Vous êtes sur la page 1sur 15

Available online at www.sciencedirect.

com

Geothermics 37 (2008) 332–346

Reservoir management at Awibengkok geothermal


field, West Java, Indonesia
Jorge A. Acuña ∗ , James Stimac,
Lutfhie Sirad-Azwar, Riza Glorius Pasikki
Chevron Geothermal Salak Ltd., 11th Floor Senayan I, Jalan Asia Afrika, Jakarta, Indonesia
Received 8 October 2007; accepted 5 February 2008
Available online 2 April 2008

Abstract
The Awibengkok geothermal field, also known as Salak, is the largest developed geothermal resource in
Indonesia, currently sustaining 377 MW of electrical generation. It is a water-dominated, naturally fractured
reservoir with benign fluid chemistry. A very large amount of produced brine is injected along the margins
of the proven reservoir. After 13 years of continuous operation, production levels have been maintained at or
above nominal turbine capacity through periodic make-up drilling and field management. The two main chal-
lenges have been taking advantage of the changing reservoir thermodynamic conditions and managing injec-
tion. Some innovations in well design and drilling procedures include sizing up the production casing from
9 58 in. to 13 38 in. with 13 38 in. and 16 in. tie-backs, and drilling shallow, relatively high-angle (55–60◦ inclina-
tion) wells to maximize production from the steam cap. Well deliverability predictions have been improved
by combining well production history and downhole measurements to construct wellbore hydraulic models.
Changes in injection strategy were made periodically to optimize heat recovery based on well perfor-
mance and trends in well chloride concentration and enthalpy. Thermally stable tracers have been used to
better understand inter-well connectivity. Make-up drilling has been managed to take advantage of evolving
thermodynamic conditions in different parts of the field. Reservoir management has required continuously
adapting production and injection strategies in response to reservoir evolution. This is only possible through
maintaining robust data gathering and monitoring programs. Efforts are currently underway to evaluate
expanding the area of production by moving injection deeper or to more distal locations.
© 2008 Elsevier Ltd. All rights reserved.

Keywords: Salak; Awibengkok; Indonesia; Reservoir management; Injection management; Tracer testing; Chloride
monitoring; Well targeting and design; Drilling practices

∗ Corresponding author. Tel.: +62 21 5798 4579; fax: +62 21 5730 981.
E-mail address: jacuna@chevron.com (J.A. Acuña).

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


doi:10.1016/j.geothermics.2008.02.005
J.A. Acuña et al. / Geothermics 37 (2008) 332–346 333

1. Introduction

The Awibengkok geothermal field is located in West Java, 60 km south of Jakarta in Indonesia.
A map of the field location is shown in companion paper by Stimac et al. in this issue. It is the
largest developed geothermal field in the country and together with the Darajat field (Hoang et
al., 2005) constitutes the geothermal capacity operated by Chevron in Indonesia. The combined
capacity of these two projects makes the company the largest geothermal operator in the country.
Awibengkok, also called Salak, was originally discovered in the early 1980s by Unocal and the
western portion was developed to supply 110 MWe of generating capacity in 1994 (Murray et al.,
1995). Production was increased to a nominal capacity of 330 MWe in late 1997 (Soeparjadi et al.,
1998) by expanding production to the eastern portion of the proven reservoir. After the economic
crisis that hit Indonesia in 1997–99, the operating contract with the government of Indonesia
called for increased production from Awibengkok. As a result the field has operated at a target of
377 MWe since late 2002. The only periods of significant decreased production have been due to
maintenance of the generating facilities and landslides that destroyed part of the pipeline system
in March 2003. Full recovery from this event took approximately 5 months.
Here we describe major changes that have occurred in reservoir production during the first 13
years of operation and highlights resource management challenges that were overcome to maintain
full steam supply. The geologic framework and conceptual model of the field are described by
Stimac et al. (2008).

