You are on page 1of 47

Elsevier Editorial System(tm) for Geothermics

Manuscript Draft

Manuscript Number: GEOT-D-08-00009

Title: Overview of the Wayang Windu Geothermal Field, West Java, Indonesia

Article Type: Indonesian Geothermal Fields

Section/Category:

Keywords: Geothermal reservoirs; Vapor-dominated; Liquid dominated; Wayang Windu; Java; Indonesia

Corresponding Author: Dr Ian Bogie,

Corresponding Author's Institution:

First Author: Ian Bogie

Order of Authors: Ian Bogie; Yudi I Kusumah; Merry C Wisnandary

Manuscript Region of Origin:

Abstract: The Wayang Windu geothermal field, West Java, Indonesia, is interpreted to be transitional
between vapor-dominated and liquid-dominated conditions with four coalesced fluid upwelling centers that
generally become younger and more liquid-dominated towards the south. Two of these centers 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 pre-exploitation
pressure versus elevation conditions were the same throughout the deep liquid-dominated reservoir it is
likely to be continuous across the field.
The deep liquid reservoir is overlain by three separate vapor-dominated reservoirs. The northernmost is the
largest as it is coalesced over two separate fluid upwelling centers. Its low gas content, size, prolonged
productivity and constant pressure at a given elevation preclude it from being a parasitic steam zone.
Mineralogical relationships demonstrate that the northern vapor zone was originally liquid-dominated with its
water level at an elevation as high as 1700 m asl. Subsequent boil off may reflect low recharge rates due to
hydrological isolation at depth. To the south, the vapor-dominated reservoirs decrease in thickness and are
characterized by progressively higher pressures, temperatures and gas contents, suggesting that the
southernmost vapor-dominated zone is the youngest and that these reservoirs become increasing older
toward the north.
Cover Letter

Dear Sue,

Attached are seven files corresponding to the Bogie et al. manuscript.


The paper is for the special issue on Indonesia and has been handled
outside EES.

Title of the manuscript: Overview of the Wayang Windu Geothermal Field,


West Java, Indonesia

Authors: Ian Bogie, Yudi Indra Kusumah and Merry C. Wisnandary

Corresponding author: Ian Bogie

Mailing address:
Ian Bogie
Sinclair, Knight Merz Ltd
PO Box 9806, Newmarket
Auckland, New Zealand

Other contact information:


Tel.: +64 9 913 8900; fax: +64 9 913 8901 E-mail address:
ibogie@skm.co.nz

Manuscript received on:16 April 2007


Manuscript accepted on 20 March 2008

The file labeled GEOT-D-08-Bogie-text.doc has 38 pages of text,


including two tables and the figure captions. The paper has a total of
six figures, all in JPEG format.

Note that I have already sent an e-mail message and a letter to Ian
Bogie informing him of the acceptance.

Please let me know if you have any questions about the manuscript or
the attached files.

Best regards,
Marcelo
****************************************************
Marcelo J. Lippmann, Ph.D.
Editor, Geothermics Tel: 1-510-486-5035
Earth Sciences Division 90R1116 Fax: 1-510-486-5686
Lawrence Berkeley National Laboratory
Berkeley, CA 94720-8126 USA e-mail: mjlippmann@lbl.gov
* Manuscript

GEOT-D-08-Bogie

Overview of the Wayang Windu Geothermal Field, West Java, Indonesia

Ian Bogie1*, Yudi Indra Kusumah2 and Merry C. Wisnandary2

1. Sinclair, Knight Merz Ltd, PO Box 9806, Newmarket, Auckland, New Zealand.
2. Mandala Nusantara Ltd., Wisma Mulia 50th Floor, Jl. Jend. Gatot Subroto no. 42,
Jakarta 12710, Indonesia.

Received: 16 April 2007; accepted 20 March 2008

Abstract

The Wayang Windu geothermal field, West Java, Indonesia, is interpreted to be

transitional between vapor-dominated and liquid-dominated conditions with four

coalesced fluid upwelling centers that generally become younger and more liquid-

dominated towards the south. Two of these centers 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 pre-exploitation pressure versus elevation conditions were

the same throughout the deep liquid-dominated reservoir it is likely to be continuous

across the field.

