings of the
Twenty-eighth Workshop on Geothermal Reservoir Engineering, Stanford University, Stanford, CA, USA, pp. 5665.
Lovelock, B.G., Cope, D.M., Baltasar, A.J., 1982. Hydrogeochemical model of the Tongonan geothermal field. In:
Proceedings of the Fourth New Zealand Geothermal Workshop, University of Auckland, pp. 259264.
I. Bogie et al. / Geothermics 37 (2008) 347365 365

Moore, J.N., Allis, R., Renner, J.L., Mildenhall, D., McCulloch, J., 2002. Petrological evidence for boiling to dryness in the
Karaha-Telega Bodas geothermal system Indonesia. In: Proceedings of the Twenty-seventh Workshop on Geothermal
Reservoir Engineering, Stanford University, Stanford, CA, USA, pp. 223232.
Moore, J.N., Christenson, B., Browne, P.R.L., Lutz, S.J., 2004. The mineralogic consequences and behavior of descending
acid-sulfate waters: an example from the Karaha-Telaga Bodas geothermal system Indonesia. Can. Mineral. 42,
14831499.
Murakami, H., Kato, Y., Akutsu, N., 2000. Construction of the largest geothermal power plant for Wayang Windu project
Indonesia. In: Proceedings of the World Geothermal Congress 2000, Kyushu-Tohoku, Japan, pp. 32393244.
Nemcok, M., Moore, J.N., Christensen, C., Allis, R., Powell, T., Murray, B., Nash, G., 2007. Controls on the Karaha-Telaga
Bodas geothermal reservoir Indonesia. Geothermics 36, 946.
Newcomb, K.R., McCann, W.R., 1987. Seismic history and seismotectonics of the Sunda arc. J. Geophys. Res. 92,
421439.
Nurruhliati, R., 1996. Hydrothermal alteration of well WWR-SH, Wayang Windu geothermal field, West Java, Indonesia.
Geothermal Diploma Project Report 96.18, University of Auckland, New Zealand, 52 pp.
Raharjo, I., Wannamaker, P., Allis, R., Chapman, D., 2002. Magnetotelluric interpretation of the Karaha Bodas geothermal
field Indonesia. In: Proceedings of the Twenty-seventh Workshop on Geothermal Reservoir Engineering, Stanford
University, Stanford, CA, USA, pp. 388394.
Reid, M., 2004. Massive collapse of volcano edifices triggered by hydrothermal pressurization. Geology 32, 373376.
Reyes, A.G., Giggenbach, W.F., Saleras, J.R.M., Salonga, N.D., Vergara, M.C., 1993. Petrology and geochemistry of Alto
Peak, a vapour-cored hydrothermal system, Leyte Province Philippines. Geothermics 22, 479519.
Stoffregren, R.E., Alpers, C.N., 1987. Woodhousite and svanbergite in hydrothermal ore deposits: products of apatite
destruction during advanced argillic alteration. Can. Mineral. 25, 201211.
Sudarman, S., Pujianto, R., Budiarjo, B., 1986. The Gunung Wayang Windu geothermal area in West Java. In: Proceedings
of the Fifteenth Annual Convention of the Indonesian Petroleum Association, pp. 141153.
Suminar, A., Molling, P., Rohrs, D., 2003. Geochemical contributions to a conceptual model of Wayang Windu field
Indonesia. Geochimica et Cosmochimica Acta Supplement 67 (18), 454 (Abstract).
Thaysa, A., 2003. Hydrothermal alteration and geochemistry of geothermal fluids study of the Wayang Windu field, West
Java, Indonesia. Final Project Report, Department of Geology, Faculty of Sciences and Technology, Bandung Institute
of Technology, Indonesia, 74 pp.
Utami, P., 2000. Characteristics of the Kamojang geothermal reservoir (West Java) as revealed by its hydrothermal alter-
ation mineralogy. In: Proceedings of the World Geothermal Congress 2000, Kyushu-Tohoku, Japan, pp. 19211926.
Whittaker, J.M., Muller, R.D., Sdrolias, M., Heine, C., 2007. Sunda-Java trench kinematics, slab window formation and
overriding plate deformation since the Cretaceous. Earth Planet. Sci. Lett. 225, 445457.
Wibowo, H., 2006. Spatial data analysis and integration for regional-scale geothermal prospectivity mapping, West Java,
Indonesia. Msc Thesis. International Institute for Geo-Information Science and Earth Observation, Enschede, The
Netherlands, 94 pp.

Das könnte Ihnen auch gefallen