2. Reservoir configuration and characteristics

The Awibengkok geothermal system is associated with a cluster of densely forested andesitic
to rhyolitic volcanoes on the western flank of Gunung Salak (Stimac et al., 2008). The project
is located in the protected Halimun-Salak Forest. Directional wells were selected to maximize
operational efficiency and minimize the project footprint. Up to nine wells have been drilled from
individual pads resulting in deep and far reaching fluid conduits with wellheads and production
facilities located together at surface. As of late 2006, 75 wells have been drilled from 18 well
sites; 47 of these wells are producers and 18 are used for brine and plant condensate injection
(Fig. 1). The remainder (10 wells) are either idle or used for pressure monitoring.
Awibengkok is a water-dominated reservoir with initial temperatures ranging from 235 to
310 ◦ C. At initial conditions the reservoir had a liquid pressure profile up to the highest point
of the reservoir at approximately 560 m above sea level (masl) (See Fig. 2). The discovery well
suggested that a very thin steam cap was present at the top of the eastern reservoir (UGI, 1985).
The main feature of the reservoir geometry as described by Stimac et al. (2008) is a relatively
shallow reservoir top on the eastern side of the field defined by wells drilled from the Awi 1, 13
and 16 pads. The reservoir top is deeper on the hotter western and southern sides of the field as
defined by wells from Awi 7, 8, 9, 11, and 10, 14 and 15 pads, respectively. Fig. 2 shows the
pre-exploitation temperature profiles for wells in each side of the reservoir, the location of the
reservoir top can be inferred in these profiles as the contact between conductive and convective
well temperature gradients.
Outflows from the system to the north and southeast are defined by temperature and fluid
chemical trends of wells and the distribution of thermal features. Elevation of the chloride spring,
believed to be connected to the reservoir, is shown in Fig. 2.
A “cupola” or thimble-shaped reservoir top allowed the development of a relatively large steam
cap on the east. Shallow eastern wells that produced brine early in the life of the resource later
334 J.A. Acuña et al. / Geothermics 37 (2008) 332–346

Fig. 1. Location of power plants, main roads, well pads and trajectories at the Awibengkok geothermal field. The boundary
of commercial production is shown by the dashed line. Wells used for injection are shown as dashed lines; production
wells by the bold lines. Well numbers correspond to the order in which well pads were established and individual wells
were drilled (e.g. Awi 1-1, 1-2, Awi 2-1). Only wells mentioned in the text are labeled. Light sinuous lines correspond to
roads.

evolved to produce dry steam. Wells in the western sector, where the reservoir top is deeper, have
remained brine producers until today.
Sub-commercial conditions have been encountered to the south [Awi 10-2OH (OH: original
hole), Awi 14-4], northwest (Awi 12) and east (Awi 4) of the central reservoir, but reservoir
boundaries remain undefined to north. Recently new geophysical surveys showed that the low-
resistivity anomaly attributed to the smectite clay cap overlying the proven reservoir extended far
to the north and west (Stimac et al., 2008). These areas are consequently being evaluated for their
potential production or injection. Thus far, two delineation wells drilled to the west of the proven
reservoir in 2006 have encountered relatively low temperature and permeability.
Awibengkok is a naturally fractured reservoir that exhibits large inter-well permeability despite
the fact that matrix permeability measured on core samples is orders of magnitude smaller than
fracture permeability (Stimac et al., 2008). Permeability in such complex fracture networks is
controlled by highly connected, large fractures, while most fluid reserves are stored in smaller
fractures and lower permeability matrix.
Tracer tests using thermally stable chemicals (Rose et al., 2000) have shown return times of
only a few days for some wells (Gunderson et al., 2002). Some injection wells exhibit rapid
tracer breakthrough to specific producers and, based on temperature measurements, sometimes
J.A. Acuña et al. / Geothermics 37 (2008) 332–346 335

Fig. 2. Pre-exploitation pressure gradient (dashed line) in Awibengkok as determined by UGI (1985) and the elevation
of chloride spring. Also shown are the pre-exploitation temperature gradient for the hottest well in the field (Awi 9-2) in
Western Awibengkok and a typical Eastern well. These profiles show the conductive–convective contact associated to the
reservoir top in both sides of the field.

only to one feed zone in those wells. The observed pressure depletion rates throughout most of
the reservoir are similar, although some areas appear to be separated by semi-permeable barriers
that correspond primarily to zones of faulting and recent intrusion (Stimac et al., 2008). Faults
play an important role as flow barriers perpendicular to the fault traces, but enhanced fluid flow
is commonly observed parallel to the traces or at fault intersections. The interaction of these
structural features leads to the formation of compartments not totally sealed, but separate enough
to provide distinctive well behavior. This feature has been observed in other fields, such as Tiwi
(Sugiaman et al., 2004) and Bulalo (Abrigo et al., 2004), both in the Philippines.

3. Early 110 MW field development and resource management

When the first power plants were put on line in 1994, the size and characteristics of the
Awibengkok reservoir were not very well constrained. A central core of production was developed
around the early discoveries provided by wells from pads Awi 7 and 8 (See Fig. 1). An injection
area at relatively low elevation was required for the project, and Awi 9 and 10 pads were drilled at
distances between 1.5 and 1 km from the main production area for this purpose. These injection
locations proved to have higher well temperatures than the production areas of Awi 7 and 8.
Well Awi 9-1 had a deep reservoir top and modest permeability, and had to be stimulated with
injection and two acid jobs until its injectivity increased to acceptable levels. A second well,
Awi 9-2, was drilled, which also found very high temperatures and modest deep permeability.
Awi 10-1 was also hot and very productive while the second well, Awi 10-2OH, proved to be
almost impermeable and was later redrilled becoming Awi 10-2RD (RD: redrilled hole). A more
detailed description of the early development in Awibengkok can be found in Murray et al.
(1995).
336 J.A. Acuña et al. / Geothermics 37 (2008) 332–346