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

1
The deep liquid reservoir is overlain by three separate vapor-dominated reservoirs. The

northernmost is the largest as it is coalesced over two separate fluid upwelling centers. Its

low gas content, size, prolonged productivity and constant pressure at a given elevation

preclude it from being a parasitic steam zone. Mineralogical relationships demonstrate

that the northern vapor zone was originally liquid-dominated with its water level at an

elevation as high as 1700 m asl. Subsequent boil off may reflect low recharge rates due

to hydrological isolation at depth. To the south, the vapor-dominated reservoirs decrease

in thickness and are characterized by progressively higher pressures, temperatures and

gas contents, suggesting that the southernmost vapor-dominated zone is the youngest and

that these reservoirs become increasing older toward the north.

Keywords: Geothermal reservoirs; Vapor-dominated; Liquid dominated; Wayang Windu;

Java; Indonesia

1. Introduction Regional 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., 2003),

Kamojang (Utami, 2000), Karaha-Telaga Bodas (Moore et al., 2002, 2004), Papandayan

(Wibowo, 2006), Patuha (Layman and Soemarinda, 2003), Tampomas (Wibowo, 2006),

2
and Tangkuban Perahu (Wibowo, 2006). These fields lie within andesitic, volcanic

highlands formed by a concentration of volcanic centers in this part of the Sunda Arc.

The city of Bandung is located in a basin (Dam, 1994) near the center of the volcanic

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 Australian-

Indian 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, 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 sidetracked to the

west, Fig. 2) was drilled by Pertamina to the west of the saddle between Gunung

3
Wayang1 and 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 vapor-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

liquid-vapor type geothermal systems. A 600-m deep slim hole (MSH-1) drilled by

Pertamina in 1993-1994 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 surficial hydrothermal alteration

The most intense surficial hydrothermal activity occurs adjacent to the small G. Wayang

and G. Windu volcanic centers (Fig. 3). Fumaroles, steaming and altered ground, and

acid-sulfate springs occur in the Wayang thermal area, which lies within a sector

collapse, with the current peak of G. Wayang representing an eastern remnant of a much

larger volcanic center, originally located to the west along the axis of alignment of other

small volcanic centers. The matrix of the debris flow from the sector collapse consists of

hydrothermal clay, suggesting slope failure related to the weakening of the volcanic

deposits by hydrothermal alteration and the increase of pore pressures with heating (Reid,

2004). A radiocarbon date of 7450 110 years was obtained from peat sampled 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.

1
The abbreviation G. will be used for Gunung (mountain) from here on.

4
Smaller areas of altered ground with acid-sulfate 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 centers (Fig. 3). The springs have temperatures ranging from 25 to 66C 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 groundwater, 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 fresh 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 Wayang Windu 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

5
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 initially productive wells were

restricted to sites immediately beneath and southwest of the young volcanic centers.

Wells drilled north of G. Bedil encountered a shallower, two-phase, vapor-dominated

reservoir than the one found in the first well (WWA-1). These new wells confirmed the

results from slim hole MSH-1 in the north and indicated a vapor-dominated zone 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 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).

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).

6
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 Wayang Windu project in 2001. The

poorly productive deep MBD-1 vertical well was sidetracked to intersect the shallow,

vapor-dominated zone and came to be the largest producer in the field at the time. That

was followed by workovers of other wells on the MBD drill pad, which reduced

production due to the installation of tiebacks (whereby the production casing is extended

to the surface reducing the internal diameter of the cased section of the well). WWQ-3,

another deep well, was also sidetracked, 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 vapor-dominated zones, indicating that they are major, sustainable

productive reservoirs rather than limited parasitic ones.

7
The Wayang Windu 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 reservoir modeling studies (Asrizal et al., 2006) is

eminently achievable.

3.1. Extent of the Wayang Windu geothermal field

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), reflecting the presence of 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 system (Fig. 3).

Southwest of G. Windu, productive wells have been drilled on the basis of earlier

Schlumberger and MT resistivity survey results, whereas later re-interpretation of new

8
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 reached higher elevations. The conductor in the south now

appears to be too impermeable for the alteration mineralogy to re-equilibrate and allow

formation 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.

