Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Geothermics 55 (2015) 150–158

Contents lists available at ScienceDirect

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

A reformulation of USGS volumetric “heat in place” resource


estimation method
Sabodh K. Garg a,∗ , Jim Combs b
a
Leidos Inc., 10260 Campus Point Drive, San Diego, CA 92121, USA
b
Geo Hills Associates LLC, 12945 Vanderbilt Dr. #207, Naples, FL 34110, USA

a r t i c l e i n f o a b s t r a c t

Article history: A reformulation of the USGS volumetric “heat in place” method is presented. More specifically, expres-
Received 4 December 2014 sions for “recoverable heat” are derived by considering specific power cycles, i.e., single-flash and binary.
Accepted 10 February 2015 The latter approach eliminates the ambiguities associated with specifications of a reference temperature
Available online 5 March 2015
and the utilization efficiency. Since the use of an arbitrarily low reference temperature such as the ambi-
ent temperature yields too optimistic an estimate for the recoverable heat, the abandonment temperature
Keywords:
instead of the ambient (or condenser) temperature should be assumed as the reference temperature in
Heat in place
order to obtain realistic estimates of recoverable heat. The standard USGS method will, however, yield a
Thermal recovery factor
Utilization efficiency
non-zero available work for temperatures between the ambient (or condenser) and abandonment tem-
Reference temperature peratures; in this case, the only way to conform the USGS method to reality is to require conversion
Monte Carlo (utilization) efficiency to also tend to zero. The probable range for the thermal recovery factor is con-
Single flash sidered, and it is argued that in the early exploration stage prior to deep drilling and well testing, the
Binary cycle proper lower limit for the recovery factor is zero. An illustrative example of the standard USGS method
Sumikawa vs. the present reformulation of the USGS method for resource estimation is presented using previously
published data from the Sumikawa geothermal field. A Monte Carlo procedure was employed to predict
the megawatt capacity at Sumikawa for three cases: (1) the reformulated USGS method, (2) the effect
of a zero minimum recovery factor, and (3) the standard USGS method. A total of 100,000 Monte Carlo
simulations were used in each case to compute the cumulative probability distribution. Comparing Cases
1 and 3, it is apparent that the original USGS method predicts a much larger capacity than the new for-
mulation. The main effect of a zero minimum recovery factor (Case 2) is to yield a considerably smaller
capacity at the 90% confidence level compared to Case 1. In a certain sense, a zero minimum recovery
factor encodes a lack of knowledge regarding the productivity of as yet undrilled wells.
© 2015 Elsevier Ltd. All rights reserved.

1. Introduction of combining estimates with uncertainties for the temperature,


area, and thickness of a geothermal reservoir into a prediction of the
During early stage exploration of geothermal resources associ- stored energy (“heat in place”) with uncertainty. Probability density
ated with an identified hydrothermal convection system, it is often functions for temperature, area, thickness, and thermal recovery
desired to obtain an estimate of the potential electrical energy that factor are assumed based on uncertain estimates in order to cal-
might be produced from the geothermal system in order to justify culate the probability distribution function for the “recoverable
the anticipated cost for development of the geothermal resource. heat”. The probability distribution function for the “recoverable
In the 1970s, researchers at the United States Geological Survey heat” can be obtained using a Monte Carlo approximation. Thus,
(USGS) developed a methodology to quantify the uncertainty of the USGS volumetric estimation method together with Monte Carlo
estimates of the geothermal resources associated with an identi- simulations is often used to provide estimates of the probable gen-
fied hydrothermal convection system (e.g., Nathenson, 1975a,b; eration capacity of a geothermal resource. Taken at face value,
Nathenson and Muffler, 1975; Muffler and Cataldi, 1977; Brook the method is deceptively simple. In two previous papers, the
et al., 1979). The USGS volumetric estimation methodology consists authors focused on the difficulties of specifying a probability den-
sity function for thermal recovery factor (Garg and Combs, 2010)
and utilization efficiency (Garg and Combs, 2011) required to com-
∗ Corresponding author. Tel.: +1 858 826 1615; fax: +1 858 826 6044. pute the probable generation capacity. More recently, Williams
E-mail address: gargs@leidos.com (S.K. Garg). (2014) has re-examined the probable range for the thermal

http://dx.doi.org/10.1016/j.geothermics.2015.02.004
0375-6505/© 2015 Elsevier Ltd. All rights reserved.
S.K. Garg, J. Combs / Geothermics 55 (2015) 150–158 151

recovery factor, and opined that the thermal recovery factor must
be greater than zero. In conformity with Garg and Combs (2010),
the authors maintain that a zero value for the thermal recovery
factor cannot be a priori excluded.
In this paper, a reformulation of the USGS volumetric “heat
in place” method is presented. More specifically, expressions for
“recoverable heat” are derived by considering specific power cycles, Brine

Temperature
e.g., single-flash and binary. The latter approach eliminates the Pinch Point
ambiguities associated with specifications of a reference temper-
ature and the utilization efficiency. The probable range for the
Secondary Fluid
thermal recovery factor is considered, and it is argued that in the
early exploration stage prior to deep drilling and well testing, the
proper lower limit for the recovery factor is zero. Finally, the new
formulation along with Monte Carlo method is applied to assess
the power potential of an actual geothermal field, and results are
Pre-heater Evaporator
compared to those derived using the conventional USGS method.

