You are on page 1of 18

Geothermics 82 (2019) 150–167

Contents lists available at ScienceDirect

Geothermics
journal homepage: www.elsevier.com/locate/geothermics

An overview of the Muara Laboh geothermal system, Sumatra T


a,⁎ a a a a
Jim Stimac , Novi Ganefianto , Marino Baroek , Mauliate Sihotang , Irvan Ramadhan ,
Wildan Mussofana, Ridwan Sidika, Alfiadya, Dayinta A. Dyaksaa, Herwin Azisa, Alfianto P. Putraa,
Rudy Martiknoa, Robi Irsamukhtia, Sonny Santanaa, Koji Matsudab, Hideki Hatanakab,
Yoshio Soedab, Laurent Cariouc, Patrick Egermannc
a
PT Supreme Energy, Menara Sentraya, 23rd Floor. Jl. Iskandarsyah Raya No 1A, Jakarta, 12160, Indonesia
b
West Japan Engineering Consultants, Inc., Denki Bldg., Sunselco Annex 8F 1-1-1, Watanabe-Dori, Chuo-ku, Fukuoka, Japan
c
Storengy (Engie group) 12 rue Raoul Nordling, Djinn Building, 92277, Bois-Colombes Cedex, France

ARTICLE INFO ABSTRACT

Keywords: The Muara Laboh geothermal system is a liquid-dominated, fracture-controlled reservoir showing some char-
Stratigraphy acteristics of both intrusion-related and fault circulation systems. The developed reservoir has moderate to high-
Alteration temperature (230–310 °C) with a long NNW outflow (160–230 °C) extending to the Sapan Malulong boiling
Conceptual model chloride springs. Reservoir fluids generally have low salinity (∼400-1600 ppm Cl) with benign chemistry and
Geochemistry
low non-condensable gas (NCG) content (∼0.5 to 2 wt% in steam). The proven reservoir is divided into distinct
Geophysics
SW and NE sectors. The principal deep upflow zone (270 to 310 °C) is located in the SW and is associated with
Porosity
Permeability Patah Sembilan volcano and satellite vents (Anak Patah Sembilan). The SW upflow is characterized by relatively
low salinity (∼400-600 ppm Cl), whereas fluids in the NE emerge from a second upflow along faults and
fractures at 240–250 °C (1200–1600 ppm Cl) and ascend to an initial-state steam cap near the Idung Mancung
fumarole. The shallow NE part of the geothermal system is hosted mainly by Quaternary to Miocene Age an-
desitic to rhyolitic rocks, whereas the deeper SW part of the system occurs in a Mesozoic to Cenozoic Age
plutonic complex and its host rocks that are cut by younger dikes. Rock porosity and permeability trends can be
related to reservoir age, volcanic deposit type, and alteration history. Fracture permeability is high, whereas
matrix porosity (< 1-7%) and permeability are moderate to low and decrease with depth. The volcanic sequence
is capped by tuffs dated at ∼34,000 to 41,000 years BP, incapsulated in debris flows. This sequence may have
formed during and shortly after likely sector collapse episodes of the Patah Sembilan volcano. The shallow clay
cap of the system wraps around the crater, suggesting that it was excavated during crater formation.
Permeable fracture patterns and their origins differ somewhat between the deep SW and the shallow NE
reservoir. In the deep SW reservoir, fluids ascend along fractured margins of stock and dike intrusions. Steeply
dipping intrusions and open fractures strike from WNW to NNW, subparallel to the main Great Sumatra Fault
trend, with a secondary set of N to NE fractures and intrusions parallel with the inferred maximum horizontal
stress direction (SHMax). In the shallower NE steam cap and outflow, N to NE steeply dipping fractures dominate,
with bounding N to NNW faults. Bulk rock alteration includes shallow argillic and transitional clay zones
overlying phyllic and propylitic zones. High-T propylitic alteration (secondary amphibole) and rare potassic
alteration (secondary biotite) are associated with intrusion. Vein paragenesis in the SW reservoir indicates that
portions of the initially permeable propylitic zone have been sealed by calcite and quartz ± prehnite. Late
calcite veins suggest ingress of steam-heated bicarbonate waters contributed to resealing of the system. These
relationships suggest that the SW system underwent extensive boiling, fluid loss, and inflow of dilute bicarbonate
waters, possibly related to sector collapse and partial excavation of the topseal of the system (now Patah
Sembilan crater).


Corresponding author at: 4210 Chaparral Rd., Santa Rosa, CA, 95409, USA.
E-mail addresses: sgcgeo00@gmail.com, wsg-advisor@supreme-energy.com (J. Stimac).

https://doi.org/10.1016/j.geothermics.2019.05.008
Received 30 January 2019; Received in revised form 6 April 2019; Accepted 12 May 2019
0375-6505/ © 2019 Published by Elsevier Ltd.
J. Stimac, et al. Geothermics 82 (2019) 150–167

1. Introduction with a maximum temperature of 104 °C. These springs have not been
linked directly to the developed Muara Laboh geothermal system al-
This paper provides a summary of exploration and development though whether they could be distal outflow from the SW sector of the
activities at the Muara Laboh geothermal area, and an overview of the field is still the subject of ongoing investigation.
current conceptual model. The Muara Laboh geothermal area is located The South Muara Laboh spring group discharges boiling, slightly
in South Solok Selatan Regency, ∼140 km SE of Padang, the capital city alkaline NaCl water (TDS of about 2.6 g/kg) at an elevation of about
of West Sumatra Province (Fig. 1). The geothermal concession extends 800 m, pointing to a deep geothermal origin (Fig. 2). Cation geo-
over about 62,300 ha (62.3 km2) at elevations from 450 to 2000 m and thermometers indicated fluid–rock equilibrium temperatures of
borders the Taman Nasional Kerinci Seblat (Kerinci National Park) to ∼230 °C. It is the South Muara Laboh springs, and in particular, Sapan
the west and south. The prospect was identified based on the presence Malulong (SM) that is considered the main outflow spring of the NE
of fumaroles, mud pools, and hot springs occurring over a wide area reservoir sector. The location of the South Muara Laboh springs appears
(http://www.supreme-energy.com/company/se-muara-laboh/ to be related to the trace of the so-called Suliti segment of the GSF (Sieh
location/). and Natawidjaja, 2000; Figs. 1 and 2). Based on their distribution and
The geothermal area lies within a right stepover of the GSF in an composition, this spring group was inferred by Hochstein and
area of Quaternary volcanism (Mussofan et al., 2018). Detailed de- Sudarman (1993) to be typical of outflow from high-elevation liquid-
scription of the regional geologic setting, stratigraphy, and alteration dominated system with a long flow path. Santoso et al. (1995) used
history of the area is given in a companion paper (Stimac et al., 2019). gravity and resistivity data to infer that the geothermal system was
associated with interpreted NW to NNW-trending faults extending from
1.1. Early exploration the Idung Mancung (IM) fumarole to the Pakan Salasa hot spring
(Figs. 2 and 3).
Abundant thermal manifestations in the lower elevations attracted Pertamina drilled three shallow temperature gradient wells in the
geothermalists to Muara Laboh. The earliest reconnaissance surveys vicinity of the low elevation thermal features from 1992 to early 1993
were undertaken by the Volcanological Survey of Indonesia between (MLB-1, MLB-2, and MLB-3) to depths of 275, 250, and 225 m, re-
1972 and 1979, and geophysical surveys (dc-resistivity, head-on pro- spectively (Wisnandary and Alamsyah, 2012). The maximum measured
filing, gravity, and magnetic surveys) were later conducted by temperature was 68 °C in MLB-2. These low temperatures suggest that
Pertamina and the Institute of Technology Bandung (ITB) group to the wells failed to intersect thermal outflow channels, which may be
further delineate the prospect and trace concealed major faults confined to discrete fractures and limited in extent. The wells have
(Hochstein and Sudarman, 1993; Santoso et al., 1995). A deep reservoir since been plugged and abandoned. MLB-2 is located just north of well
of hot fluids, however, could not be confirmed at that time. C1 (Fig. 3), which also failed to intersect the geothermal system.
Early workers separated low elevation hot springs into two groups Heat flux from the Muara Laboh geothermal prospect was measured
based on location and composition, designating them the North and in May 2011 in an area extending from Patah Sembilan (PS) fumarole in
South Muara Laboh springs (Figs. 1 and 2). Both spring groups are lo- the south to SM hot springs in the north (Sinclair Knight Mertz, 2011).
cated within what appears to be a narrow fault bounded basin. The The total measured heat loss from the Muara Laboh South resource
North Muara Laboh spring group occurs at about 400 m elevation (purple box in Fig. 2) was calculated to be approximately 104 MWt.
mostly along the western margin of the basin, discharging almost High elevation thermal features including the PS and IM fumaroles
neutral-pH, dilute sodium-sulfate-bicarbonate water (TDS∼0.8 g/kg) accounted for about 68 MWt, equating to a mass flow of 26 kg/s steam.