Combining new well data with those of recent MT surveys, which have a greater station

density and have yielded better quality information, 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 E-W

direction and to extend approximately 14 km to the south beneath a series of aligned,

small volcanic centers, where it narrows down to approximately 2 km across. This gives

an overall potential resource area of approximately 40 km2. However, since core 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 center, as will be

discussed below. This feature is found in other Javan geothermal fields. Layman et al.

(2002) identified three centers at Dieng, multiple centers (referred to as cells) were

9
suggested for Karaha-Telega Bodas (Nemok et al., 2007) and two geothermal centers

were recognized at Awibengkok (Hulen and Anderson, 1998).

4. Geology of the field

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,

which define a series of overlapping andesitic piles. Their cross section has been

extended in Fig. 4; its the trace is 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 centers 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 Soemarindo, 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. 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 to

10
interlayered smectite-illite. As these clays are usually found at temperatures below

200C (Anderson et al., 2000), we suggest the beds must have very low permeability,

since measured temperatures at the corresponding depths (> 300C in some instances) are

much greater than the typical stability limit of the clay.

Gunung 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

volcano summit (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).

Gunung Bedil, the next volcanic center to the south, was dated at 0.19 Ma, and G. Windu,

the southernmost volcanic center at 0.10 Ma. A 0.49 Ma date for G. Wayang breaks the

trend of having younger centers towards the south. While the other young volcanic

centers are well preserved and samples for dating were obtained from their youngest

parts, G. Wayang has undergone sector collapse, and it is likely that the dated sample

from that eruptive center was taken from a much older part of the volcanic pile.

The geochemistry of the younger volcanic centers 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 volcanics. A

more extensive, but less intensive geochemical study of samples of older units collected

11
from 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 30-40 and 330340. In the north, along the

southern boundary of the Bandung Basin, further deformation results from movement

along the boundary fault. Gunung 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 rocks penetrated by the wells have highly localized structural permeability; with the

most permeable geologic structures following the regional 40 strike. As these structures

have trends similar to regional faults, then it is likely they are strike-slip faults, which

would tend to have lower permeabilities than normal faults because of shearing and

consequent rock comminution (i.e. trituration). 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 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

12
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 the Bandung Basin.

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 that 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 feed the springs.

Possible condensate aquifers, not connected to the regional groundwaters, are indicated

by high porosities measured by using a downhole magnetic resonance imaging logging

tool; these aquifers are characterized by strong alteration to kaolinite, calcite, quartz and

anhydrite. The conductor itself is made up mainly of an argillic assemblage dominated

13
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 still deeper. 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 well WWA-4. In the deep liquid

reservoir, wairakite and prehnite, along with epidote, are common as alteration and vein

minerals, with less frequent adularia.

Other than a generally prograde transition, with illite-smectite and corrensite detected

shallow in the vapor-dominated reservoirs, and prehnite in the lower parts of the deep

liquid reservoir, there is no clear difference in the hydrothermal alteration of the liquid-

and vapor-dominated reservoirs. Wairakite, however, is common in the vapor-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 caprock of both the

vapor-dominated and deep liquid reservoirs. Minerals of this assemblage include

14
pyrophyllite, diaspore, woodhouseite and dickite, with accompanying quartz, anhydrite

and pyrite. As these hydrothermal minerals are created under high-temperature, acid

conditions, and considering that the reservoir pH is now near neutral, and has

temperatures below those at which these minerals formed (Reyes et al., 1993), we

consider this deep advanced argillic alteration to be relict. This may possibly reflect the

earlier presence of acidic condensed magmatic volatiles, particularly since woodhouseite

(CaAl3PO4SO4(OH)6; Stoffengren 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 up to 1330 m asl. The shallowest appearance of this mineral is above

the vapor-dominated reservoir, and all its first occurrences in well samples are above the

deep liquid reservoir (Figs. 4 and 6). Since generally in this type of geothermal reservoir

epidote forms at temperatures above 240C under near neutral pH conditions (Browne,

1978), we infer that the water level was previously higher. If the geothermal system is

related to recent volcanism, then it is no older than 0.23 Ma. In that case, the lowering of

the water level could not have been 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 that the 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 centers, the top of

15
the conductor is at 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 tops of the Waringin volcanic unit (Fig. 4)

and of the vapor-dominated reservoir (Fig. 6) reflecting, perhaps, a porosity variation. In

the northern part of the field, the Malabar volcanic unit consists mainly of lavas whose

average porosity is ~1% (Asrizal et al., 2006). The underlying Waringin unit consists

mainly of lapilli tuffs with 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 geothermal system.