Before the field started operation it was debated whether to use wells Awi 9 and 10 as producers
(Murray et al., 1995), but it was decided to proceed with the original plan of utilizing them as
injectors due to financial and scheduling constraints on the project. The decision was based
primarily on their lower elevation and deep permeability. Although it was a difficult decision as
no operator would choose to inject into productive wells, it was made knowing that the effects of
injection can often be reversed by transforming injection wells into producers later in the life of
the resource.
The field started to produce 110 MW in 1994 from the west using seven producing wells located
on the Awi 7, 8 and 11 pads, and seven injection wells on the Awi 9 and 10 pads. During the early
period of exploitation from 1994 to 1997, rapid chemical breakthrough was observed in production
wells Awi 11-1 and Awi 11-2 (see Fig. 3). The primary source of this breakthrough was Awi 10-
1. During this period more than 250 kg/s of brine was injected into Awi 10-1 and the chloride
concentration in nearby Awi 11-1 and 11-2 wells jumped from the original pre-exploitation value
of approximately 6500 ppm to close to 8000 ppm. Chloride mass balance suggested that Awi 11-1
produced approximately 50% injectate, while Awi 11-2 about 40%. A tracer test done at Awi
10-1 indicated that injectate from this well returned to Awi 11-1 in 2.2 days or with a migration
rate of 450 m/day. The rapid connection between the two wells was attributed to entry locations
in the same permeable tuff formation known as the rhyodacite marker (Stimac et al., 2008). The
very rapid increase in chloride in produced brine can be related to a very short residence time of
the injectate in the reservoir, and this was taken as an indication that thermal breakthrough would
inevitably follow.
Injection was shifted from Awi 10-1 to Awi 10-2RD in early 1995, resulting in a small decline
of injectate content in Awi 11 wells. Termination of injection in Awi 10-1 and 10-2RD in late
1996 to early 1997 further reduced the amount of injectate produced by Awi 11. In 1997, Awi 11-1

Fig. 3. Reservoir chloride concentration histories for wells Awi 11-1 and 11-2. Both wells had initial chloride concentra-
tions of about 6500 ppm (see box to the left). Periods of injection into Awi 10-1 and 10-2RD indicated by the solid and
dashed lines are shown for reference. The abrupt change that occurred in 1994 was caused by the start of injection into
Awi 10-1, while the decline that begins in late 1996 is due to stopping injection into wells at pad Awi 10.
J.A. Acuña et al. / Geothermics 37 (2008) 332–346 337

Fig. 4. Reservoir chloride concentration histories for selected production wells at pad Awi 8. The changes are mostly due
to injection into wells at pad Awi-9. These production wells were impacted by injection into Awi 9 wells prior to start-up;
probably their initial concentrations were about 6500 ppm.

and 11-2 produced less than 10% injectate. These changes were estimated from the evolution of
chloride shown in Fig. 3.
Despite the injectate proportion being produced, the productivity of Awi 11 wells did not suf-
fer significantly. Enthalpy measurements as well as downhole and deliverability measurements
indicated minor to nil thermal decline. High injectate production rates interpreted from chloride
trends declined rapidly once injection in Awi 10 wells was halted. When Awi 10-1 and 10-2RD
were converted from injectors to producers in 1997, wells Awi 11-1 and Awi 11-2 produced
less than 20% injectate; chloride concentrations did not reach their 1996 levels again until 2002
(Fig. 3).
During this same period (1994–1997) all the separated brine from Awi 8 and Awi 7 wells was
injected into the Awi 9 wells. Awi 8 wells produced 30 to 50% injectate with the exception of Awi
8-3 which produced less than 20% because it is directed far and deep to the east, crossing a semi-
sealing fault. Tracer tests revealed that of the five injectors on the Awi 9 pad, well Awi 9-1 had
the most rapid returns (5–8 days with transit time of 100–340 m/day), and was consequently shut
in late 1998. The more moderate increases in chloride observed in Awi 8 wells when compared to
Awi 11 wells (Fig. 4), were related to a longer and more convoluted flow path for Awi 9 injection
returns. That is, the abrupt increase in chloride in Awi 11 wells is caused by the more direct return
path of Awi 10 injectate.
The producing wells during this early stage were all deep brine producers. The overall strategy
for field development was peripheral injection to the west (Awi 9) and southwest (Awi 10). By the
338 J.A. Acuña et al. / Geothermics 37 (2008) 332–346