0% 100%
2. Recoverable heat Heat Transfer

Consider a geothermal reservoir with volume V. Heat stored in Fig. 1. Heat transfer in a simple binary cycle.
the geothermal reservoir, qR , is given by:

qR = V c (TR − Tr ) (1) A simple example will suffice to illustrate the importance of


specifying a reasonable value for the reference temperature Tr .
where cl , cg , cr = heat capacity of liquid water, steam, and rock Consider a typical high temperature geothermal resource with an
grains; A = reservoir area; H = reservoir thickness; Sl , Sg = liquid average temperature of 250 ◦ C. It is assumed that a single flash
water and steam saturation in pores (Sl + Sg = 1); Tr = reference tem- cycle with a separator pressure of 5 bars (saturation tempera-
perature; TR = average reservoir temperature; V = reservoir volume ture = 151.831 ◦ C) is used for generating power. In this case, the
(=AH);  = porosity; c = volumetric heatcapacity of fluid-saturated “abandonment temperature” equals 151.831 ◦ C. With a reference
rock = Sl l cl + Sg g cg + (1 − ) r cr ; l , g , r = density of liq- temperature of 15 ◦ C, a value commonly used in previous applica-
uid water, steam, and rock grains. tions of the USGS method, the heat recovered in Joules (J) at the
The choice of the reference temperature is very important since wellhead is:
it has a large effect on the computed value for the “recoverable
heat”. In previous applications of the USGS method, two choices qwUSGS = 235˛ J
for Tr , i.e., the ambient (say ∼15 ◦ C) or the condenser (∼40 ◦ C) tem-
peratures, have been used (see e.g., Brook et al., 1979; Williams, Since the wells will not be discharged below the “abandonment
2014). While any value for reference temperature, including abso- temperature”, a more reasonable value for the reference temper-
lute zero, may be employed to compute the “stored heat” in the ature is 151.831 ◦ C, and the heat recovered at the wellhead in the
reservoir, the choice of a reference temperature is restricted for actual case is:
calculating the “recoverable heat”. For the latter purpose, it is use-
ful to introduce the concept of an “abandonment temperature”, qwActual = 98.169˛ J
which is defined as the temperature below which a geothermal
reservoir will not be produced. Values for “abandonment temper- The two values for the heat recovered at the wellhead differ
ature” will depend upon the power cycle. Thus, for a flash-type by a factor of more than two. Use of an arbitrarily low reference
power plant, the lower limit for the “abandonment temperature” temperature such as the ambient temperature yields too opti-
is given by the saturation temperature Tsep corresponding to the mistic an estimate for the recovered heat. It is suggested that the
separator pressure, and for a binary power plant it equals the so- “abandonment temperature” instead of the ambient (or condenser)
called pinch point temperature Tp (see Fig. 1). The pinch point is temperature should be assumed as the reference temperature in
the location in the heat exchanger with the least temperature dif- order to obtain realistic estimates of recovered heat.
ference T between the primary (water) and secondary fluids, and
corresponds to the bubble point Tb for the secondary fluid. Thus, 3. Available work and conversion efficiency
the pinch point temperature Tp is given by:

Tp = Tb + T (2) In a geothermal reservoir most of the heat is contained in the


reservoir rocks and not in the reservoir fluid. The fluid is needed
A geothermal thermal recovery factor Rg is defined as the ratio to mine the heat from the rocks. Injection of cooled brine after its
of the heat recovered at the wellhead, qw , to the heat stored in the use in the power plant and/or influx of cold water from reservoir
reservoir, qR . boundaries is required to mine heat from the reservoir. The tem-
perature of the discharged fluid gradually declines from its initial
Rg = qw /qR (3)
value TR to the “abandonment temperature”. Average temperature
Combining Eqs. (1) and (3), the heat recovered at the wellhead of the produced fluid T̄ is in general different from TR .
qw can be expressed by the following expression: Assuming isenthalpic flow in the wellbore and neglecting the
work required to raise water to the wellhead, the enthalpy of pro-
qw = ˛(TR − Tr ) (4) duced fluid at the wellhead, hw , is equal to that of liquid water at T̄ .
where Thus
 
˛ = Rg V c (5) hw = hw T̄ (6)
152 S.K. Garg, J. Combs / Geothermics 55 (2015) 150–158

The amount of fluid produced at the wellhead, mw , is given by: The separated brine is injected into the reservoir, and the steam
is used to generate power. The mass of the fluid produced at the
mw = qw / (hw − hr ) (7)
wellhead is given by Eq. (8) with Tr = Tsep . The steam fraction of the
where hr is the enthalpy of liquid water at the reference tempera- produced fluid is:
ture, Tr . Substituting from Eq. (4) into Eq. (7), there follows:       
mstm = mw hw T − hw Tsep /hgl Tsep (12)
mw = ˛ (TR − Tr ) / (hw − hr ) (8)
where hgl (Tsep ) denotes the heat of vaporization. Combining Eqs.
The concept of availability plays an important role in the (8) and (12), there follows:
USGS volumetric “heat in place” method. Availability is defined    
mstm = ˛ TR − Tsep /hgl Tsep (13)
as the maximum work (or power) output that can theoretically
be obtained from a substance (water) at specified thermodynamic Substituting mstm for m in Eq. (10), the available work for the
conditions (wellhead) relative to its surroundings. DiPippo (2008) single-flash case is given by:
observes:  
“To achieve this ideal outcome, there are two thermodynamic ˛ TR − Tsep   
WAflash =   hstm Tsep − hw (Tc )
conditions that must be met: hgl Tsep
   
(1) All processes taking place within the system must be perfectly −TcK sstm Tsep − sw (Tc ) (14)
reversible.
in Eq. (14), Tc denotes the condenser temperature (◦ C) and TcK is
(2) The state of all fluids being discharged from the system must
the absolute condenser temperature.
be in thermodynamic equilibrium with the surroundings.”