The permeability associated with vertical faults that cut these 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 volcanic unit contains a higher proportion of lava flows in its

upper parts. The overlying Pangalengan unit, which contains many ash deposits and fine-

grained sediments, could have formed the original cap in this part the field, and thus

might have influenced the distribution of epidote. However, a plot of the depth of the

first appearance of epidote has a shape similar to that of the 300C isotherm, suggesting

instead that the initial temperature distribution may outweigh the influence of porosity

variations in the south.

16
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 centers 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 higher resistivity areas in the conductor reflect near

present conditions, whereas its base was defined during an early stage in the development

of the system. Similar higher resistivities 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 vapor-dominated reservoirs.

North of the wells drilled at Wayang Windu the main conductor is found at elevations up

to 1800 m asl. In this part of the system it is possible that the water level may have

reached these highs during the history of the field. Moreover, the lower conductivities

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

17
above the vapor-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 vapor-dominated resource, at least in its upper parts. This may explain why this

steam zone is best developed in the north where there is the steepest drop off in the depth

of the margin of the conductor, forming a dome which encloses the vapor-dominated

reservoir.

Hydrothermal alteration in two-phase, vapor-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 240C (Bogie et al., 2002), 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.

18
Fluid inclusion work has been limited by the amount of suitable sample material.

Abrenica (2007) reports homogenization temperatures between 228 and 255C, with the

mode at 235C, in primary fluid inclusions in quartz from a vein at 590 m asl in well

MBD-5; the current estimated temperature is 246C (the downhole logging tool did not

reach this depth). Melting point measurements for these inclusions indicate salinities of

0.53 to 1.05 wt% NaCl equivalent (Abrenica, 2007), reflecting 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 ~2 wt% NaCl equivalent. Therefore, these inclusion

results likely reflect early liquid reservoir conditions.

Secondary fluid inclusions in that quartz sample are vapor-rich and have higher

vapor/liquid ratios in successive generations, consistent with the presence of increasing

vapor-dominated conditions with time. Homogenization temperatures of the secondary

inclusions range from 241 to 334C, but since it is unlikely that a single-phase fluid was

trapped, they may not reflect temperatures prevailing at the time of their formation.

Primary inclusions in calcite from a sample collected two meters below the MBD-5

quartz sample (i.e., at 592 m asl) gave higher homogenization temperatures in the 264-

295C range (Abrenica, 2007), with the mode at 275C. Like the secondary inclusions in

quartz, these calcite-hosted inclusions are vapor-rich and have increasing vapor/liquid

ratios in successive generations. and homogenization temperatures may not be reflecting

trapping temperatures.

19
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 vapor-dominated conditions now prevail at that depth, it is inferred that the

liquid level was previously higher and that the epidote is relict.

5. Reservoir characteristics and geochemistry

The Wayang Windu geothermal resource has a deep, hot, neutral pH, liquid reservoir

that, in the area drilled, is overlain by perched vapor-dominated, two-phase reservoirs

(Fig. 6). Throughout the field within the deep liquid reservoir, pressures and temperatures

versus elevation are similar (Table 2; Fig. 5). In the north its top is at 400 m asl, and it

deepens towards the south where it is found near sea level elevation. Since under natural-

state (pre-exploitation) conditions the reservoir had almost the same vertical pressure

distribution throughout, it can be considered to be a continuous body, and since it was

near pressure equilibrium there was 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 regions 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 vapor-dominated reservoir is

more limited. In the recharge areas the shallowest well feed zones from the deep liquid

20
reservoir are cooler than elsewhere in that reservoir, but without actual temperature

inversions deeper in the well, indicating some limited ingress of meteoric waters into the

shallow zones.