end of 1997 the deep producers at Awi 7 and 8 had not shown any sign of thermal deterioration
due to injection into Awi 9 wells, even though the percentage of injectate they produced was
high based on chloride mass balance calculations. Since a major field expansion to the east was
underway, it was decided not to invest in reconfiguring the western injection system at that time.
In retrospect it seems that injecting in the hottest portion of the reservoir, albeit not planned in
this case, provided some benefits from the point of view of heat recovery as it allowed the colder
injectate to extract additional heat from the reservoir rocks as compared to injecting into a colder
marginal area.
Microseismic monitoring showed that a dense cloud of remarkably deep events was related to
Awi 9 injection (Stimac et al., 2008). These events suggest pressure communication extended
well below the drilled depth. Whether this implies that some limited volume of injectate
circulates and extracts heat along deep flowpaths is still open to debate. In spite of this uncer-
tainty, there was a keen awareness that the area of production could be expanded significantly
if alternative injection locations could be identified and developed. However, as described
below, shuttling injection among the available edge-field locations to minimize its impacts
proved to be the prevailing short-term injection management strategy as field expansion was
implemented.

4. Field expansion and development of steam cap

By late 1997 the installed capacity of Awibengkok was expanded to 330 MWe by drilling 24
production and 12 injection wells in 28 months. The additional steam capacity was developed
from three well pads (Awi 1, 13 and 16) in the eastern part of the field that were sited to exploit
the shallow reservoir top there. Injection was accommodated with two new pads (Awi 14 and 15),
and the use of Awi 2, 3, and 4 on the field edges. More details regarding field expansion and the
new operating configuration can be found in Soeparjadi et al. (1998).
The field was only operated in this new configuration for a short time before a financial crisis
rocked Indonesia, forcing several new geothermal projects to halt development due to reduced
power consumption. As economic conditions slowly returned to normal, renegotiation of the
Salak operating contract with the government of Indonesia allowed for an increase in electricity
generation from the existing plants. As a result of the new contract terms, the field has been
operated at a target of 377 MWe since 2002, taking advantage of the larger capacity of the units
already installed. Obtaining and maintaining steam supply for the higher generation capacity
required make-up drilling campaigns in 2002, 2004 and 2006.
Some innovations were incorporated into the expansion and make-up drilling programs that
ultimately reduced the cost of steam produced. It was recognized from testing of the early wells
completed with 9 58 in. production casing that larger wellbore diameters would result in higher
flowrates (Noor et al., 1992; Murray et al., 1995). Consequently wells drilled from Awi 10-1
onward were completed with 13 38 in. production casings and 13 38 in. or 16 in. tie-backs, and this
resulted in initial production rates of up to 30 MWe for some wells.
Unstable formations containing swelling clays proved to be a problem in the shallow reservoir,
resulting in a high number of stuck pipe incidents in early wells (Stimac et al., 2008). Mapping and
correlation of the clayey units based on drill-cuttings and wellbore image logs allowed drilling
programs to be tailored to case off the most problematic layers shortly after they were drilled.
Drilling with aerated polymer mud rather than brine reduced fluid losses, the need to cement
shallow loss zones, and formation exposure time and reactivity, thus further alleviating the severity
of formation sloughing and collapse.
J.A. Acuña et al. / Geothermics 37 (2008) 332–346 339

Well design technology also evolved with changing well targeting requirements. Long-reach
wells, with bottomhole locations up to 1.6 km from the wellhead were used to delineate reservoir
margins (Awi 9-6 and 14-4). More recently low-angle trajectories were successfully implemented
on several wells in order to maximize the number of fractures encountered within the shallow
steam cap. Awi 16-7 drilled in 2006 reached a deviation angle of ∼60◦ . This well produces more
than 40 kg/s of steam without penetrating beneath the liquid level of the reservoir.
Reservoir permeability has proven to be quite variable, resulting in a relatively high rate of
non-commercial penetrations within the proven reservoir area. Many penetrations have been
plugged and redrilled to more permeable zones. This has led to an increasing effort to improve
the methodology to target wells.