3.2. Simple binary cycle


The first of these conditions amounts to neglecting losses due
to friction, turbulence, and other sources of irreversibility. The sec-
The produced brine at an average wellhead temperature T̄ is
ond condition requires that any fluids discharged from the system
used to heat the working fluid (e.g., isobutane), and the spent brine
are in temperature equilibrium with the surroundings. None of
is injected back into the reservoir. Given the turbine inlet pressure
the real power cycles can meet these conditions, and the “elec-
pin , the corresponding saturation temperature of the secondary
trical energy” that is generated is always less than the “available
fluid Tb may be determined from thermodynamic data for the fluid
work”. The conversion efficiency (utilization factor) is the ratio of
(see e.g., NIST, 2010). The secondary fluid is assumed to be satu-
the “actual electrical energy” to the “available work”.
rated at the turbine inlet. With a temperature differential T at
Neglecting kinetic or potential energy effects, the maximum
the pinch point, the temperature of the brine at the pinch point Tp
energy output per unit mass of the substance e is given by (DiPippo,
is given by Eq. (2). The heat provided by hot brine to vaporize the
2008):
secondary fluid (from incipient bubbling at Tb ) is given by:
e = h − hex − TexK (s − sex ) (9)       
Q = mw hw T − hw Tp = ˛ TR − Tp (15)
where h and s denote the enthalpy and entropy of the substance
The mass of secondary fluid msf vaporized by heat Q is:
(e.g., steam or hydrocarbon vapor) at turbine inlet conditions with
 
temperature T, TexK is the absolute exit temperature (Kelvin), and msf = ˛ TR − Tp /hsfgl (Tb ) (16)
sex is the entropy of liquid phase (water or hydrocarbon liquid) at
the exit temperature. For mass m of the substance, the available where hsfgl (Tb ) denotes the heat of vaporization for the secondary
work is therefore given by: fluid at temperature Tb .
Substituting from Eq. (16) into Eq. (10), the available work for
WA = me = m [h − hex − TexK (s − sex )] (10) the binary cycle becomes:
 
The available work, as used in the conventional USGS method ˛ TR − Tp   
(Brook et al., 1979), is computed by replacing mass m in Eq. (10) WAbinary = hsfg (Tb ) − hsfl (Tc ) − TcK ssfg (Tb ) − ssfl (Tc )
hsfgl (Tb )
by mw from Eq. (8), h by hw (TR ), and putting Tex = To . Here To is the

ambient temperature and is often assumed to be 15 ◦ C. Thus −Vsf (Tc , psfb ) [pin − psfb (Tc )] (17)
˛ (TR − To )
WAUSGS = [hR − ho − ToK (sR − so )] (11)
(hR − ho ) The last term in Eq. (17) denotes the work required to raise
In general, mass that enters the turbine is not the same as mw the pressure of the secondary fluid (liquid) from the condenser
computed from Eq. (8). In single (or dual) flash systems, liquid brine pressure (=saturation pressure at condenser temperature) to the
(and the energy contained in it) is rejected at the separator tem- turbine inlet pressure, Vsf is the specific volume, and subscripts g
perature; only the separated steam is used to generate power. The and l refer to gas and liquid.
binary technology involves the use of a secondary fluid, heated by
the brine, to generate power; the brine is in any event rejected at a 3.3. Illustrative examples
higher temperature than the ambient temperature. For a real power
cycle, m, h and s in Eq. (10) should be evaluated at the turbine inlet We consider two typical geothermal resources with temper-
conditions rather than the wellhead, and the most appropriate exit atures of 150 ◦ C and 250 ◦ C. The lower temperature resource is
temperature (Eq. (10)) is the condenser temperature. In the follow- suitable for binary applications while the higher temperature
ing, expressions are developed for available work for single-flash resource will most likely be exploited using flash technology. Before
and binary power cycles. embarking on an analysis of power cycles, it is useful to evaluate
available work using the USGS methodology (Eq. (11)). Available
3.1. Single-flash work computed from Eq. (11) for the two resource temperatures
(150 ◦ C and 250 ◦ C) and a reference temperature of 15 ◦ C is pre-
It is assumed that the produced fluid with an average tempera- sented in Table 1. All the water properties were evaluated along
ture T̄ at the wellhead is separated at a separator temperature Tsep . the saturation line.
S.K. Garg, J. Combs / Geothermics 55 (2015) 150–158 153