Fig. 1. Left: Map of Sumatra, the Barisan Mountains, and the Great Sumatra Fault (GSF). Tectonic setting of Sumatra is attributed to oblique subduction of the Indo-
Australian Plate beneath Eurasian Plate resulting in the transcurrent GSF, arc volcanism and associated geothermal systems. Quaternary volcanism occurs
130 ± 20 km above the downgoing oceanic plate (Acocella et al., 2018) and mostly along or astride the GSF within the Barisan Mountain Range. Right: Map of
Muara Laboh area and right stepover of the GSF system resulting in a mature pull-apart basin with young volcanism along its SW margin. Thermal discharge is mainly
along the GSF and in Patah Sembilan Crater.

151
J. Stimac, et al. Geothermics 82 (2019) 150–167

Fig. 2. Map of the main thermal manifestations in the Muara Laboh area compiled on an early geologic map (modified from Rosidi et al., 1996). North Muara Laboh
spring group is highlighted in the upper red dashed box, meanwhile the South Muara Laboh spring group, along with higher elevation fumaroles, are highlighted in
the lower purple dashed box. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

Total heat loss from the lower elevation thermal springs was measured quarter of 2019.
at approximately 36 MWt which equated to a mass flow of 108 kg/s of
98 °C water.
2.1. Surface exploration and drilling

2. Project exploration and development chronology Exploration activities from 2008 to 2013 included geoscience stu-
dies to further characterize the extent and temperature of the ex-
PT Supreme Energy conducted pre-feasibility studies in the “Liki ploitable Muara Laboh resource, followed by drilling of six deviated
Pinangawan Muaralaboh” geothermal prospect based on permissions exploration wells (A1, B1, C1, E1, H1, H2). These studies defined an
granted by the Ministry of Energy and Mineral Resources in 2008. extensive shallow clay layer that was spatially associated with NE and
Exploration studies integrating preliminary geologic, geochemical and NW trending faults and the known thermal features. Geochemical
geophysical surveys were completed within a year. These studies con- sampling of thermal features confirmed that the reservoir fluid is likely
firmed the high potential of the area and allowed reserves assessment of benign and high enough temperature for exploitation. Gas geochem-
the concession. Following the tender award of the Muara Laboh con- istry validated a model where upflow is likely associated with the PS
cession to Supreme Energy Consortium in early 2010, a geothermal volcano or its satellite vents, and outflow reaches the surface within the
license was issued to the project company. Power purchase agreements Muara Laboh basin along the GSF. The superheated PS fumarole, lo-
were signed in March of 2012 clearing the way for exploration drilling. cated in the southeasterly crater of the volcano indicated possible up-
As described later, an 80 MWe dual flash power station was eventually flow temperatures of 300–340 °C, whereas IM fumarole and SM
tendered and is expected to reach commercial operation by the last chloride springs indicated temperatures of 270 and 230 °C, respectively.

152
J. Stimac, et al. Geothermics 82 (2019) 150–167

Fig. 3. Exploration-stage conceptual model. (Top) shaded relief


map shows inferred upflow near the PS fumarole with lateral flow
parallel to the NW-trending GSF until it reaches N and NE
trending faults and fractures formed in the stepover. The purple
shaded area delineated the subsurface region > 230 °C reachable
from the geothermal concession (solid orange line). Blue dashed
lines show the likely maximum temperature at reasonable drilling
depth, and projected beyond the limits of drilling to the south.
Yellow shaded areas are regions of thick clay that mark the
eastern edge of fluid circulation and outflow based on MT and B1
and C1 wells. (Bottom) approximately N–S cross section showing
the conceptual model of Muara Laboh. Arrows in left figure in-
dicate the main directions of lateral flow. (For interpretation of
the references to colour in this figure legend, the reader is referred
to the web version of this article.)

The conceptual model after completion and testing of the first six wells found that the TOR, as confirmed by temperature and permeability,
is shown in Fig. 3. conformed to the base of smectite clay inferred from MT data near Pad
The 10 Ωm contours from 1D and 3D inversions of resistivity data A (near IM fumarole), but that it was deeper in the SW where it was
were used to map the base of the conductor (BOC) which served as a well below the first occurrence of epidote (Dyaksa et al., 2016 and
proxy for the base of hydrothermal smectite clay. The BOC and SM Fig. 4).
chloride spring elevation were used to infer a reservoir liquid level at Exploration drilling and well testing confirmed the existence of an
about 800 m asl and boiling point for depth (BPD) curve in the A Pad exploitable geothermal reservoir with good permeability and tem-
area (Fig. 4). This approach was used to predict the top of reservoir peratures from 232 to 312 °C (Situmorang et al., 2016; Fig. 3). Well
(TOR) prior to drilling the first wells and design them accordingly temperatures and testing, and hydrothermal alteration patterns formed
(Dyaksa et al., 2016). The 1D MT inversion tended to exaggerate the basis for revision of the conceptual model (Fig. 3). The highest
changes near discontinuities and to displace them because of dimen- temperature (312 °C) was observed to the south in H2, and the lowest
sional distortion, while 3D inversion shown in the figure smoothed low temperature of about 160 °C was to the NNW in well E1, between IM
resistivity and made the layers extend to greater depth than in the 1D fumarole and SM springs. Based on well temperature trends, the most
inversion. likely upflow area was located within Kerinci National Park near the PS
As mentioned, well locations and trajectories were planned based fumarole (S of the area accessible by drilling). Lateral fluid flow was
primarily on the extent of inferred hydrothermal clay and thermal thought to be parallel to segments of the NNW-trending GSF in the area
features. Wells also targeted specific structures that were thought to of upflow, primarily N to NE within the stepover and pull-apart basin,
play a role in enhancing permeability and supporting fluid flow. and N to NNW again as it reaches the N outflow of the reservoir (see
However, permeable zones could not be related to any single inferred arrows in Fig. 3). The A1 well PT profile and flow characteristics also
structure or linear trend like in some fault-controlled reservoirs. It was provided evidence for an initial-state steam cap in that area. Well C1

153
J. Stimac, et al. Geothermics 82 (2019) 150–167

Fig. 4. N–S 3D MT profile approximately along trend of the


system outflow (NNW) with exploration wells shown. The BOC
based on 5 Ωm contour (red dashed line), TOR interpretation from
MT at about 7 Ωm and, and actual TOR locally revised from well
data are shown relative to the elevation of the SM chloride spring.
Well C1 is outside the reservoir, and both C1 and E1 are somewhat
out of the plane of the section. (For interpretation of the refer-
ences to colour in this figure legend, the reader is referred to the
web version of this article.)

penetrated older basement rocks which were impermeable. Well B1 was also used as the basis for evaluating expansion opportunities (Stage 2).
drilled in a thick clay-altered package of andesitic volcanic and volca- Eleven new wells and one redrill of an existing well (H2) were com-
niclastic rocks that forms the eastern margin of the reservoir (Dyaksa pleted during the development program (Table 1). Four of the original
et al., 2016). A numerical model based on these conceptual under- exploration wells will also be used in Stage 1 (A1, H1 for production;
pinnings indicated that the resource could support at least 60 MWe of E1, B1 for injection), providing a total of sixteen wells for project uti-
power generation with single flash turbine technology (Situmorang lization.
et al., 2016), and this was upgraded through an EPC tender process to
80 MWe using dual flash condensing plant technology. A second 2.3. Production capacity
amendment to the PPA was signed in August 2016, and loan agree-
ments to finance the project were concluded in January 2017. The planned dual flash development will utilize 11 wells for pro-
duction, four for brine injection, and one for condensate injection. Six
producers will tap the shallow NE reservoir, and five will utilize the
2.2. Development drilling
deep SW reservoir. Five development wells (A2, A3, A4, H3, and H4)
were sufficiently tested to confirm their capability to flow commercially
A development strategy for 80 MWe of dual flash capacity was fi-
to the planned steam gathering system (Fig. 6). Together with the ex-
nalized in 2016 that included plans for 11 new wells and two con-
isting A1 and H1 wells (tested during the exploration program), the new
tingency wells to meet production and injection requirements (SEML,
production wells have an estimated production of 260 kg/s and 40 kg/s
2016). Wells were planned from existing pads A, H, and E, and two new
at high pressure (HP; 9.1 bara) and low pressure (LP; 4.5 bara) re-
pads designated D and F (Table 1 and Fig. 5). Production was planned
spectively. This far exceeds required targets of 123 kg/s and 26 kg/s for
from pads A, H and F, whereas injection was designated for pads E, D,
HP and LP steam respectively (Ganefianto, 2018). The above mentioned
and B. Most of the wells were designed as “bigholes” (13-3/8″ pro-
capacity estimate has been discounted by 20% relative to individual
duction casing) to maximize production, minimize drilling problems
well tests to account for production interference. This value was con-
and provide flexibility for contingency sidetracks.
sidered appropriate because some nearby wells have feed zones < 200-
The Stage 1 development drilling campaign and other supporting
400 m from each other. The total mass flow production and enthalpy
geoscientific studies were carried out from the end of October 2017 to
values, collected on a daily basis from an atmospheric flash tank using
May 2018. Well and geoscientific study results were used to develop a
the James Tube method, were periodically verified using the Tracer
3D geologic model in Leapfrog, and a revised conceptual and numerical
Flow Test method.
model of the field (Dyaksa et al., 2018; SEML, 2018). The models were
Injection wells (E2, D1, and D2) encountered permeability in the
reservoir that together with the existing well E1, will accept about
Table 1
1300 kg/s of 140 °C separated brine at 20 barg and 30 barg WHP at E
Exploration and Development Well Locations.
pad and D pad, respectively, well above the estimate of required in-
jection capacity of 420 kg/s. It is worth noting that the E2 injection
capacity was initially low, but significantly improved after being used
for injection of brine produced from A and H wells during the pro-
duction test. Such improvement following a long period of cold brine
injection has been observed in other geothermal wells (e.g., Lovekin
et al., 2017). Plant condensate will be injected into exploration well B1.