4.1. Well deliverability trends

Due to the shallower reservoir top in the eastern part of the Awibengkok field, it was expected
a steam cap would develop and wells in that area would have high steam rates and lower brine
re-injection requirements similar to what was experienced in the western Tiwi field (Sugiaman
et al., 2004). Reservoir pressure and enthalpy monitoring has confirmed that the steam cap has
steadily expanded with time. Starting in 1994 the steam–water interface dropped from 560 m
above sea level to sea level. The rate of descent increased dramatically after late 1997–early 1998
when Units 4–6 started operation. As the water level has declined, the area of the steam cap has
expanded to include some well-feed zones in the Awi 3, 7, 8, 10 and 11 locations.
The evolution of the average enthalpy for eastern and western wells is shown in Fig. 5. As the
steam cap expanded to include shallow feed zones of wells that were initially producing liquid,
excess steam increased and dry steam production expanded. By 2002 the eastern side of the field
had three dry steam producers and many more wells producing fluids with enthalpies intermediate
between liquid and steam.
Predicting the timing of enthalpy and deliverability changes of wells transitioning from liquid
to steam proved to be a challenging task. Many wells did not evolve entirely into dry steam but
showed an increased enthalpy due to steam production in their shallow feed zones, while their
deep zones remained liquid. This typically resulted in a net productivity increase for these wells,
rendering conventional decline-curve analysis inappropriate.
A technique was developed to combine reservoir pressure and enthalpy of individual feed
zones, with wellbore simulation to construct hydraulic models for each well. The technique
made it possible to calibrate individual wells to their historic deliverability changes and then use
the calibrated models to predict the change in well deliverability for given reservoir pressure and
enthalpy trends (Acuña, 2003). This technique is currently used to match historic well performance
and make short-term steam predictions. Fig. 6 shows the steam production and enthalpy evolution
of Awi 1-3, which is typical of wells that evolve into dry-steam producers.
As a consequence of the development of the steam cap in the eastern reservoir, several new
make up wells were targeted to shallow depths in that part of the Awibengkok field in order to
exploit this new thermodynamic condition. This resulted in prolific dry steam wells drilled at a
relatively low cost. The average enthalpy of the eastern reservoir has continued to increase through
drilling campaigns in 2002, 2004, and 2006–2007 as shown in Fig. 5.
Increased production from the shallow steam zone has led to an increase in non-condensable
gas (NCG) concentration in the steam. The concentration of NCG in the produced steam has risen
with time and has periodically exceeded plant-handling capabilities on the eastern side of the
field. Units 4–6 have therefore been upgraded to handle larger NGC concentrations. The western
340 J.A. Acuña et al. / Geothermics 37 (2008) 332–346

Fig. 5. Evolution of the average enthalpy of fluids produced from the eastern and western sectors of Awibengkok (or
Salak) reservoir. Several wells in the eastern reservoir have become dry steam producers. Some new eastern wells drilled
in 2002, 2004 and 2006–2007 were targeted to exploit the steam cap that had developed.

reservoir does not have this condition because most wells produce mainly from liquid zones.
Predicting future gas trends as a function of production and injection strategy is perhaps one of
the most important areas of geothermal reservoir management and an area ripe for research.
The performance of a few wells has improved through injection of brine or condensate or acid
stimulation. In the last 3 years four sub-commercial wells have been successfully acidized and
now contribute more than 43 kg/s of steam to the system (Pasikki and Gilmore, 2006).

4.2. Injection management strategy

Injection management was established very early as the most important resource manage-
ment issue at Awibengkok, and not surprisingly, devising and reaching consensus on an injection
strategy has proved to be the most difficult aspect of field management (Soeparjadi et al., 1998;
Gunderson et al., 2002). As described in Section 3, the low average steam flash at field start-up
required a very large volume of brine to be injected within 1.5 km of the western production area.
The original strategy was to produce primarily from the central-west reservoir, and inject deep on
its western and southern edges (Noor et al., 1992; Murray et al., 1995).
Brine production remained essentially constant at a little over 1000 kg/s during the years
when fluid production was exclusively from the deep western reservoir (Fig. 7). As described
above, extensive chemical breakthrough was observed on the west side of the field from 1994
to 1998, but was mitigated to some extent by moving injection away from Awi 10 pad. In fact
it is somewhat surprising that edge-field injection at the Awi 9 pad resulted in relatively minor
J.A. Acuña et al. / Geothermics 37 (2008) 332–346 341

Fig. 6. Evolution of well Awi 1-3 from liquid to dry-steam producer. A sharp decline in steam production and enthalpy
in 1998 preceded an abrupt increase in steam production and enthalpy as the well began producing from the steam cap.

Fig. 7. History of brine production in the Awibengkok geothermal field. A declining trend in brine production since field
expansion to 330 MW in late 1997 is related to steam cap expansion, and has made possible a greater degree of flexibility
in redistributing brine injection in the field.
342 J.A. Acuña et al. / Geothermics 37 (2008) 332–346