Overlying the deep liquid-dominated zone we find two-phase vapor-dominated

reservoirs. The largest in the north appears to be coalesced over two fluid upwelling

centers (associated with Puncak Besar and G. Gambung), while the two farther south

(associated with G. Wayang and G. Windu) appear to be separate. In other words, the

Wayang Windu geothermal system seems to present three vapor-dominated reservoirs

that are located over four fluid upwelling centers. The characteristics of the vapor-

dominated zones change progressively towards the south; i.e. their pressures,

temperatures and gas contents increase, their thicknesses decrease, and are found at

greater depths.

The degree of communication between the vapor-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 centers are also temporally related to the fluid upwelling

centers) 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 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 vapor-dominated reservoirs may reflect the

21
distribution of impermeable, regional ash deposits that restrict vertical permeability (and

fluid flow), 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), while MBE-2 and the major

producing wells on the MBA and MBD drill pads are associated with the southern dome.

In the north, a fourth area related to the northernmost domed structure in the base of the

conductor has now been drilled and found to be productive [well MBB-1 (see Figs. 2-4)

produces enough steam to generate 23 MWe].

The four possible fluid upwelling centers areas in the field are spatially associated with

four eruptive centers (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 (e.g. 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.

22
The waters of the deep reservoirs in the G. Gambung3 and G. Wayang areas have similar

Cl/B ratios than those from the G. Windu region, but are much more saline and have

lower gas contents suggestive of boiling at high temperatures, with very limited mixing

with groundwaters (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 fluid upwelling centers (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 centers, 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.

7. Discussion

As the pre-exploitation pressure-versus-elevation in the deep Wayang Windu liquid

reservoir was the same throughout the drilled area and the deep reservoir was under-

pressured with respect to local cold-water aquifers, the hydrology of the system cannot be

interpreted in terms of liquid upflows and outflows typical of most geothermal fields

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 vapor-dominated

reservoirs, and the limited depth range where vapor-dominated conditions prevail at

3
Fluid chemistry data on the deep liquid reservoir is lacking for the northernmost wells
as they are too shallow (i.e. they were completed above it). Thus, all the deep liquid
geochemistry data come from well MBE-2.

23
Wayang Windu also mean that the field cannot be strictly compared to the solely vapor-

dominated systems of Darajat or Kamojang. Wayang Windu must therefore be regarded

as a new type of geothermal field, transitional between liquid-dominated and vapor

dominated. 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 vapor-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 vapor-dominated

conditions represented in the geothermal field.

The Wayang Windu system is largely sealed off from surrounding and overlying cold-

water aquifers, but still receives deep heat and fluid recharge from possibly four fluid

upwelling centers that are inferred to be progressively younger to the south. Vapor-

dominated reservoirs have developed over the fluid upwelling centers, with a vapor-

dominated reservoir in the north coalesced over the fluid upwelling centers associated

with the Puncak Besar and G. Gambung volcanic centers. Two possibly separate vapor-

dominated reservoirs further south, lie above separate fluid upwelling centers 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 multiphase

intrusion, which would be the ultimate heat source. The diorite xenoliths observed in the

younger volcanics may be from the older parts of this intrusion.

24
The locations of the four fluid upwelling centers 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 centers can be related to

the fluid upwelling centers, 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 geothermal system.

The deep magmatic degassing is now heating and partially recharging a large exploitable

reservoir. As this deep recharge is composed of water vapor 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. 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.

25
The Wayang Windu field is part of a cluster of Indonesian high-temperature geothermal

fields, which include the vapor-dominated reservoirs of Darajat and Kamojang that may

represent the end point of the liquid-to-vapor transition. Vapor-capped reservoirs at

Patuha (Layman and Soemarinda, 2003) and Karaha Telaga Bodas (Moore et al., 2002,

2004) exhibit magmatic vapor 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

the 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 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

26
alteration and activity have not been reported from these areas and it is more difficult to

relate them to changes in the hydrology of the system.