3. Reservoir characteristics

3.1. Stratigraphy and geochronology

The stratigraphic section exposed at the surface and penetrated to


about 3100 m MD (about 1300 m bsl) by wells is summarized in Fig. 7.
These sections are generalized from well logs that considered descrip-
Notes: Production sectors (NE, SW) are based on well locations, temperatures, tion of cuttings and core, GR logs, and resistivity images of formations
and fluid compositions, as described in more detail below. Reservoir sectors are (Baroek et al., 2018). The stratigraphic section was dated by 14C ages at
so named because they are on the NE and SW sides of the proven production the surface and U-Pb ages on zircon from subsurface samples and is
zone. described in more detail by Stimac et al. (2019). Ages confirm that

154
J. Stimac, et al. Geothermics 82 (2019) 150–167

Fig. 5. Map of all Muara Laboh wells and inferred major structures (blue lines are revised faults using high resolution digital terrain model obtained from LiDAR data,
new mapping and drilling results). (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

intrusions of both Mesozoic and Tertiary age host the SW reservoir.


Weakly altered microdiorite dikes that cut these older intrusions were
only sampled from cuttings that could not be dated due to mixing with
other lithologies. Volcanic sequences of likely Mesozoic to Quaternary
age and ranging in composition from rhyolite to basalt were en-
countered. The NE reservoir is largely capped by the PS Andesite for-
mation and hosted primarily by a sequence of dacitic to rhyolitic tuffs of
Plio-Pleistocene age. The deeper SW reservoir is hosted mainly by un-
derlying Miocene and probable Late Mesozoic age mafic to silicic se-
quences cut by the aforementioned intrusions.

3.2. Formation porosity and permeability

Core samples and logs provide data on reservoir petrophysical


Fig. 6. Status of high pressure (HP) and low pressure (LP) steam availability
properties. Matrix porosity values measured on core plugs are relatively
based on the testing results. Dashed black lines represent steam requirement for
low, ranging from about 7 to < 1% (Baroek et al., 2018; Table 2 and
HP and LP. Values are discounted 20% from capacities determined by in-
dividual tests.
Fig. 8). As has already been observed in other geothermal fields, there is
a general trend of decreasing porosity with depth, especially for frag-
mental rock types (e.g., Stimac et al., 2004; Siratovich et al., 2016). At
Muara Laboh this results in part because intrusive rock porosities,

155
J. Stimac, et al. Geothermics 82 (2019) 150–167

Fig. 7. Stratigraphy of selected Muara Laboh wells and likely correlative regional formations. Major units are further described in Stimac et al. (2019). Intrusive units
(INT-D for diorite and INT-GD for granodiorite) crosscut other units as shown on the left.

mostly from mid-Tertiary and Mesozoic age encountered in deep H properties. However, some samples were so pervasively fractured that
wells, are only from 1 to 2% (Table 2). However even these rocks have veinlets with some remaining open space could not be avoided. Ex-
structurally enhanced porosity near ubiquitous veins and breccias. amination of core plugs shows that the samples with the higher per-
Partially sealed fracture porosity and permeability are not easily mea- meability include such veinlets, although no veinlet fully transected the
sured using conventional methods, so values from plugs are likely near sample in the direction of permeability measurement (Baroek et al.,
the minimum of a wider range. Fig. 8 shows that when multiple plugs 2018).
are measured from a given depth, a range of values is observed that Petrography of open and partially open veins indicates that porosity
highlights the variability of porosity at the plug scale. near veins has generally been substantially increased by local micro-
Matrix permeability is low to moderate and varies as a function of fracturing, brecciation, and mineral dissolution. Plagioclase pheno-
pressure applied during measurement (Table 2). There is no clear re- crysts commonly show dissolution cavities in which minerals such as
lationship between matrix porosity and permeability although all the epidote, calcite, and adularia have grown (Baroek et al., 2018). Based
higher matrix permeabilities are from samples with > 3% porosity. on these observations, we expect that porosity is on the order of 5–15%
Core plugs were generally selected away from open fractures and par- near open fractures, providing some local fluid storage, and con-
tially open veins and represent hydrothermally altered primary rock nectivity of the main fracture pathways to the rock matrix. Since

Table 2
Petrophysical measurements on core plugs.

Notes: Bulk density calculated from grain density and water-saturated pore volume; sample permeability listed at 0.0005 is below
the threshold of measurement. Data from CoreLab (A2, A3, H4) and Geoservices ITB (A1, B1, and E1).

156
J. Stimac, et al. Geothermics 82 (2019) 150–167

Fig. 8. Core matrix porosity versus vertical depth and by rock type. There is a general trend of decreasing porosity with depth (r2 = 0.6) but a range of values is
observed in a given core interval when multiple plugs are measured.

reservoir rocks are mainly of Mesozoic and Tertiary age in the SW, there be more recent and developed mainly in the stepover between the main
has been ample time for infilling of primary porosity during hydro- GSF segments but have also been found in some wells near the Siulak
thermal alteration events. The shallower NE reservoir is hosted by Plio- fault segment. Interpretation of Bouguer gravity suggests N–S trending
Pleistocene volcanic rocks and retains a higher percentage of its original horsts and grabens have formed within the broader stepover area
primary porosity. (Mussofan et al., 2018). This N–S fracture set is well oriented for ex-
tensional failure in the current stress field and is thought to be im-
portant in controlling permeability and fluid flow in the Muara Laboh
3.3. Reservoir structure and permeability
geothermal system, especially in the NE reservoir (A Pad) and far N
outflow. The NE-SW fault trend is interpreted to be antithetic to the
Structural controls on the Muara Laboh system have been described
GSF. It is considered to be the youngest fracture set based on the field
by Mussofan et al. (2018) and Baroek et al. (2018). The Muara Laboh
mapping data, where examples of NE-SW faults crosscutting the older
geothermal system is a hybrid between magmatic-hydrothermal intru-
NW-SE and N–S structures were observed. Fig. 9 shows effective frac-
sion-related and fault-controlled circulation systems. Shallow perme-
ture orientations interpreted from coupled borehole image and PTS log
ability is related to fault and fracture systems formed in a stepover of
interpretation. Effective fractures are defined as those that support fluid
the GSF, where a large pull-apart basin has formed. Deeper perme-
flow based on drilling losses, well completion, and flow testing data.
ability is related to magmatic-hydrothermal fracturing associated with
Fractures striking NW-SE, N–S and NE-SW are most common, consistent
intrusions emplaced over an extended period of time at a deeper level
with the pull-apart basin setting. A subpopulation of WNW-ESE striking
into the evolving fault and fracture system. Accurate mapping of faults
fractures is present in some deep wells in the SW (F2, H2, H3). The
is however impeded by young volcanic cover and dense tropical vege-
majority of fractures are high angle (> 60°), consistent with a strike-
tation.
slip faulting environment; however, a significant subset of the popula-
Present day in situ stress orientations were determine from borehole
tion dip from 40-60°, indicating that normal faults are also common
breakout and drilling induced tensile fractures. In near-vertical wells
(see Baroek et al., 2018, Fig. 8).
(maximum 30° inclination), the minimum principal horizontal stress
The reservoir clay cap, defined by the MT conductive layer, me-
(Shmin) is parallel to the orientation of borehole breakouts, whereas the
thylene blue analysis (MeB), petrographic analysis of well cuttings, and
maximum principal horizontal stress (SHMax) is parallel to tensile
limited XRD data is found mainly within the Quaternary Andesite
fractures (Baroek et al., 2018). The SHMax and Shmin orientations de-
volcanics of the Patah Sembilan Complex. Smectite clay is most strongly
rived from well D1 are consistent with those determined from modeling
developed within tuff and volcaniclastic layers that originally contained
for exploration wells H1 and H2, where SHMax is NE-SW and Shmin is
abundant glassy material. The shallow steam cap intersected in Pad A
NW-SE (GMI, 2013, 2012). This local in situ stress orientation is con-
wells is associated with the stratigraphic contact between Patah
sistent with some regional in situ stress data in Sumatra based on Tingay
Sembilan Andesite and underlying Undifferentiated Silicic Formation,
et al. (2012) who showed SHMax is commonly NE-SW. However, the
whereas liquid feed zones at Pad A and Pad E are present within the
GSF is a segmented strike-slip fault system that can generate stress
Undifferentiated Silicic Formation and are thought to be related mainly
perturbation due to the stepover mechanism between segments gen-
to steeply dipping faults and fractures. Core samples from A Pad show
erating pull-apart basins and pop-up structures. This can change the
hydrothermal veins and breccias are common, highlighting the inter-
direction of the SHMax locally as shown by the focal mechanism data
play of faulting and episodic fluid overpressure.
from N–S to NNW.
Deep permeability encountered by Pad H wells corresponds mainly
Based on surface mapping and borehole image logs, the dominant
with highly fractured contact zones within a multiple intrusive complex
fault and fracture trends of the area are NE-SW, N–S and NW-SE
composed of dikes, sills and stocks emplaced from Mesozoic to recent
(Baroek et al., 2018). The NW trend (NNW to WNW) is represented
time. Open fractures and dike margins commonly show the same strike
mainly by the GSF and subsidiary structures and is considered a long-
orientations, suggesting that dikes and fractures share a common origin.
lived fault/fracture system. The N–S and NE-SW trends are thought to