Table 1 of over 40% (referred to ambient temperature) have been reported


Comparison of available work computed with USGS method (Eq. (11)) with that
for a couple of power plants utilizing advanced power cycles, most
obtained using power cycle Eqs. (14) and (17). For the sake of simplicity WA /˛ instead
of WA is given. See text for additional details. operating ORC power plants have relatively low (less than 25%) sec-
ond law efficiencies. GeothermEx (2004) used a utilization factor
Resource temperature TR (◦ C) Available work WA /˛ (◦ C)
of 0.45 (referred to 15 ◦ C) in a resource assessment of several low
USGS method Power cycle temperatures geothermal fields in Nevada and California. Although
150 24.53 10.47 the latter utilization factor appears to be theoretically possible,
250 65.00 29.10 resource characteristics (e.g., change in resource temperature over
time) and economic considerations will usually dictate a much
lower utilization factor.
For the geothermal resource with an average temperature of It is apparent from Eqs. (14) and (17) that as the difference
250 ◦ C, it is assumed that the produced fluid is separated at 5 between the average reservoir temperature and the separator (or
bars (saturation temperature: 151.831 ◦ C). The separated brine is pinch point temperature) temperature tends to zero, available work
injected into the reservoir, and the steam is used to generate power. for conversion to electricity will also approach zero. The standard
The turbine inlet pressure is set equal to the separator pressure, and USGS method (Eq. (11)) will, however, yield a non-zero available
the condenser temperature is assumed to be 40 ◦ C. Evaluating the work; in this case, the only way to conform USGS method to reality
water properties along the saturation line, Eq. (14) yields: is to require conversion (utilization) efficiency to also tend to zero.
WAflash /˛ = 29.10 o C Stated somewhat differently, the conversion (utilization) efficiency
used in concert with Eq. (11) must be assumed to be a function of
The available work given by Eq. (14) is thus only about 45% of the reservoir temperature and the “abandonment temperature”. To
that obtained using Eq. (11). avoid ambiguities with a temperature-dependent conversion (uti-
The produced brine from the geothermal reservoir with an aver- lization) efficiency, it is recommended that the available work be
age temperature of 150 ◦ C is used to heat a secondary working fluid computed by considering the power cycle (e.g., Eqs. (14) and (17)).
(isobutane), and the spent brine is injected back into the reservoir.
The turbine inlet pressure is assumed to be 20 bars; the satura-
tion temperature for isobutane at the latter pressure is 100.36 ◦ C 4. Thermal recovery factor
(NIST, 2010). Isobutane is assumed to be saturated at the turbine
inlet. With a temperature differential of 5 ◦ C at the pinch point, The thermal recovery factor Rg was introduced in Section 2 as
the temperature of the brine at the pinch point is 105.36 ◦ C. Eval- the ratio of the heat recovered at the wellhead, qw , to the heat
uating the isobutane properties using NIST tables (NIST, 2010) and stored in the reservoir, qR . The latter parameter depends on the per-
substituting in Eq. (17), there follows: meability structure (fracture vs. matrix, permeability anisotropy,
faulting, etc.), production and injection well depths and patterns,
WAbinary /˛ = 10.47 o C
and the thermal (heat and fluid recharge from depth) and hydraulic
The available work obtained from Eq. (17) is about 43% of that boundary conditions (recharge/discharge along the ground surface,
computed using Eq. (11). and assumed lateral boundaries for the geothermal reservoir). Since
many of the reservoir properties that affect the thermal recovery
3.4. Conversion (utilization) efficiency factor are likely to be poorly known until the reservoir has been pro-
duced for several years, specification of the thermal recovery factor
Modern turbo-generators are quite efficient; the dry turbine for a specific reservoir is more often than not a matter of conjec-
(turbines operating with dry steam or gas) efficiency is about 85% ture. A major goal of reservoir engineering- including well drilling
(DiPippo, 2008; Zarrouk and Moon, 2014). Geothermal turbines and testing, data collection and synthesis, and detailed reservoir
generally operate in the wet region. Presence of moisture lowers modeling—is to obtain reliable estimates of the thermal recovery
the efficiency; thus a one percent (1%) moisture will lower the effi- factor for a particular reservoir.
ciency by about 1% (DiPippo, 2008; Zarrouk and Moon, 2014). For Williams (2014) presents estimates of thermal recovery fac-
flash power plants, the efficiency is further reduced by the pres- tor based on both theoretical grounds and data from operating
ence of non-condensable gases. Including losses in the generator, hydrothermal fields, and suggests that the appropriate range for
it is suggested that a conservative value for conversion efficiency fracture-dominated geothermal reservoirs is from 0.08 to 0.20. For
of 70% to 80% be used together with Eqs. (14) and (17) to estimate sediment-hosted reservoirs, the suggested range is slightly higher
the electrical generating capacity of a geothermal reservoir. from 0.10 to 0.25. According to Williams (2014), if a permeable
The illustrative examples presented in Section 3.3 show that reservoir exists, then Rg is non-zero; we agree.
the available work calculated using the standard USGS method During the exploration phase and prior to geothermal well
is much higher than that obtained considering the actual power drilling and testing, it will not in general be possible to infer the
cycles. Thus, if Eq. (11) is used to compute the available work, the presence of a permeable reservoir. Indeed, many hydrothermal
conversion efficiency will need to be reduced from 70% to 80% to exploration projects (e.g., Valles Caldera, Rye Patch, Lee Hot Springs,
no more than about 40%. Dixie Meadows, Newberry, Honey Lake, Mangakino) have been
As an example, consider a geothermal turbo-generator supplied abandoned due to lack of sufficient permeability (see also Grant,
with 5-bar steam and a condenser temperature of 40 ◦ C; in this 2015). Ultra low formation permeability usually results in unac-
case, the available power per kg/s of steam from Eq. (14) is about ceptably large pressure drop and uneconomic discharge rates. Thus,
0.625 MW. With a 70% conversion efficiency, about 2.3 kg/s will be the possibility of the thermal recovery factor being zero cannot
required to generate 1 MWe; this value for steam consumption is be discounted during the exploration phase. Therefore, we dis-
close to that used by modern geothermal power plants. Comparing agree with Williams (2014) that a non-zero Rg is appropriate for
available work computed from Eqs. (11) and (14), the appropriate “resource” evaluation. In accordance with our earlier work (Garg
conversion efficiency for use with the standard USGS method would and Combs, 2010), it is suggested that the proper range for ther-
be about 31.3%. mal recovery factor is from 0 to 0.20 (the latter value is believed to
DiPippo (2004) has analyzed the conversion efficiency for sev- be the maximum credible value based on world-wide experience
eral operating ORC power plants. Although second law efficiencies with production from liquid-dominated reservoirs). Unfortunately,
154 S.K. Garg, J. Combs / Geothermics 55 (2015) 150–158