This concentration of vapor-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 perpendicular 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 rollback (Whittaker et al., 2007) is taking place all along the Javan part of the

Sunda Arc because of the subduction of the old (95 - 135 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 extension in the overlying

plate, which may be responsible for the large number of volcanic centers and associated

geothermal fields in the Bandung area. The rollback may also be responsible for the

occurrence of multiple fluid upwelling centers in Javan geothermal fields, as the location

27
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 vapor-dominated reservoirs.

Acknowledgements

The authors wish to gratefully acknowledge the permission and support of the

management of Mandala Nusantara Ltd. 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 Greg Bignall, Sabodh Garg, Marcelo

Lippmann and Joe Moore. 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

28
Engineering Department, Faculty of Engineering, Gadjah Mada University, Yogyakarta,

Indonesia, 151 pp.

Allis, R., 2000. Insights on the formation of vapor-dominated systems. In: Proceedings

of the World Geothermal Congress 2000, Kyushu-Tohoku, Japan, pp. 2489-2496.

Allis, R., Moore, J.N., McCulloch, J., Petty, S., DeRocher, T., 2000, "Karaha-Telaga

Bodas, Indonesia: A Partially Vapor-Dominated Geothermal Systems, Geothermal

Resources Council Transaction 24, 217-222.

Alzwar, M, Akbar, N., Bachri, S., 1992. Geology of the Garut and Pameungpeuk

Quadrangle, Jawa. Geologic map issued by Republic of Indonesia, Department of Mines

and Energy, Directorate General of Geology and Mineral Resources, Geological

Research and Development Centre.

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. 107-112.

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. 909-914.

29
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. 235-242.

Bandyopadhay, I., Juandi, D., Baroek, M., Hadi, J., Pramono, B., 2006. In situ stress

analysis in deviated wells using inversion method A case study from volcanic

pyroclastics of Wayang Windu Field, Java, Indonesia. In: Proceedings of the Jakarta

2006 International Geosciences Conference and Exhibition, 11 pp.

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. 1003 1009.

Bogie, I., Mackenzie, K.M., 1998. The application of a volcanic facies model to an

andesitic stratovolcano hosted geothermal system at Wayang Windu, Java, Indonesia. In:

Proceedings of the Twentieth New Zealand Geothermal Workshop, University of

Auckland, pp. 265- 270.

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. 173-178.

Browne, P.R.L., 1978. Hydrothermal alteration in active geothermal fields. Ann. Rev.

Earth Planet Sci. 6, 229-250.

30
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, The Netherlands, 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 Na-K-Ca geothermometer for natural

waters. Geochim. et Cosmochim. Acta 37, 1255-1275.

Giggenbach, W.F., 1992. Magma degassing and mineral deposition in hydrothermal

systems along convergent plate boundaries. Economic Geology 87, 1927-1944.

Ganda, S., Hantono, D., 1992. Alteration mineralogy of the Wayang Windu geothermal

field, West Java, Indonesia. In: Kharaka, Y.K. and 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. 309-314.

Gorbatov, A., Kennett, B.L.N., 2003. Joint bulk-sound and shear tomography for

Western Pacific subduction zones. Earth and Planetary Science Letters 210, 527-543.

31
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, 11 pp.

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, University of Auckland, Vol. 2, pp. 313-319.

Hochstein, M.P., Sudarman, S., 2007. 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 theTwenty-third Workshop on Geothermal Reservoir

Engineering, Stanford University, Stanford, CA, USA, pp. 256-263.

Layman, E.B., Soemarinda, S., 2003. The Patuha vapor-dominated resource, West Java,

Indonesia. In: Proceedings of the Twenty-Eighth Workshop on Geothermal Reservoir

Engineering, Stanford University, Stanford, CA, USA, pp. 56-65.

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. 259-264.

32
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. 223-232.

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. Canadian Mineralogist, 42, 1483-

1499.

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. 3239-3244.

Nemok, M., Moore, J.N., Christensen, C., Allis, R., Powell, T., Murray, B., Nash, G.,

2007. Controls on the KarahaTelaga Bodas geothermal reservoir, Indonesia.

Geothermics 36, 9 46.