157
J. Stimac, et al. Geothermics 82 (2019) 150–167

Fig. 9. Stereonets of effective fracture strikes from image logs (from Baroek et al., 2018). The dominant trends are NW-SE, N–S and NE-SW, similar to identified
faults. N is number of fractures.

Petrography suggests that some dikes are weakly altered and post-date the low temperature of well B1 and is consistent with a discontinuity in
the bulk of hydrothermal alteration (Stimac et al., 2019). The heat the MT (1D TE Mode) (Fig. 11, right), and a thickening and plunging of
source of the Muara Laboh geothermal system may be related to these the conductive layer (3D) to the east. This has been interpreted to be
younger structurally controlled intrusions. Their exact locations are associated with a prominent N to NNE trending fault (F2). The west
unknown, but the distribution of young vents indicates they are likely boundary could possibly go as far as F10 as marked by the dis-
beneath the Anak PS vents and/or the PS volcano sector collapse. appearance of the clay cap.

3.4. Reservoir geophysics 3.5. Hydrothermal alteration and paragenesis

The SW extension of the geophysical conductor has been used to Bulk rock alteration at Muara Laboh includes extensive shallow ar-
estimate the maximum extent of the reservoir and therefore its upside gillic and transitional clay zones that overlie phyllic (quartz-sericite-
potential as shown in Fig. 10. A NW-SE elongated conductive layer in pyrite ± adularia ± titanite) and propylitic (epidote-chlorite-quartz-
the eastern part of the system is associated with volcanic sediment pyrite) zones that host the permeable reservoir. High-T propylitic al-
(based on drilling results of B1), interpreted not to be associated with teration (amphibole-epidote-adularia ± prehnite-garnet) and rare po-
the current geothermal reservoir. The absence of a conductive layer tassic alteration (biotite-adularia-amphibole) are associated with intru-
inside the PS crater may be in part related to its excavation by sector sions. Some of this high temperature alteration, especially potassic zones,
collapse or long-term erosion. 14C ages on tuffs associated with surficial now occurs in rocks at < 250 °C and is therefore relict. Phyllic alteration
debris flow deposits indicate sector collapse may have occurred as re- is more prevalent in silicic rocks, whereas propylitic alteration is domi-
cently as 34 to 41 ka (see Stimac et al., 2019), but is most likely nant in intermediate volcanics and intrusives (Stimac et al., 2019).
somewhat older than this. The geothermal system at Lihir Island was Methylene blue analysis conducted on cuttings at the wellsite has
affected by partial removal of its clay cap by sector collapse, however provided a simple and effective means of determining the extent of low-
the system resealed itself and remained extant (White et al., 2010; Rae temperature smectite-rich alteration (Fig. 12).
et al., 2010). This suggests that a geothermal reservoir extending below Vein paragenesis has been described in detail by Baroek et al.
a young sector collapse crater may survive these events, although it (2018) and Stimac et al. (2019). In general, adularia and epidote were
would likely be impacted by cooling. As described above regarding the deposited on fracture initiation and are found in open-space veins in the
exploration-stage conceptual model, the distribution of low resistivity system upflows, whereas quartz ± prehnite and wairakite dominate in
clay also suggests that geothermal circulation may be ongoing in a the shallow steam cap region and are found as the last minerals to have
narrow fault-controlled NNW-SSE corridor extending from well H1 to formed in open fractures. An episode of cooler steam-heated inflow is
the E of the PS fumarole (cf. Figs. 3 and 10). represented by anhydrite ± calcite in the NE. Fluid inclusion data in-
The maximum northern extension of the reservoir is limited by dicate some cooling of the NE reservoir from a maximum of 250–260 °C
disappearance of the conductive layer (Fig. 11, left) and interpreted to to the current temperature of < 240 °C (Stimac et al., 2019). Wairakite
be associated with a NE-trending fault (F4). Meanwhile the southern is the dominant vein mineral currently forming in the shallowest part of
boundary is still uncertain due to the nature of the clay cap within the the reservoir where temperatures may currently not support formation
crater. A limit on the eastward extension of the reservoir is defined by of epidote.

158
J. Stimac, et al. Geothermics 82 (2019) 150–167

Fig. 10. 3D MT total conductance to 1000 m


below surface. Section lines A-A’ and B-B’ are
shown in Fig. 11. Inverted triangles are MT
stations. Dashed black line is likely reservoir
extension based on conductor extent. Dotted
black line is a possible reservoir extension be-
yond the conductor provided clay cap has been
partially removed. Black triangles are inferred
volcanic vents based on LiDAR image inter-
pretation.

Vein paragenesis in the SW reservoir indicates that portions of the sector collapse and partial excavation of the topseal of the system (now
initially permeable propylitic zone have been sealed by infilling of Patah Sembilan crater). This process may have also led to formation of
fractures by calcite and quartz ± prehnite. Late calcite veins suggest the NE steam cap, but the two areas are now poorly connected. In the
downflow of steam-heated bicarbonate waters following infilling of the deeper SW reservoir open space veins lined by epidote and adularia
higher temperature veins, and resealing of the system. These relation- remain with only minor infilling of later minerals. Temperatures mea-
ships suggest that the SW system underwent extensive boiling, fluid sured in the wells are near the maximum represented by fluid inclusions
loss, and ingress of overlying bicarbonate waters, possibly related to in the permeable reservoir below about −200 m elevation. Alteration

Fig. 11. MT cross section lines from Fig. 10 (1D – top and 3D – bottom). Left) line A-A’ showing north and possibly south reservoir boundaries. Right) Line B-B’
showing east and possibly west reservoir boundaries.

159
J. Stimac, et al. Geothermics 82 (2019) 150–167

Fig. 12. Cross section A-B-C along the main axis of the Muara Laboh geothermal system with contoured resistivity (Ωm) methylene blue (MEB) index values from
well cuttings and 3D MT data. Wells B1 and C1 are further from section line and show deeper low temperature alteration (see Fig. 10). Wells E1 and E2 also diverge
from the section line at deeper levels.