in most instances prior to geothermal well drilling and testing, estimate of the minimum temperature may be computed using the
the minimum thermal recovery factor is assumed to be non-zero; average regional temperature gradient.
this more often than not leads to overly optimistic projections of During the exploration phase, chemical geothermometers
electrical generating capacity. provide the best estimate of the possible range for reservoir tem-
peratures. Shallow temperature gradient data, in conjunction with
5. Electric generating capacity chemical geothermometers, may then be used to estimate the
depth range for the geothermal reservoir. A minimum estimate of
The parameters required for the computation of electric capac- possible reservoir area may be obtained from the distribution of
ity with the “heat in place” method are provided in Table 2; these hot springs/fumaroles and high temperature gradient areas. Geo-
parameters can be divided into two groups. The second of these physical data (e.g., resistivity surveys) or geological mapping (e.g.,
groups (Group 2 Parameters) contains parameters whose value does surface alteration, faulting, etc.) may be useful for estimating the
not either vary substantially from case to case (volumetric heat probable maximum geothermal reservoir area. As discussed in Sec-
capacity, power plant or project life, power plant load factor) or tion 4, the appropriate range for the thermal recovery factor during
can be specified sufficiently accurately using available engineer- the exploration phase is from 0 to 0.20. The assumed value for the
ing data (reference temperature and conversion efficiency; see also minimum thermal recovery factor has a major influence on the pre-
Sections 2 and 3). The situation is completely different as far as the dicted electrical power capacity at the 90% confidence level (Garg
first group (Group 1 Parameters) of parameters is concerned, and and Combs, 2010). On the other hand, the power capacity at the 10%
estimates for these parameters are highly dependent on the specific confidence level remains more or less unaffected by the changes
phase of development of the geothermal reservoir. in the minimum values for the recovery factor and the reservoir
Specification of statistical distributions for the parameters in thickness.
the first group (reservoir area, reservoir depth, reservoir thickness, A geothermal well does not penetrate a region of uniform per-
reservoir temperature, thermal recovery factor) is at best a difficult meability. Very often a geothermal well is completed with a large
task and as demonstrated below is highly dependent on the stage of open-hole section (100 to 1000 or more meters); actual production
development of a geothermal system. Reservoir area, thickness, and usually comes from one or more feed zones with a thickness of the
temperature are required to compute the “heat in place”. Reservoir order of 1 to 100 m. In any event, during the exploration phase, the
depth, usually ignored in Monte Carlo simulations, is important in reservoir thickness should be assumed to vary within a rather wide
that it determines the depth to which the wells must be drilled in range (say between 100 and 2000 m).
order to access the geothermal resource. Thermal recovery factor A triangular probability distribution is assumed for parameters
is needed to compute the fraction of “heat in place” that may be for which maximum, mode, and minimum values can be estimated.
recovered using a system of production and injection wells. If only minimum and maximum values are available, then a rect-
angular (i.e., uniform) probability distribution is used. It should
5.1. Exploration phase be pointed out that any particular choice or prescription of the
distribution of parameter probability over the postulated range
Prior to deep well drilling and testing, data that may be used to (minimum to maximum) will also impact the cumulative results,
estimate reservoir parameters include shallow temperature gradi- and that this issue is not well-studied. Presumably, different dis-
ent surveys (temperature data from heat flow holes usually less tributions (uniform vs. triangular vs. Gaussian vs. log-normal, etc.)
than 100–200 m in depth), chemical geothermometer values for will give different answers; thus, providing additional uncertainty
fluid samples from any hot springs/fumaroles, and surface alter- in the estimation process.
ation (e.g., sinter, carbonate) surveys. Surface alteration, if present, Adoption of the above procedure for specifying parameter
can provide an indication of probable reservoir temperatures; as values during the exploration phase is liable to yield a wide dis-
an example, presence of sinter indicates fluid temperatures in tribution for estimated electrical capacity with a relatively large
excess of about 175 ◦ C. Chemical geothermometer data, if available, ratio between the capacity at 10% confidence level and that at
provide a reasonable first estimate of the reservoir temperature. 90% confidence level. The estimated capacity at 10% level would
Extrapolation of the high shallow temperature gradient data to almost certainly be higher than what will ultimately prove to be
great depth should be done with considerable care since the rela- the case. Similarly, the estimated capacity at the 90% confidence
tively high permeability associated with a geothermal reservoir will level will go up as we learn more about the geothermal reser-
tend to produce a near isothermal (and low temperature gradient) voir. The major goal of the resource estimation at this stage should
zone. Temperatures at depth, obtained by extrapolating shallow be to determine whether an exploitable geothermal resource
gradient data and not supported by any other evidence, should be exists and if it may be economically worthwhile to undertake a
regarded as an estimate of maximum reservoir temperature; a first program of geothermal well drilling and testing to reduce the
uncertainty in the geothermal reservoir parameters and thus
obtain a more reliable estimate of the probable resource megawatt
Table 2 capacity.
Parameters required for the calculation of the electric generation capacity using the
USGS volumetric “heat in place” method.
5.2. Well drilling and testing phase
Group 1 parameters
Reservoir area (km2 )
Drilling and testing of geothermal wells is essential for obtain-
Reservoir depth (m)
Reservoir thickness (m) ing reliable estimates of important reservoir parameters. Downhole
Reservoir temperature (◦ C) temperature and pressure surveys in deep wells can be used to
Thermal recovery Factor (%) establish (1) depth to the top of the geothermal reservoir (defined
here as the transition from the linear conductive thermal gradi-
Group 2 parameters
Volumetric heat capacity (kJ/m3 -K)
ent interval in the well to either a near isothermal or low thermal
Reference temperature (◦ C) gradient interval), (2) formation temperature, and (3) formation
Conversion efficiency (%) pressure. Fluid samples from discharging geothermal wells may be
Plant or project life (years) used to obtain chemical geothermometers temperatures; gener-
Plant load factor (%)
ally speaking, different chemical geothermometers yield a range
S.K. Garg, J. Combs / Geothermics 55 (2015) 150–158 155

of temperatures. Pressure transient tests (and in particular pres-


sure interference tests) and tracer tests can be used to establish
reservoir continuity in the region penetrated by the wells, and to
obtain the probable area of the permeable zone. Drilling records
(“mud losses”) are also helpful in determining depths of permeable
horizons. At the end of the initial well drilling and testing phase, it
should be possible to define the following parameters with a high
degree of confidence:

(1) Reservoir depth: Temperature profiles in wells (i.e., transition


from conductive to convective profiles) should be used to esti-
mate the minimum and maximum depths to the top of the
permeable zone.
(2) Reservoir temperature: Measured temperatures in wells
together with chemical geothermometry may be employed to
define the range of probable temperatures for the “hot region”
of interest in the geothermal system.
(3) Reservoir area: Since it is unlikely that the entire reservoir
area with elevated temperatures has been penetrated by the
geothermal wells, it is almost certain that the minimum reser-
voir area is greater than that investigated by drilling. Estimates
of maximum reservoir area may be obtained from pressure
transient data, geophysical surveys (e.g., electrical resistivity),
and geological mapping (e.g., alteration surveys).
(4) Reservoir thickness: Estimates of minimum and maximum
geothermal reservoir thickness may be obtained by examining
the convective zone from downhole temperature surveys.

Provided geothermal well drilling and testing has shown ade-


quate well productivity, it is justified at this stage to assume a Fig. 2. Sumikawa–Ohnuma area. Cross-sections AA and BB are shown in
non-zero minimum value (say 0.05) for the thermal recovery factor. Figs. 3 and 4, respectively.
At the end of geothermal well testing, it should thus be possible
to place much closer limits on reservoir parameters (i.e., Group 1
and Sarmiento (2005) postulated that while numerical simulation
Parameters) than those in the exploration phase. It need hardly be
is more sophisticated than the volumetric method, the latter can
stressed that even after initial geothermal well drilling and testing,
be readily conducted in a rigorously probabilistic way while the
it will only be possible to narrow (but not eliminate) the uncer-
former cannot; in principle, however, there is no reason why a
tainty in reservoir properties. Use of these “parameter ranges” will
Monte Carlo process cannot be used in conjunction with “detailed
result in a considerably narrower distribution of probable electric
numerical reservoir model” to assess the impact of uncertainty in
megawatt capacity than that obtained in the exploration phase. A
reservoir parameters on the probable electrical megawatt capacity.
measure of the success of the geothermal well drilling and testing
Availability of production and injection history as well as
would be a reduction in the ratio of the estimated electric megawatt
detailed reservoir modeling are essential for narrowing the possible
capacity at the 10% confidence level to that at the 90% confidence
ranges for important reservoir parameters (i.e., Group 1 Parameters),
level.
and in turn for obtaining a well-constrained estimate of the future
electrical megawatt capacity of the geothermal reservoir.
5.3. Production and injection phase

Following a positive outcome from the initial geothermal well 5.4. An illustrative example: Sumikawa geothermal field, Japan
drilling and testing program, the production and injection well-
field must be developed. The knowledge of reservoir parameters The Sumikawa geothermal field is located in the Hachimantai
increases with each new geothermal well that is drilled and tested. volcanic area, Honshu, Japan. Pritchett et al. (1989) described the
However, a production/injection history (and consequent changes natural fluid circulation system at Sumikawa based on static pres-
in reservoir temperature, pressure, and fluid state) is essential sure and temperature logs in several wells drilled in the field, and
for understanding important geothermal reservoir processes such a limited amount of pressure transient testing. It is worth noting
as boiling due to a reduction in pressure, recharge from the that the latter study was published after several deep wells had
boundaries, and possible short-circuiting between production and been drilled and tested but prior to the start of power production
injection wells. in 1995.
At this stage, a detailed geothermal reservoir model may be The Sumikawa geothermal field is located in the western part of
constructed by calibrating model predictions with available mea- the area shown in Fig. 2; to the east, the Ohnuma geothermal power
surements (e.g., temperature and pressure measurements, surface station has been generating about 10 MWe since 1974. The area lies
heat and mass flows, production/injection induced changes in the within a north–south graben; the Sumikawa field is located along
reservoir, etc.). The calibrated model may then be used to fore- the western edge of the graben. Based on drilling logs, the geologic
cast the electrical megawatt capacity of the reservoir. The current sequence at Sumikawa consists of the following formations (see
practice in the geothermal industry is to develop “determinis- Figs. 3 and 4):
tic” reservoir models. In large part, this practice is driven by the
“expert” time required to construct a geothermal reservoir model (1) ST formation: Surficial andesitic tuffs, lavas, and pyroclastics of
that is in accord with most of the known facts. Although Sanyal recent origin.
156 S.K. Garg, J. Combs / Geothermics 55 (2015) 150–158

West East South North


1500 1500
B

A A’ B’
1000
1000 “ST”
“ST”
“LS”

500 “DA”
“DA” HM-3
500
“LS”
“MV” S-3
“MV”

z (m ASL)
“MV” 0
z (mASL)

S-3
“AA”
0 “MV” “AA”
S-2 SB-2
SM-2
–500 S-4

SB-3 KY-1 “AA”


–500 “GR” S-4
“AA”
SA-2
SN-5 SA-4 KY-2 –1000 “GR” SB-1

SA-1
“AA” “GR”
–1000 SB-1
“GR”
–1500

“GR”
Fault?