resource performance issues in comparison to those created by proximal injection in other fields
(Stefansson, 1997). This is primarily due to the relatively high temperature and compartmentalized
permeability pattern in this part of the reservoir.
The amount of brine produced in Awibengkok peaked at more than 3000 kg/s during start-
up of the 330 MWe development in 1998, and has declined to about 2000 kg/s with time, even
though generation has increased to 377 MWe (Fig. 7). The peak brine injection load was accom-
modated by establishing a number of new injection sites on the periphery of the field. Since
that time a steady reduction of injection requirements has occurred due to growth of the shallow
steam cap. This, along with improvement in the capacity of some injection wells due to thermal
effects, has caused that many injection wells are not used anymore. This excess injection capac-
ity has allowed the negative impact on production enthalpy to be minimized by re-location of
injection to the least deleterious injectors, by shuttling injection in response to chemical moni-
toring, tracer test results, and field performance monitoring. This has proven to be an effective
strategy to minimize thermal breakthrough at Awibengkok, with specific examples highlighted
below.
An extensive tracer-testing program was carried out to determine inter-well subsurface con-
nectivity, with the aim of establishing the likelihood of thermal breakthrough for each injection
well and thereby prioritize their use (Gunderson et al., 2002). Early tests were conducted using
fluorescein, but it was found that recovery was strongly impacted by the limited thermal stability
of this chemical, as well as other unknown factors. Later tests were conducted using a variety of
more thermally stable tracers including naphthalene sulfonates that yielded return patterns more
consistent with those indicated by trends in reservoir chloride (Gunderson et al., 2002). All tests
highlighted that injection wells are open within the highly permeable reservoir, with peak tracer
concentrations generally observed between 10 and 30 days.
A dominant pattern of NE-directed fluid flow was observed in reservoir, that complemented
surface structural and borehole image log data indicating that this is the dominant fracture strike
in the field (Stimac et al., 2008). Tracer return patterns to individual wells also supported earlier
interpretations based on pressure interference data that the field is divided into at least three major
compartments that are in moderate communication (Murray et al., 1995). Changes in injection
allocation to optimize production performance became the first key element in the injection
management strategy.
Soon after the field was expanded to 330 MWe rapid chemical breakthrough was observed from
Awi 2-2 to some of Awi 13 and Awi 1 wells; the trend of this rapid breakthrough was NNE-SSW.
Injection was stopped in Awi 2 in 1998. Later in 1998 the percentage of injectate produced by
some wells in the east field declined significantly. At the same time, the brine in the central eastern
part of the field also became more dilute in chloride and this was attributed to condensate injected
into Awi 4 (see Fig. 1) progressing to this area. Microearthquake events correlated with injection
start-up also formed a linear array reaching into the adjacent production area. These observations
prompted a test in late 1998 in which tracer was injected into Awi 4, and all eastern wells were
monitored for tracer recovery. This test verified that nearly all Awi 1, 13 and 16 wells showed
returns from Awi 4, and this well was therefore shut during 1999.
As mentioned earlier expanding the area of production through conversion of injection areas
was also a goal of the long-term resource management strategy at Awibengkok. This required
developing methods of estimating the thermal recovery time and production potential of former
injectors and injection areas. The Awi 10 wells provided one example. The Awi 10 pad had been
used for injection for almost 3 years, and it took about 1 year for Awi 10-1 and 2 years for Awi
10-2RD to be able to produce again once injection stopped. After production started, thermal
J.A. Acuña et al. / Geothermics 37 (2008) 332–346 343

Fig. 8. Observed and computed thermal recovery of well Awi 10-1 in relation to the production and injection history of
the field. The thermal recovery of the well after being used as an injector for almost 3 years was accelerated when the
well started to produce.

recovery accelerated significantly, and by 2001 both wells had returned to their pre-injection
temperature conditions.
An important lesson from trying to model this behavior with reservoir simulation was that
conventional Warren-Root type dual porosity models underestimated the time required for the well
to heat up. In order to properly model the recovery it was necessary to use the MINC approach for
reservoir modeling (Pruess and Narasimhan, 1985). Fig. 8 shows the observed thermal recovery
of Awi 10 as well as the results of the numerical simulator calibrated to replicate this behavior.
It was concluded that if thermal recovery is an issue in a particular geothermal reservoir then the
dual-porosity system must be implemented using the MINC formulation. The experience with
Awi 10 demonstrated that it was feasible to convert injectors to producers and reverse the local
cooling caused by injection.
When cooler fluid such as injectate reaches a particular production well, the enthalpy of
the extracted fluids decreases. The thermal impact of injection is monitored through produc-
tion enthalpy measurements and produced fluid chemistry. To diagnose which production zones
are being impacted by inflows of cooler fluid requires the application of production spinner sur-
veys. When deep liquid feed zones become affected by colder injection, the cooled zones can be
easily recognized in the production spinner surveys as a localized changes in well temperature in
the liquid part of the profile. Many times, however, shallower feed zones become colder, and this
is not evident from spinner analysis because the fluid enters the wellbore in the two-phase flow
region and does not produce any localized pressure or temperature change. To address this issue
a technique to interpret spinner surveys was developed that calculates the mass contribution and
enthalpy of each feed zones in the wellbore (Acuña and Arcedera, 2005).
344 J.A. Acuña et al. / Geothermics 37 (2008) 332–346