history indicates that the distribution of matrix porosity and perme- Meanwhile A3 and A4 geothermometry (Tmeas. < TQtz < TNKC) suggests
ability is tied to the history of fracture formation, infilling, and fluid that these wells are nearer the reservoir margin. H1 fluid has an NE
sources and temperatures. upflow composition with some mixing with cooler fluid
(TQtz < Tmeas. < TNKC).
3.6. Fluid geochemistry Upflow temperature for the SW reservoir compartment is in excess
of 300 °C, possibly associated with intrusion and young volcanic vents
The distribution of thermal features and related heat flow were of Anak PS. Wells completed to the SW (H2RD, H3, and H4) have re-
described in Section 1.1. Fumaroles are present at higher elevations servoir chloride of 400–500 ppm (maximum in H2OH is 900 ppm)
(∼1400 to 2000 m asl) and chloride springs at lower elevations (800 m confirming a different fluid source than A Pad area (Fig. 15). All wells
asl). drilled to the SW have lower geothermometry (except TQtz) than mea-
Representative well chemistry is given in Table 3 and portrayed in sured temperatures and inferred major feed zones. The most recent H3
Figs. 13 and 14. The data show that fluid produced from the NE area (A measured temperature, however, is almost equal to geothermometry
and H1 area) is distinct from fluid produced in the SW area (all other H (TQtz), suggesting a nearby upflow at ∼300 °C. The low chloride con-
wells), suggesting that each compartment has its own deep fluid source. centration and high temperature suggest heating of deeply circulated
Measured temperatures are also consistent with poor connection be- meteoric fluid with little or no magmatic input.
tween these reservoir sectors. Upflow locations were constrained During the numerical model matching process, locations of the SW
through iterative initial-state temperature matching during numerical upflow(s) had to be moved closer to bottom hole locations of H2RD, H3
model construction (Situmorang, pers. comm., 2018). The final upflow and H4 to better match measured well temperatures than originally
locations implemented in the model are those that provide best initial- envisioned in the conceptual model (Situmorang, pers. comm., 2018).
state temperature and pressure match to combined well data. This is geologically reasonable because all H wells drilled SW en-
Upflow temperature of the NE reservoir sector is estimated at countered entries related to intrusive damage zones, indicating young
≤250 °C based on fluid geothermometry. This upflow is thought to be magmatism (Stimac et al., 2019); and the bottom of the wells were
structurally focused but ultimately related to an intrusion heat source below the relatively young Anak PS volcanic vents.
with some magmatic contribution (high-chloride, highest 3He/4He The current conceptual model of the system is shown in Fig. 15.
isotope composition). All A pad wells and H1 have reservoir Drilling results indicate that reservoir fluid beneath the NE area mainly
Cl > 1200 ppm (Fig. 13) with relatively similar geothermometry (TNKC, outflows to SM springs along deep-rooted structures (narrow corridor
TQtz) ranging from 238 to 250 °C. Geothermometry of well A1 shows the between faults F13 and F2). Limited fluid may also migrate south to H1
most equilibrated fluid (TNKC=TQtz = 250 °C), suggesting it is closest to based on the chemistry and temperature of this well. Some of the fluid
an upflow. This well is deviated to the SE and has a minor reversal near from the SW sector is assumed to outflow to Pad A through a weak
bottom. Thus the NE upflow was placed to the west of the well at a connection, likely located near H3. That outflowing fluid from the SW
possible fault intersection, rather than near its bottom (Fig. 15). mixes and re-equilibrates with upflow fluid from Pad A.

160
J. Stimac, et al. Geothermics 82 (2019) 150–167

Table 3
Representative well fluid chemical data.
Well Date FCV (%) Concentration (brine) NCG (wt% in stm) Sample P (barg) Enthalpy (Kj/kg) Notes

pH TDS SiO2

ML-A1 08 Apr 2018 100 6.87 3820 547 0.45 11.00 1314
ML-A1 14 Apr 2018 100 6.89 3790 533 0.45 10.80 1310 During FPT
ML-A1 15 Apr 2018 30 6.65 3730 530 0.52 14.10 1329
ML-A1 15 Apr 2018 10 6.24 3530 507 0.67 21.00 1714
ML-A2 08 Jul 2017 100 2.10 11.30 2800
ML-A2 07 Jan 2018 100 0.49 8.54 2800
ML-A2 16 Feb 2018 100 0.47 8.46 2800
ML-A3 11 Aug 2018 100 7.27 3680 531 0.90 9.30 1069 During FPT
ML-A3 14 Aug 2018 20 7.19 3790 539 0.94 7.20 1059
ML-A3 14 Aug 2018 3 7.28 4050 593 1.10 2.90 1193
ML-A3 03 Oct 2018 100 7.44 3810 565 0.42 8.80 1084
ML-A3 18 Oct 2018 100 6.36 3840 565 0.48 9.50 1201
ML-A3 09 Dec 2017 100 6.85 3960 553 0.36 9.90 1371 During FPT
ML-A3 10 Dec 2017 25 6.45 3680 514 0.44 18.30 1524
ML-A3 10 Dec 2017 4 5.79 2940 409 0.45 25.10 2568
ML-A3 16 Feb 2018 100 5.95 4040 558 0.20 9.00 1137
ML-A4 12 Sep 2017 100 6.73 3780 542 0.59 10.20 1208
ML-A4 19 Sep 2017 100 7.01 3780 551 0.50 10.20 1213
ML-A4 12 Nov 2017 100 6.50 3760 499 0.48 9.93 2184
ML-A4 25 Jan 2018 100 6.88 3930 570 0.28 10.60 1328 During FPT
ML-A4 27 Jan 2018 20 6.37 3640 478 0.35 18.30 1467
ML-A4 27 Jan 2018 5 6.10 3100 410 0.39 24.80 2500
ML-H2RD 23 Nov 2017 100 8.35 2020 748 0.43 3.51 1612
ML-H2RD 08 Dec 2017 100 8.26 3030 951 0.21 2.24 1920
ML-H3 28 Jan 2017 100 7.36 1900 843 0.29 11.80 1273
ML-H3 10 Jan 2018 100 7.09 1860 816 0.39 11.60 1334
ML-H3 01 Feb 2018 100 7.31 1920 851 0.77 11.70 1287 During FPT
ML-H3 03 Feb 2018 25 6.77 1810 788 0.87 18.40 1268
ML-H3 03 Feb 2018 13.3 6.34 1700 769 0.95 27.50 1250
ML-H4 18 Feb 2018 100 6.63 1930 758 1.00 9.50 1169
ML-H4 07 Apr 2018 100 6.83 2010 974 0.74 12.70 1312
ML-H4 22 Apr 2018 100 6.80 1980 956 0.75 12.60 1303 During FPT
ML-H4 22 Apr 2018 30 6.36 1950 946 0.86 15.80 1283
ML-H4 23 Apr 2018 10 6.09 1770 867 1.25 31.30 1299

Notes: FCV (flow control valve opening), TDS and SiO2 in mg/kg. TDS, total dissolved solids; FCV, flow control valve; NCG, non-condensible gas; FPT, flow
performance test.

Current understanding and interpretation based on the available is indicated by temperature profiles in wells E1, E2, and C1, but these
data suggest that most outflowed fluid from the SW area exits the descending fluids must be separated from the drilled reservoir by sealed
system to the SW instead of S of Pad H, based on the following ob- faults. This infiltration is likely more widespread than shown, but are
servations: 1) distribution of conductive layer (and likely removal of the supported by some well data for these particular locations.
conductor by crater formation); 2) Surface alteration found inside the
crater indicating hydrothermal activity after the collapse; and 3) the
3.6.1. Stable isotopes
characteristics of PS fumarole, located in the south of the field, are
Stable isotope data indicate that meteoric water is the dominant
inconsistent with it being outflow from Pad H fluids (PS fumarole is
component of reservoir fluid, with all well samples plotting near the
higher in N2). But these features do not preclude the existence of ex-
meteoric water line (Fig. 16). The fluid (total discharge) of SW wells
ploitable reservoir in the crater area (see Stimac et al., 2019).
indicates almost no positive oxygen shift, while the fluid of NE wells
The significant topography around Muara Laboh potentially influ-
exhibits a positive shift about 1.5 permil. This small shift is likely re-
ences the patterns of recharge and discharge. Possible sources of in-
lated to a higher degree of water-rock interaction for the NE reservoir
filtrating deep meteoric recharge shown in Fig. 15 include SW of the PS
fluid, also consistent with the higher Cl concentration of this sector. The
Crater, along F10, and E of F2 near well B1. Additional cool infiltration
difference in lithology of reservoir rock, crystalline granodiorite in the

Fig. 13. Cl-SO4−HCO3 ternary plot (left) and B versus Cl plot (right) indicate distinct fluid chemistry between the NE and SW reservoir sectors.

161
J. Stimac, et al. Geothermics 82 (2019) 150–167

Fig. 14. Cl-TNKC plot, showing different chloride concentrations between NE and SW sectors.

SW and intermediate to silicic volcanics (probably richer in glassy The range of helium isotope values (R/Ra) in Muara Laboh is
material) in the NE, would be one important factor in the difference in 1.12–3.14 (Mussofan, 2018), rather low compared to other subduction-
fluid chemistry between these sectors. related geothermal systems (typically between 4–7) or subduction-re-
lated volcanic gas (typically ≥7). These low He isotopic values may
3.6.2. Gas chemistry indicate contributions of near-surface crustal fluids (e.g. He derived
The helium isotopic composition (R/Ra) in geothermal fluids re- from uranium and thorium decay in crustal rocks). Mixing between
flects a variation in the heat source, with Average Crust (R/Ra ∼0.02) mantle and crustal fluid sources seems reasonable since thick silicic
as one end-member and the mantle (MORB, R/Ra ∼7-8) as the other. volcanic and granitic intrusive rocks of Mesozoic and Tertiary ages are
MORB-like helium isotopic compositions, such as found at The Geysers, widely distributed in Muara Laboh. The highest He isotope value is
Lassen, Coso, Cerro Prieto, Steamboat, and Long Valley, indicates a from IM fumarole (3.14) compare to PS (1.95) fumarole. This indicates
recently or presently active magmatic heat source, whereas lower more that IM has a higher “mantle” component to its gas. SM has a low value
crustal like helium isotopic compositions, such as found at Dixie Valley, of 1.12 which is consistent with meteoric contributions in outflow
indicates a source of heat that is predominantly crustal (Kennedy, features (Fig. 17). One possible explanation for the higher He isotope
2002). This suggests more deeply circulated meteoric waters along fault value of IM fumarole is that is it located in an area of recent extension
systems in regions of elevated heat flow. that lacks or has a thinner underlying Mesozoic to Tertiary basement

Fig. 15. Map view of conceptual model showing reservoir sectors


(SW in green, NE in orange and pink labelled Pad A and H1 (NE)),
proven reservoir area (purple dashed line), likely upflow regions
(red dashed lines), outflows (orange and pink regions with blue
arrows showing flow direction), and recharge locations (light blue
dashed circles) and directions relative to major structures
(number F1 to F20 in dark blue) and volcanic vents (black trian-
gles). SM Cl spring lies N of D1OH off map. MR signifies areas of
known (from well profiles) and inferred downward infiltration of
meteoric water recharge. Question marks indicate a higher degree
of uncertainty with the model element. (For interpretation of the
references to colour in this figure legend, the reader is referred to
the web version of this article.)