SC-1
–1500
–2000
Fault?

0 500 1000 1500 2000 2500 3000 3500

y (m North)

–2000
–1000 –500 0 500 1000 1500 2000 Fig. 4. A north–south cross-section across Sumikawa geothermal field.

x (m East)

Fig. 3. An east–west cross-section across Sumikawa geothermal field.


2000 m, and the average temperature is between 260 ◦ C and 320 ◦ C.
The Sumikawa power plant has a nominal capacity of 50 MW. The
initial design for the power plant called for the supply of 5-bar
(2) LS formation: Lake sediments; Pleistocene tuffs, sandstones, silt- steam. Ignoring any pressure drop between the separator and the
stones and mudstones. turbine inlet, the separator temperature (saturation temperature
(3) DA formation: Pliocene dacites, dacitic tuffs, and breccias. corresponding to 5 bar) is 151.831 ◦ C.
(4) MV formation: Interbedded Miocene dacites volcanic rocks and The parameters used to compute the megawatt capacity for the
marine shales and sediments. Sumikawa geothermal field are listed in Table 3. Triangular distri-
(5) AA formation: Altered andesitic rocks. butions are assumed for reservoir area and temperature. Reservoir
(6) BA formation: Crystalline intrusive rocks (mainly granodiorite thickness and thermal recovery factor are represented by rectangu-
and diorite). lar (uniform) distributions. All the other parameters have constant
values. Cases 1 and 2 were evaluated using the new formulation
The lake sediments have low permeability and separate the (Eq. (14)), and the original USGS formulation (Eq. (11)) is used
overlying cold groundwater from the hot geothermal reservoir for Case 3. Since the well productivity had been confirmed prior
below. Also, the MV formation is an aquitard. Geothermal wells to Pritchett et al.’s study, a non-zero minimum recovery factor is
mainly feed from DA, AA, and upper BA formations. justified; thus, the recovery factor is taken to range from 0.05 to
Using the information presented by Pritchett et al. (1989), it is 0.20 for Case 1. To illustrate the effect of a zero minimum recover
reasonable to assume that the Sumikawa geothermal field has a factor on the predicted capacity, Case 2 was also run; other than
minimum area of 3 square kilometers, and a maximum area of 5 the minimum value for the recovery factor, parameters for Case 2
square kilometers. The formation thickness ranges from 1000 m to remain unchanged from those for Case 1. The recovery factor and

Table 3
Parameters for Sumikawa geothermal field. See text for description of the three cases.

Parameter Case 1 Case 2 Case 3

Area (km2 ) Min: 3 Same as Case 1 Same as Case 1


Mode: 4
Max: 5
Thickness (m) Min: 1000 Same as Case 1 Same as Case 1
Max: 2000
Reservoir temperature (◦ C) Min: 260 Same as Case 1 Same as Case 1
Mode: 290
Max: 320
Volumetric heat capacity (kJ/m3 -◦ C) 2700 Same as Case 1 Same as Case 1
Thermal recovery factor Min: 0.05 Min: 0.00 Min: 0.08
Max: 0.20 Max: 0.20 Max: 0.20
Separator temperature (◦ C) 151.831 Same as Case 1 Not applicable
Condenser temperature (◦ C) 40 Same as Case 1 Not applicable
Ambient (rejection) temperature (◦ C) Not applicable Not applicable 15
Conversion efficiency 0.75 Same as Case 1 0.40
Plant life (years) 30 Same as Case 1 Same as Case 1
Plant capacity (Load) Factor 0.95 Same as Case 1 Same as Case 1
S.K. Garg, J. Combs / Geothermics 55 (2015) 150–158 157

Table 4 distribution for the electrical megawatt capacity. After the start of
Megawatt capacity vs. probability for Sumikawa geothermal field.
large-scale production and injection operations and the availability
Minimum MW capacity with a probability > Case 1 Case 2 Case 3 of data on production-induced changes in the geothermal reservoir,
90% 33 10 51 it should in principle be possible to further constrain the range of
50% 65 53 82 probable electrical megawatt capacity.
10% 110 104 127 The USGS volumetric “heat in place” method together with
10% (MW)/90% (MW) 3.3 10 2.5 Monte Carlo simulations is an important tool for assessing the elec-
trical capacity of a geothermal reservoir. The method should be
Cumulative Probability of Energy Reserves
modified in several respects to obtain valid results.
1
case #1 Firstly, available work and conversion efficiency should be eval-
0.9 case #2
case #3 uated using parameters for the power cycle (binary or flash) that
0.8 may be used for power conversion.
Secondly, specification of the probability distributions of the
0.7
reservoir parameters should take into account the stage of the
Cumulative Probability

0.6 development of the project. Very often, these parameters are pre-
scribed based on data from other geothermal reservoirs.
0.5 Thirdly, at least during the early exploration phases and prior to
0.4
deep well drilling and testing, the minimum thermal recovery fac-
tor should be assumed to be zero.
0.3

0.2
At present, there exist insufficient data in the public domain
to specify probability distributions for most reservoir parameters.
0.1 Moreover, conditions vary widely between and within the vari-
ous geothermal provinces around the world. For this reason, it is
0
0 50 100 150 essential that as far as possible, actual field data should be used
Power (MW) when prescribing reservoir parameters. Without data-driven reser-
voir parameters, use of Monte Carlo simulations is liable to generate
Fig. 5. Cumulative probability of energy reserves for Cases 1–3. See text for descrip-
tion of different cases. only unreliable estimates of reservoir megawatt capacity.