The presence of a liquid zone under saturated conditions that separates the steam cap from the
liquid reservoir has been documented at Awibengkok (Acuña, 2005). Feed zones located in this
boiling region decline in temperature and enthalpy when there is reservoir pressure drawdown
as long as they are below the steam–liquid contact at the top of the boiling zone. These zones
eventually evolved into dry steam zones when the steam–liquid contact gets below them as it
drops with time. In the past few years, however, as the steam–liquid contact moved closer to the
level of brine injection, this process was suppressed by the dominance of cooler injectate in the
major fractures. In this case the feed zone does not evolve to dry steam but continues declining
in temperature and enthalpy due to the prevalence of cooler injectate. The steam–liquid contact
drops very slowly or nor at all. The presence of colder fluids at the boiling level can sometimes
be seen in temperature profiles as a departure from boiling conditions at a depth close to the
steam–liquid interface. Flowing spinner logs and the technique described above can also be used
to detect these colder feed zones.
Despite significant effort to minimize the negative impacts of injection, chemical and thermal
breakthrough still threaten numerous production wells and reduce make-up drilling options. Wells
Awi 1-2, 13-5 and 1-7 continue to produce a high percentage of injectate due to injection at the Awi
14 pad based on correlations between changes in produced chloride and injection rates. Injection
at the Awi 15 pad could have been keeping the percent injectate at Awi 13-4 and 13-1 high after
termination of injection into Awi 2-2. This prompted the idea of reducing injection at this pad, a
change that was implemented gradually between 2005 and 2006.
In the western reservoir, wells Awi 10 and 11 seem to be the least affected by injection (i.e.
they produce between 0 and 20% injectate) even though they are flanked by injectors Awi 9 and
15. Here, higher reservoir pressure drawdown has been observed which is consistent with more
limited pressure support from injection than in surrounding areas. This counterintuitive behavior
is attributed to the compartmentalized nature of the reservoir.
Awi 7-1, 8-1, 8-2, 8-4 and 8-5 have produced between 40 and 55% injectate from Awi 9 wells
since early production. Thermal degradation in these deep wells is clearly taking place although
production declines have been minimal due to the high-pressure support they receive. Phasing
out injection at Awi 9 therefore has remained a high long-term priority due to its proximity to
these wells and its own production potential, but awaits results of ongoing efforts to identify an
economically viable alternative (Stimac et al., 2008).

5. Recovery of injection areas for production

The capacity of most Awibengkok injection wells has increased with time as a result of reservoir
pressure drawdown as well as permeability increases attributed to thermal effects due to cooling
of near wellbore area. After having reduced the adverse impacts of injection as much as possible,
and recognizing that significant thermal decline is still occurring through use of the currently
available injection sites, efforts are shifting to evaluate the far western and southeastern margins
of the field for deeper or more distant injection zones (Stimac et al., 2008).
The recovery of injection areas for production offers the promise of improving the performance
of nearby producers and obtaining additional steam at very low cost by using converted injection
wells for production. The feasibility of changing former injectors into producers was proven with
wells Awi 10-1 and 10-2RD in 1998. This success was followed by the conversion of Awi 3
injectors in 2004; these wells were clearly connected to the steam cap, and therefore capable
of production immediately after stopping injection at this location. Awi 3-3 has been one of the
largest steam producers in the field since then. Phasing out injection at Awi 15 by moving virtually
J.A. Acuña et al. / Geothermics 37 (2008) 332–346 345

all remaining eastern injection to Awi 14 in 2006, has led to improved performance in some nearby
wells with deep feed zones. An increase in the temperature of Awi 2 sufficient to make the well
productive may be due the result of stopping nearby injection.
A reduction in the injection flow rate at the Awi 9 pad and shifting of a significant fraction
of the total injection to Awi 9-6, a deep well drilled to the far west, has led to some temperature
recovery in Awi 9-1. During late 2006, Awi 9-1 was tested and shown to sustain production at
commercial rates, providing additional evidence that thermal recovery of the area is possible.

6. Discussion and conclusions

Awibengkok is a geothermal field that was developed in stages with incomplete knowledge
of the resource size and characteristics. Extensive drilling, reservoir monitoring and field testing
have gradually led to a clearer understanding of the reservoir and its response to production, and
of critical resource management issues. Numerical simulation has been used by integrating most-
likely, pessimistic and optimistic models (Acuña et al., 2002) to evaluate alternative development
opportunities and injection and production scenarios. Monitoring programs and reservoir simula-
tion have been used successfully to improve and refine the initial reservoir exploitation strategy,
while making the best of the well configuration available at any given time. Improvements in
prediction of field performance and drilling efficiency have been driven by technical innovation
and process improvement.
Maintaining a balance between pressure support and thermal breakthrough has been, and will
continue to be critical to the successful management of the Awibengkok field. Current efforts are
focused on the long-term plan of relocating injection to deeper or more distal locations along the
field margins, in order to expand the current area of commercial production and stimulate steam
cap expansion. This strategy has been determined to be the most economic means of maintaining
the current generation capacity as it will expand the area available for production, lead to growth
of the steam cap and lower overall steam capacity decline rates. Challenges remain in identifying
alternative injection locations, anticipating changes in the production field that major changes in
the injection configuration will create, and modifying the current field facilities at an acceptable
cost.