162
J. Stimac, et al. Geothermics 82 (2019) 150–167

Fig. 18. Liquid tracer return patterns for A3 and A4 versus time. Tracer returns
were detected after 64 days for A4 and 66 days for A3.

Fig. 16. Stable isotope plot for selected Muara Laboh well samples with World
Meteoric Water Line shown for reference.

Fig. 19. Vapor tracer PFC analyses versus time. No returns of PFC observed in
A3, A4, and A2.

samples were analyzed by the Thermochem laboratory in Bandung.


Well A2 did not have liquid tracer data because it produces 100%
steam. For the steam tracer samples, a total of 28, 31, and 37 samples
were collected from A2, A3, and A4, and four samples from IM, but
there were no returns of the vapor tracer in either the wells or the
fumarole (Fig. 18). Four NDSA samples were collected from SM hot
springs but no tracer returns were detected.
Fig. 17. Helium isotope composition of Muara Laboh compared to other geo- These results show that there is connectivity between production
thermal systems. wells (Pad A) and injection well E1 within the liquid zone. The samples
from A3 yielded one positive tracer return 66 days after tracer injection.
The samples from A4 had the first positive returns 64 days after in-
section relative to the heavily intruded corridor along the southern
jection (Fig. 19). The last samples from A3 and A4 have the highest
strand of the Great Sumatra Fault. Deep fault penetration of this ex-
NDSA concentrations, and the curve does not tail-off, suggesting that
tended crustal section would allow dike intrusion and gases emanating
tracer was being produced when the flow test was terminated. Data
from deep mantle contributions to dominate over crustally-derived
available from the test is unfortunately not sufficient to provide a
gases. Alternatively, there is more recent intrusion with a higher mantle
quantitative analysis. There is an interesting observation that NDSA in
gas component associated with the NE upflow that is reaching the PS
A4 arrived two days earlier than A3 which is closer to injection well E1.
fumarole.
These results confirm that there is connectivity between production
wells (A pad) and injection well E1. However, this is a relatively long
3.6.3. Tracer test interval for first arrival, indicating that the injected fluid will have time
Liquid- and steam-phase tracer tests were conducted to understand to thermally equilibrate as it migrates toward the production field.
potential connectivity between injection well E1 and the production Experience in other fields has indicated thermal breakthrough (cooling
wells A2, A3, and A4 (Alfiady, 2018). The tracers and amounts used related to injection returns) is very unlikely in Muara Laboh in the short
were 100 kg of 1, 6 Naphthalene Di Sulfonate Acid (NDSA) for liquid term given this return interval (e.g., Vicedo et al., 2008). Additional
and 25 kg of Per-Fluor-Carbon (PFC) for steam. These tracers were in- longer-term tracer tests are planned once the field reaches full pro-
jected into well E1 on October 5, 2017, during the period of parallel duction to better assess connectivity between injection wells (E1, E2,
production testing of A3 and A4 wells. Tracers were injected using the D1, D2) and Pad A producers. Excess brine injection capacity allows
PT Thermochem Rapid Injection Pulse Skid to deliver a highly con- some flexibility in the use of each injector.
centrated pulse of tracer solution to the reservoir. The vapor tracer,
PFC, was injected as an emulsified solution using a peristaltic pump. 3.7. Reservoir pressure, temperature and permeability
After tracers were injected, samples were taken from A2, A3 and A4
wells during production tests of each well. For the first 10 days after the Review of well temperature, pressure, and fluid chemistry data in-
tracers were injected, fluid samples were taken from each well every dicate the field lies on a single pressure gradient. Injection and pro-
day. The sampling frequency was reduced to every two days after that. duction PTS and PT surveys provide constraints on well feedzone lo-
Monitoring was also conducted in the thermal manifestations, SM cations (Fig. 20). They illustrate that the Pad A area has the shallowest
chloride spring, and IM fumarole. Liquid tracer samples were analyzed reservoir top and a thin initial state steam cap from about 800 m to
by Thermochem laboratory in Santa Rosa USA, whereas the steam 620 m asl. This fluid level is slightly lower than the SM chloride springs

163
J. Stimac, et al. Geothermics 82 (2019) 150–167

Fig. 20. Feed-zone locations and phase conditions based on well testing. Colored boxes show wells in the steam cap area (red), adjacent NE sector (blue), and deep
SW sector (green). (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

of the system interpreted as the main outflow of the system (Fig. 21, Pads define the SW sector (Fig. 15 and Table 1) which has more dilute
also see Fig. 4). Wells F1ST and H1 appear to be located at the margins fluid, a deeper reservoir top, and higher temperature (260 to > 310 °C).
of the A Pad shallow cap and no shallow steam was encountered. The NE reservoir sector is defined by F1ST, H1, and A Pad wells,
Further to the SW, wells F2 and other H wells encountered limited whereas the system outflow is defined by wells drilled from Pad E and
permeability until about sea level. Permeability is found in most wells D. The reservoir boundary was confirmed to extend further southwest
from sea level to 800 m bsl 1300 m bsl in the deepest well. and south and to greater depth than previously thought, but to the NE it
was proven to be more restricted than anticipated from exploration.
4. Reservoir conceptual model Review of well temperature, pressure, and fluid chemistry data indicate
that although the field lies on a single pressure gradient, it has two
The system conceptual model was revised after development dril- distinct reservoir compartments in the SW and NE in terms of depth,
ling and well testing (Figs. 15 and 22). Most wells drilled from H and F temperature, and salinity. In particular, differences in fluid chemistry

Fig. 21. Map and cross section (note bend) of reservoir top (TOR) and temperature distribution based on drilled wells. The SM spring lies out of section to the NNE of
the line. The reservoir boundary to the south is uncertain.

164
J. Stimac, et al. Geothermics 82 (2019) 150–167

Fig. 22. NNE-trending slice of 3D model showing temperature isosurfaces, pads and wells, main thermal features, and topography. The system or an adjacent system
probably extends to the PS fumarole but is not shown.