Newcomb, K R., McCann, W.R., 1987. Seismic history and seismotectonics of the

Sunda arc. Journal of Geophysical Research 92, 421-439.

33
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. 388-394.

Reid, M, 2004. Massive collapse of volcano edifices triggered by hydrothermal

pressurization. Geology 32, 373-376.

Reyes, A.G., Giggenbach W.F., Saleras, J.R.M., Salonga, N.D., Vergara, M.C., 1993.

Petrology and geochemistry of Alto Peak, a vapor-cored hydrothermal system, Leyte

Province, Philippines. Geothermics 22, 479-519.

Stoffregren R.E., Alpers C.N., 1987. Woodhousite and svanbergite in hydrothermal ore

deposits: products of apatite destruction during advanced argillic alteration. Canadian

Mineralogist 25, 201-211.

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. 141-153.

34
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), Abstract, pp. 454.

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 alteration mineralogy. In: Proceedings of the World

Geothermal Congress 2000, Kyushu-Tohoku, Japan, pp.1921-1926.

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 and Planetary Science Letters 225, 445-457.

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.

35
Table 1
Alteration mineralogy (Possible relict phases shown in italics)

Location Initial Alteration Over Print


Opal, cristobalite, kaolinite, alunite,
Above Conductor natro-alunite and sulphur
Smectite, illite-smectite, quartz, chlorite,
albite, calcite, pyrite, heulandite,
mordenite, clinoptlolite, stilbite, Kaolinite, anhydrite,
Conductor analcime, laumonite calcite, quartz
Quartz, chlorite, calcite, albite, pyrite,
illite-smectite, corrensite, epidote, illite,
Vapor-Dominated Reservoirs chalcedony, wairakite Anhydrite, calcite, pyrite

Amphibole

Pyrophyllite, diaspore, quartz, anhydrite


Quartz, chlorite, illite, pyrite, wairakite,
epidote, prehnite, adularia, albite,
tourmaline
Deep Liquid Reservoir
Dickite, pyrophyllite, quartz,
woodhouseite, pyrite

Amphibole, orthoclase, magnetite

Diopside, oligoclase, magnetite

36
Table 2. Geochemical and pressure-temperature properties of Wayang Windu reservoir
areas prior to production

Deep Liquid Reservoir Two-Phase Reservoirs


Area
Measured Measured
Reservoir Cl 3 NaKCa Temp.1 NCG 2 NCG2 Temp. 3 Pressure 3 Elevation
(Wt
(ppm) (C) (Wt %) %) (C) (bar) (masl)
Puncak
4 4 4 4 4 4
Besar ? - 1120
0.3 - 0.6 - 400 -
Gambung 12, 000-13,000 295 - 300 0.6 2.6 250 - 260 35 - 45 1100
0.5 -
Wayang 12,00013,000 295 - 308 0.6 2 - 4.5 255 - 267 50 - 55 200 - 700
Windu 6000-8000 285 - 300 3.5 10 260 - 290 80 -85 80 - 400

Notes
1
NaKCa geothermometer of Fournier and Truesdell (1973).
2
Non-condensable gases in weight percent.
3
Prior to production.
4
Well MBB-1 drilled in this area produces 20 MWe on initial discharge.
It did not penetrate into the deep liquid reservoir and fully stabilised gas,
temperature and pressure data are not yet available.

37
Figure captions

Fig. 1: The distribution of Quaternary volcanic rocks and high-temperature geothermal

fields in West Java, Indonesia.

Fig. 2: Topographic map of the drilled portion of the Wayang Windu geothermal field;

contour interval: 100 meters; bold contour: 2000 m asl. Also shown are the locations of

drill pads, well tracks and of section A-B described in Figs. 4 and 6.

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.

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).

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.

Fig. 6: Section across the Wayang Windu geothermal field showing well tracks,

isotherms and the known location of the tops of the vapor-dominated and deep liquid

reservoirs (section location is given in Fig. 2).

38
Figure
Click here to download high resolution image
Figure
Click here to download high resolution image
Figure
Click here to download high resolution image
Figure
Click here to download high resolution image
Figure
Click here to download high resolution image
Figure
Click here to download high resolution image

You might also like