Acknowledgements

We thank Chevron Geothermal Indonesia for permitting us to publish this work. We are grateful
to Dave Rohrs of Chevron for providing helpful comments on the preliminary manuscript. Sabodh
Garg and Keshav Goyal are thanked for reviews that substantially improved the final paper.

References

Abrigo, M.F., Molling, P., Acuña, J.A., 2004. Determination of recharge and cooling rates using geochemical constraints
at the Mak-Ban (Bulalo) geothermal reservoir, Philippines. Geothermics 33, 11–36.
Acuña, J., Parini, M., Urmeneta, N., 2002. Using a Large Reservoir Model in the Probabilistic Assessment of Field Man-
agement Strategies. In: Proceedings of the 27th Workshop on Geothermal Reservoir Engineering, Stanford University,
Stanford, CA, USA, pp. 8–13.
Acuña, J., 2003. Integrating wellbore modeling and production history to understand well behavior. In: Proceedings of
the 28th Workshop on Geothermal Reservoir Engineering, Stanford University, Stanford, CA, USA, pp. 16–20.
Acuña, J., 2005. Salak Engineering Studies. Part 2. Awibengkok Cooling and Boiling Zone Behavior. Unpublished Unocal
report, August 2005, 21 pp.
346 J.A. Acuña et al. / Geothermics 37 (2008) 332–346

Acuña, J., Arcedera, B., 2005. Two-phase flow behavior and spinner data analysis in geothermal wells. In: Proceedings
2005 World Geothermal Congress, Antalya, Turkey, p. 6 (paper 1152).
Gunderson, R., Parini, M., Sirad-Azwar, L., 2002. Fluorescein and naphthalene sulfonate liquid tracer results at the
Awibengkok geothermal field, Indonesia. In: Proceedings of the 27th Workshop on Geothermal Reservoir Engineering,
Stanford University, Stanford, CA, USA, pp. 53–58.
Hoang, V., Alamsyah, O., Roberts, J.W., 2005. Darajat geothermal field expansion—A probabilistic forecast. In: Proceed-
ings World Geothermal Congress 2005, Antalya, Turkey, (paper 1153).
Murray, L.E., Rohrs, D.T., Rossknecht, T.G., Aryawijaya, R., Pudyastuti, K., 1995. Resource evaluation and development
strategy, Awibengkok field. In: Proceedings 1995 World Geothermal Congress, Florence, Italy, pp. 1525–1529.
Noor, A.J., Rossknect, T.G., Ginting, A., 1992. An overview of the Awibengkok geothermal field. In: Proceedings of the
21st Annual Convention of the Indonesian Petroleum Association, pp. 597–604.
Pasikki, R.G., Gilmore, T.G., 2006. Coiled tubing acid stimulation: the case of Awi 8-7 production well in Salak geothermal
field, Indonesia. In: Proceedings of the 31st Workshop on Geothermal Reservoir Engineering, Stanford University,
Stanford, CA, USA, pp. 308–314.
Pruess, K., Narasimhan, T.N., 1985. A practical method for modeling fluid and heat flow in fractured porous media. Soc.
Pet. Eng. J. 25, 14–26.
Rose, P., Benoit, D., Goo Lee, S., Tandia, B., Kilbourn, P., 2000. Testing the naphthalene sulfonates as geothermal tracers at
Dixie Valley, Ohaaki, and Awibengkok. In: Proceedings of the 25th Workshop on Geothermal Reservoir Engineering,
Stanford University, Stanford, CA, USA, pp. 36–42.
Soeparjadi, R., Horton, G.D., Wendt, B.E., 1998. A review of the Gunung Salak geothermal expansion project. In:
Proceedings of the 20th New Zealand Geothermal Workshop, University of Auckland, Auckland, New Zealand, pp.
153–158.
Stefansson, V., 1997. Geothermal reinjection experience. Geothermics 26, 99–139.
Stimac, J., Nordquist, G., Suminar, A., Sirad-Azwar, L., 2008. An overview of the Awibengkok geothermal system,
Indonesia. Geothermics, this issue.
Sugiaman, F., Sunio, E., Molling, P., Stimac, J., 2004. Geochemical response to production of the Tiwi geothermal field,
Philippines. Geothermics 33, 57–86.
UGI Union Geothermal of Indonesia, Ltd., 1985. Awibengkok Project Resource Feasibility Report. Unpublished Unocal
report, February 1985, 265 pp.

Vous aimerez peut-être aussi