indicate that there is little connectivity between these sectors in its 0.5 Ma, mostly in the form of ignimbrites that ponded in the developing
current state. Further information on this issue will be available as more basin (Mussofan et al., 2018; Stimac et al., 2019). This indicates that
wells are simultaneously flowed during plant start-up and operation. pull-apart basins do not necessarily inhibit volcanism, as suggested by
The SW sector is thought to be large in extent and represents the Muraoka et al. (2010). Instead the volcanism that occurs there is
main area for future development, although its extent to the S and SW is commonly silicic, and forms caldera or graben-type depressions rather
constrained mainly by geologic and geophysical considerations with than tall edifices. This seems to be true at Ranau (Bellier and Sébrier,
considerable uncertainty. For these reasons, conservative assumptions 1994) and, possibly Suoh as well. In these cases, dilational jogs at GSF
regarding reservoir extent and fracture permeability were made in the stepovers have closely associated magmatic systems, both active and
undrilled portion of the SW sector. The NE sector hosts a highly dormant. Once magmatism becomes established in an area of extension,
permeable shallow steam cap of limited extent attached to a long liquid intrusions may reach a high crustal level and evolve to more silicic
outflow to the NNW. This sector has moderate temperature composition with time. This results in heating and softening of the
(232–245 °C) and salinity and low gas concentration consistent with a upper and middle crust, with a concomitant loss of strength. Ductile
recent loss of steam. The water level observed in the Pad A area is deformation and ballooning of narrow fault zones at depth accom-
currently at about 620 m asl and the reservoir top at about 800 m asl. modates dilation and assures brittle extension above. Meteoric waters
The distal SM chloride spring, considered the main discharge of the descending from above and along the brittle margins of ductile zones
system, is at about 795 m asl, but other springs in the S Muara Laboh are heated and ascend nearby as geothermal upflow plumes.
group range up to 887 m asl. The SW reservoir may outflow to the west Muraoka et al. (2010) highlighted the importance of pull-apart
or to the far N Muara Laboh springs, consistent with the deeper re- basins in hosting Sumatran geothermal systems. The Muara Laboh
servoir top and lower spring elevation (Fig. 2), but this cannot be geothermal system is localized in one of 13 such pull-apart basins
confirmed at this time. formed in a right stepover of the GSF (Mussofan et al., 2018). Structures
subparallel to the GSF and oriented roughly perpendicular to it are
5. Discussion locally permeable and provide pathways for fluid circulation in the
basin. Paleozoic basement rocks are present on the NE side of this basin
It is likely that many geothermal systems in Sumatra will share some but have not been documented beneath it. It is therefore possible that
common features with the Muara Laboh system. Therefore its devel- the western edge of the Paleozoic crustal block is located on the eastern
opment well data and conceptual model provide useful information for margin of the basin and has remained a zone of weakness and intrusion
future explorationists and developers. Sumatra appears to have a higher since at least Mesozoic time. This zone may have eventually accom-
fraction of silicic volcanic centers of Quaternary and Pleistocene age modated transform faulting and intrusion related to oblique subduction
than most other islands in Indonesia that are dominated by andesite. that has been ongoing since at least the Miocene (Stimac et al., 2019).
Silicic intrusive complexes tend to have large volume and locally The Muara Laboh geothermal system has characteristics of both
shallow magma emplacement that may drive geothermal circulation. intrusive and fault-controlled geothermal systems. Fracturing and high
The localization of Quaternary vents of andesitic to rhyolitic composi- temperature upflow in the deep SW reservoir seems to be associated
tion along the Siulak segment of the GSF at Muara Laboh potentially with relatively recent intrusion below Anak PS vents. Relatively dilute
provides a wider distribution of shallow magma emplacement and heat fluid, low gas, and lower temperature (< 240 °C) suggest the im-
energy than is typical in areas dominated by andesitic stratovolcanoes portance of fault circulation in the NE sector and its main outflow.
without complex structures or significant satellite vents (Fig. 1). Linear These features underscore the need for cautious resource capacity es-
arrays of volcanic vents and underlying intrusions also encourage de- timation, as permeability may be more localized, limiting reservoir
velopment of multiple hydrothermal upflows, as appear to occur at volume compared to more distributed systems. On the other hand, the
Muara Laboh. Drilling results indicate at least two areas of upflow for combination of relatively high temperatures and high fracture perme-
the Muara Laboh system, accounting for difference in fluid chemistry ability can deliver unusually high power density like at Silangkitang
between the NE and SW reservoir sectors. It is possible additional up- (Drestanta et al., 2018).
flow is also occurring to the S or SE of the drilled area. The open fracture trends observed at Muara Laboh are similar to
At Muara Laboh, well stratigraphic evidence confirms the im- those inferred at other geothermal areas in central and south Sumatra.
portance of rhyolitic volcanism from the Plio-Pleistocene until about For example, Nurseto et al. (2015) identified several major lineaments

165
J. Stimac, et al. Geothermics 82 (2019) 150–167

in the Ulubelu geothermal area. According to their interpretation, these 6. Conclusions


structures trend NW-SE and WNW-ESE, while associated subsidiary
faults trend NNW-SSE. The NW-SE faults are synthetic to the Semangko The Muara Laboh geothermal system has been recently developed
fault (dextral sense of slip) and formed prior to the Pliocene or early for 80 MWe of power generation, scheduled for commissioning in late
Pleistocene. Similar to Muara Laboh wells near the Siulak fault, bore- 2019. Six exploration wells (one later redrilled) and eleven develop-
hole image logs from wells UBL-3 and UBL-26 indicate open, productive ment wells were completed (Table 1), and along with data from surface
fractures that strike NW-SE and dip mainly to the SW. A few open exploration and geophysical surveys provide the basis for a revised
fractures also show NE-SW and N–S strikes (Nurseto et al., 2015). conceptual understanding of the exploitable reservoir. The system hosts
Based on image log and cuttings interpretation at Muara Laboh, we two sources of hot upflow with distinct fluid compositions and tem-
conclude that dikes and related fractures play an important role in peratures. The primary upflow, identified in deep SW wells
permeability in the deep SW reservoir. It is well documented that diking (270–310 °C), has a more dilute chemistry. A second upflow with higher
generates a variety of fractures at the dike tip and along its path that salinity and lower temperature (≤250 °C) feeds a shallow steam zone in
could locally enhance permeability (Dering et al., 2019). Seismicity the NE near the IM fumarole. The bulk of outflow is channeled along N-
associated with diking events also shows more complex patterns (lo- S to NNW trending structures and reaches the surface at the SM spring.
cations, focal mechanisms) than would be expected for simple extension Additional untapped reservoir may extend from the proven H and F Pad
at a dike tip, including shear failure on planes parallel to the inflating area further to the SW and S toward the PS fumarole which is located in
dike (e.g., White et al., 2011). Using detailed mapping of well exposed a young sector collapse of PS volcano. Alternatively, there could be a
dike swarms, Dering et al. (2019) have shown that the number of faults limited zone of reservoir along the F2 fault extending from near the PS
and joints increases towards the dike swarm indicating they formed in fumarole toward H1 and A Pad based on lower resistivity in this area
the same event. They argue that faults and shear fractures form a da- (Fig. 10).
mage zone around the dikes even though the dikes occupy Mode I ex- Reservoir rocks range in age from Plio-Pleistocene to Late
tension fractures. This is consistent with the observation that many Cretaceous. Alternating sequences of andesitic lavas and tuff, and silicic
geothermal systems appear to be related to dikes and are elongate in the ash-flow tuffs and related deposits filled an expanding tectonic basin that
most likely direction of dike propagation (e.g., Dubti, Ethiopia – see formed in a right stepover of the GSF. Fractures producing geothermal
Stimac et al., 2014) or are in part above or astride silicic domes that are fluids are mostly subparallel to the GSF in the SW (NW to NNW), and
likely fed by dikes such as East Salak (Stimac et al., 2008), Wiyang mostly N to NE in the basin interior, similar to other pull-apart basins in
Windu (Bogie et al., 2008), and Bulalo (Vicedo et al., 2008). Sumatra. Fracture permeability is highest in the shallow steam cap
According to Dering et al. (2019), characteristics of damage zones where formation contacts may also play a role in fluid circulation.
related to dike emplacement (“dike corridors”) differ from damage Matrix porosity and permeability are generally moderate to low, varying
zones related to major faults. The distribution of macrofracture in- as a function of rock type, depth, and sample alteration history.
tensity across dike corridors is not organized around a high strain zone Hydrothermal alteration assemblages are typical of neutral-Cl geo-
such as a fault core, and there is no large associated fault. Instead, a thermal systems, with a phyllic assemblage (illite/sericite-quartz-
relatively broad corridor of increased fracturing is related to the em- pyrite) more prevalent in silicic rocks, and a propylitic assemblage
placement of a dike swarm itself (Delaney et al., 1986; Dering et al., (epidote-chlorite-quartz-pyrite) in intermediate volcanics and in-
2019). trusives. Sealing of the upper part of the propylitic zone has reduced
Dilute fluid chemistry, low NCG, and 3He/4He ratios suggest that permeability in the area of F and H pads, resulting in a reservoir top
the heat sources fueling the Muara Laboh geothermal system are mostly that is deeper than would be anticipated from exploration surveys.
crystallized, and not contributing a significant mass component to fluid Wairakite is the dominant vein mineral currently forming in the shal-
circulation. Instead shallow meteoric and steam-heated bicarbonate lowest part of the reservoir where temperatures appear to be too low to
waters, infiltrating along deep faults and fractures and in young basin support formation of epidote.
sedimentary and volcanic sequences appear to mainly be extracting and The Muara Laboh geothermal system appears to have had a long
recirculating heat. The NE upflow and shallow culmination of the re- history, with the SW and NE sectors being better connected in the past.
servoir at about 700 m depth and 237 °C, and its cooler outflow appear Sealing of the upper propylitic zone in the SW and formation of a
to be controlled by N to NE-trending faults in the central part of the shallow steam cap in the NE might both reflect response of the system
basin, and possibly along a horst block (Mussofan et al., 2018). This to relatively recent sector collapse of the Patah Sembilan volcano,
part of the system bears some features typical of fault-controlled sys- which excavated the clay cap in that area.
tems, but fracture patterns and gas chemistry indicate deeper magmatic
controls as well. Acknowledgments
The unexpected deepening of the reservoir top in SW Muara Laboh
is related to infilling of fractures with late-stage minerals as the system Many thanks to Supreme Energy Muara Laboh and its partners for
evolved and cooled in that area (Baroek et al., 2018; Stimac et al., permission to publish this work. Discussions with the Resource Team at
2019). Based on mineral paragenesis, boiling and later ingress of cooler Supreme, including Irene Wallis and Anna Colvin were most helpful.
waters into the system at shallow levels led to deepening of the re- Constructive reviews by David Sussman and Erik Layman are also
servoir top. Sector collapse of the PS volcano, probably shortly before greatly appreciated.
34 to 41 ka tuff eruptions may have played an important role in pres-
sure drop and boil down of the system. A similar sector collapse is References
implicated in the extreme boildown of the Karaha Telaga Bodas geo-
thermal system (Moore et al., 2002). In that system, chalcedony was Acocella, V., Bellier, O., Sandri, L., Sébrier, M., Pramumijoyo, S., 2018. Weak tectono-
formed at high temperature and is interpreted as having resulted from magmatic relationships along an obliquely convergent plate boundary: Sumatra,
Indonesia. Front. Earth Sci. 6, 3.
rapid depressurization and boiling. High-temperature chalcedony is Alfiady, 2018. Muara Laboh Result of Tracer Test (RTT), Unpublished Supreme Energy
locally present in the sealed propylitic interval at Muara Laboh, but it Report Dated April 13, 2018.
does not appear to be widespread. It is therefore possible that sealing at Baroek, M., Stimac, J., Sihotang, A.M., Putra, A.P., Martikno, R., 2018. Formation and
fracture characterization of the Muara Laboh geothermal system, Sumatera,
Muara Laboh took place more gradually and over a longer time interval, Indonesia. Geothermal Resource Council Trans. 42.
or the system was less well sealed from sources of meteoric recharge.