utilization efficiency for Case 3 are the ones recommended by Acknowledgments


Williams (2014).
A Monte Carlo procedure was employed to predict the megawatt We thank Malcolm Grant and Joel Renner for their careful read-
capacity for the three cases (Table 4). A total of 100,000 simulations ing of the original manuscript and suggesting improvements.
were used in each case to compute the cumulative probability dis-
tribution; results are illustrated in Fig. 5 and Table 4. Comparing References
Cases 1 and 3, it is apparent that the original USGS method predicts
a larger capacity at the 90% confidence level by about 55% than the Brook, C.A., Mariner, R.H., Mabey, D.R., Swanson, J.R., Guffanti, M., Muffler, L.J.P., 1979.
Hydrothermal convection systems with reservoir temperatures ≥90 ◦ C. In: Muf-
new formulation. The main effect of a zero minimum recovery fac- fler, L.J.P. (Ed.), Assessment of Geothermal Resources of the United States—1978,
tor (Case 2) is to yield a much smaller capacity at the 90% confidence 790. U.S Geological Survey Circular, 170p.
level compared to Case 1. In a certain sense, a zero minimum recov- DiPippo, R., 2004. Second Law assessment of binary plants generating power from
low-temperature geothermal fluids. Geothermics 33, 565–586.
ery factor encodes a lack of knowledge regarding the productivity DiPippo, R., 2008. Geothermal Power Plants: Principles, Applications, Case Studies
of as yet undrilled wells. and Environmental Impact, second ed. Elsevier, Amsterdam, 493p.
Published data (Fuchino, 2000; Kawazoe and Shirakura, 2005; Fuchino, H., 2000. Status of geothermal power generation in Japan. In: Proceedings
World Geothermal Congress 2000. Tohoku and Kyushu, Japan, 6p.
Sugino and Akeno, 2010) indicate that the Sumikawa power plant Garg, S.K., Combs, J., 2010. Appropriate use of USGS volumetric “Heat in Place”
generated about 40 MWe between 1998 and 2007; this is about method and Monte Carlo Calculations. In: Proceedings 34th Workshop on
20% less than the nominal plant capacity. The predicted capacity Geothermal Reservoir Engineering. Stanford University, Stanford, CA, 7p.
Garg, S.K., Combs, J.,2011. A Reexamination of USGS volumetric “Heat in Place”
at the 90% confidence level for Case 1 (Case 3) at 33 MW (51 MW)
method. In: Proceedings 35th Workshop on Geothermal Reservoir Engineering.
is somewhat smaller (larger) than the actual power production. At Stanford University, Stanford, CA, 5p.
the 50% confidence level, predicted capacity for Case 1 (65 MW) is GeothermEx, Inc., 2004. New Geothermal Site Identification and Qualifica-
much closer to the actual one than that obtained for Case 3 (82 MW). tion, Consultant Report for the California Energy Commission PIER Report
No. P500-04-051, April 2004, located at http://www.geothermex.com/CEC-
PIER Reports.htm.
6. Summary and conclusions Grant, M., 2015. Resource assessment, a review, with reference to the Australian
code. In: Proceedings World Geothermal Congress 2015, Melbourne, Australia,
6p.
Geothermal reservoir assessment and the resulting prediction Kawazoe, S., Shirakura, N., 2005. Geothermal power generation and direct use in
of its electrical megawatt capacity should be regarded as a con- Japan. In: Proceedings World Geothermal Congress 2005, Antalya, Turkey, 7p.
Muffler, P., Cataldi, R., 1977. Methods for regional assessment of geothermal
tinuing process—from the early exploration phase to the time
resources. U.S. Geological Survey Open-File Report 77-870, 78p.
when the reservoir becomes depleted. During the early exploration Nathenson, M., 1975a. Some reservoir engineering calculations for the vapor-
phase, estimates of important reservoir parameters are poorly con- dominated System at Larderello, Italy. U.S. Geological Survey Open-File Report
75-142, 47p.
strained; application of the USGS “heat in place” evaluation method
Nathenson, M., 1975b. Physical factors determining the fraction of stored energy
along with Monte Carlo simulations yields a rather wide distribu- recoverable from hydrothermal convection systems and conduction-dominated
tion for the probable electrical megawatt capacity. As initial deep areas. U.S. Geological Survey Open-File Report 75-525, 38p.
geothermal wells are drilled and tested, it becomes possible to Nathenson, M., Muffler, L.J.P., 1975. Geothermal resources in hydrothermal convec-
tion systems and conduction-dominated areas: In: White, D.E., Williams, D.L.
refine estimates of reservoir parameters (i.e., narrow the range of (Eds.), Assessment of Geothermal Resources of the United States—1975, 726.
possible values), and in turn obtain a much narrower probability U.S Geological Survey Circular, 160p.
158 S.K. Garg, J. Combs / Geothermics 55 (2015) 150–158

NIST, 2010. Thermophysical Properties of Fluid Systems, Available at Sugino, H., Akeno, T., 2010. 2010 Country update for Japan. In: Proceedings World
http://webbook.nist.gov/chemistry/fluid/. Geothermal Congress 2010, Bali, Indonesia, 7p.
Pritchett, J.W., Garg, S.K., Maki, H., Kubota, K., 1989. Hydrology of the Sumikawa Williams, C., 2014. Evaluating the volume method in the assessment of identified
geothermal prospect, Japan. Energy Sources 11, 251–262. geothermal resources. Geotherm. Resour. Counc. Trans. 38, 967–974.
Sanyal, S.K., Sarmiento, Z.F., 2005. Booking geothermal energy reserves. Geotherm. Zarrouk, S., Moon, H., 2014. Efficiency of geothermal power plants: a worldwide
Resour. Counc. Trans. 29, 467–474. review. Geothermics 51, 142–153.

You might also like