166
J. Stimac, et al. Geothermics 82 (2019) 150–167

Bellier, O., Sébrier, M., 1994. Relationship between tectonism and volcanism along the Geothermal Congress, 2010.
Great Sumatran Fault Zone deduced by SPOT image analysis. Tectonophysics 233, Rosidi, H.M.D., Tjokrosapoetro, S., Pendowo, B., Gafoer, S., Suharsono, 1996. The
215–231. Geology of the Painan and Northeastern Part of the Muarasiberut, Quadrangles
Bogie, I., Kusumah, Y.I., Wisnandary, M.C., 2008. Overview of the Wayang Windu geo- (0814-0714), Systematic Geological Map of Indonesia Scale 1:250000. Geological
thermal field, West Java, Indonesia. Geothermics 37, 347–365. Research and Development Centre, Bandung.
Dering, G.M., Micklethwaite, S., Cruden, A.R., Barnes, S.J., Fiorentini, M.L., 2019. Santoso, D., Suparka, M.E., Sudarman, S., Suari, S., 1995. The geothermal fields in central
Evidence for dyke-parallel shear during syn-intrusion fracturing. Earth Planet. Sci. part of the Sumatra fault zone as derived from geophysical data. Proceedings,
Letters 507, 119–130. Proceedings World Geothermal Congress. pp. 1363–1366.
Drestanta, Y.S., Soeda, Y., Drakos, P., Astra, D., Lobato, E.M.L., 2018. Building a 3D earth SEML, 2018. The Muara Laboh Conceptual and Numerical Models Update & Stage-2
model of Silangkitang Geothermal Field, North Sumatra, Indonesia. Proceedings 6th Assessment. Unpublished Report Dated July 2018.
Indonesia International Geothermal Convention Exhibition (IIGCE). SEML, 2016. Muara Laboh Development Drilling Strategy, Unpublished Report Dated
Dyaksa, D.A., Ramadhan, I., Ganefianto, N., 2016. Magnetotelluric reliability for ex- Feb. 2016.
ploration drilling stage: study cases in Muara Laboh and Rantau Dedap geothermal Sieh, K., Natawidjaja, D., 2000. Neotectonics of the Sumatran fault, Indonesia. J.
project, Sumatera, Indonesia. Proceedings, 41st Workshop on Geothermal Reservoir Geophys. Res. 105, 28295–28326.
Engineering, Stanford University, SGP-TR-209. Sinclair Knight Mertz, 2011. Muara Labuh Geothermal Prospect Heat Loss Survey,
Dyaksa, D.A., Stimac, J., Baroek, M., Sihotang, A.M., Ramadhan, I., Martikno, R., O’Brien, Unpublished Report for Supreme Energy, Dated October 2011.
J., 2018. An application of 3D temperature modeling during and after development Siratovich, P.A., Heap, M.J., Villeneuve, M.C., Cole, J.W., Kennedy, B.M., Davidson, J.,
drilling at Muara Laboh, Sumatera, Indonesia. Proceedings New Zealand Geothermal Reuschlé, T., 2016. Mechanical behavior of the Rotokawa Andesites (New Zealand):
Conference. insight into permeability evolution and stress-induced behavior in an actively utilized
Ganefianto, N., 2018. Muara Laboh Stage-1 Development Drilling and Well Testing geothermal reservoir. Geothermics 64, 163–179.
Report. Unpublished Supreme Report Dated Nov. 2018. Situmorang, J., Martikno, R., Perdana Putra, A., Ganefianto, N., 2016. A Reservoir si-
GMI, 2013. Geomechanical Model and Fracture Analysis for ML-H2 Well, Muara Laboh mulation of the Muara Laboh Field, Indonesia. Proceedings 41st Workshop on
Field, Unpublished Geomechanics International Report. Geothermal Reservoir Engineering, SGP-TR-209.
GMI, 2012. Geomechanics International: Geomechanical Model and Fracture Analysis for Stimac, J., Armadillo, E., Kebede, S., Zemedkun, M., Kebede, Y., Teclu, A., Rizzello, D.,
ML-H1 Well, Muara Laboh Field, Unpublished Geomechanics International Report. Mandeno, E., 2014. Integration and modeling of geoscientific data from the Tendaho
Hochstein, M.P., Sudarman, S., 1993. Geothermal resources of Sumatra. Geothermics 22, geothermal Area, Afar rift, Ethiopia. 5th ARGeo Conference.
181–200. Stimac, J.A., Nordquist, G., Aquardi, S., Sirad-Azwar, L., 2008. An overview of the awi-
Lovekin, J.W., Morrison, M., Champneys, G., Morrow, J.W., 2017. Temperature recovery bengkok geothermal system, Indonesia. Geothermics 37, 300–331.
after long‐term injection: case history from Soda Lake, Nevada. Trans. Geothermal Stimac, J.A., Powell, T.H., Golla, G., 2004. Porosity and permeability of the Tiwi
Resource Council 41, 2770–2779. Geothermal Field, Philippines based on continuous and spot core measurements.
Moore, J.N., Allis, R., Renner, J.L., Mildenhall, D., McCulloch, J., 2002. Petrologic evi- Geothermics 33, 87–107.
dence for boiling to dryness in the Karaha-Telaga Bodas geothemal system, Indonesia. Stimac, J., Sihotang, A.M., Mussofan, W., Baroek, M., Jones, C., Moore, J.N., Schmitt,
Proceedings 27thWorkshop on Geothermal Energy, Stanford University. A.K., 2019. Geologic Controls on the Muara Laboh Geothermal System, Sumatra,
Muraoka, H., Takahashi, M., Sundhoro, H., Dwipa, S., Soeda, Y., Momita, M., Shimada, Indonesia. Paper to Geothermics in Revision. this issue. .
K., 2010. Geothermal systems constrained by the Sumatran fault and its pull-apart Tingay, M., Morley, C.K., King, R., Coblentz, D., 2012. Present-day stress Field of
basins in Sumatra, Western Indonesia. Proceedings World Geothermal Congress. Southeast Asia. Adapted from Oral Presentation at AAPG International Conference
Mussofan, W., 2018. Gas Chemistry of the Muaralaboh Thermal Manifestations. and Exhibition.
Unpublished Supreme Energy Report, April 5, 2018. Vicedo, R.O., Stimac, J.A., Capuno, V.T., Lowenstern, J.B., 2008. Establishing Major
Mussofan, W., Baroek, M.C., Stimac, J., Sidik, R.P., Ramadhan, I., Santana, S., 2018. Permeability Controls in the Mak-Ban Geothermal Field, Philippines. GRC
Geothermal resource exploration along Great Sumatera Fault segments in Muara Transactions 32, 309–313.
Laboh: perspectives from geology and structural play. Proceedings, 43rd Workshop on Wisnandary, M.C., Alamsyah, O., 2012. Zero generation of muara laboh numerical model:
Geothermal Reservoir Engineering, Stanford University. SGP-TR-213. role of heat loss and shallow wells data on preliminary natural state modeling. GRC
Nurseto, S.T., Sardiyanto, S., Koestono, H., Thamrin, M.H., Kamah, M.Y., 2015. Transactions 36, 825–830.
Evaluation of the productive geologic structure in Ulubelu geothermal system. White, R.S., Drew, J., Martens, H.R., Key, J., Soosalu, H., Jakobsdóttir, S.S., 2011.
Proceedings World Geothermal Congress, Melbourne, 2015. Dynamics of dyke intrusion in the mid-crust of Iceland. Earth Planet. Sci. Lett. 304,
Rae, A.J., Ramirez, E., Villafuerte, G., Kilgour, G., Milicich, S.D., Fraser, H., Bignall, G., 300–312.
2010. Recent exploration drilling at Lihir geothermal Field, PNG: effects of cata- White, P., Ussher, G., Hermoso, D., 2010. Evolution of the Ladolam geothermal system on
strophic sector collapse on a magmatic-hydrothermal system. ProceedingsWorld Lihir Island, Papua New Guinea. Proceedings World Geothermal Congress 2010.

167

You might also like