You are on page 1of 11

Energy 35 (2010) 4955e4965

Contents lists available at ScienceDirect

Energy
journal homepage: www.elsevier.com/locate/energy

Numerical modeling of aquifer thermal energy storage system


Jongchan Kim a, b, Youngmin Lee a, *, Woon Sang Yoon c, Jae Soo Jeon c, Min-Ho Koo b, Youngseuk Keehm b
a
Korea Institute of Geoscience and Mineral Resources, Geothermal Resources Department, 92 Gwahang-no, Yuseong-gu, Daejeon 305-350, South Korea
b
Kongju National University, Department of Geoenvironmental Sciences, 182 Singwan-dong, Gongju-si, Chungnam 314-701, South Korea
c
nexGeo Inc., 134-1 Garak 2-dong, Songpa-gu, Seoul 138-807, South Korea

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

Article history: The performance of the ATES (aquifer thermal energy storage) system primarily depends on the thermal
Received 25 February 2010 interference between warm and cold thermal energy stored in an aquifer. Additionally the thermal
Received in revised form interference is mainly affected by the borehole distance, the hydraulic conductivity, and the pumping/
9 August 2010
injection rate. Thermo-hydraulic modeling was performed to identify the thermal interference by three
Accepted 19 August 2010
Available online 22 September 2010
parameters and to estimate the system performance change by the thermal interference. Modeling
results indicate that the thermal interference grows as the borehole distance decreases, as the hydraulic
conductivity increases, and as the pumping/injection rate increases. The system performance analysis
Keywords:
Thermal energy
indicates that if h (the ratio of the length of the thermal front to the distance between two boreholes) is
ATES lower than unity, the system performance is not significantly affected, but if h is equal to unity, the
Korea system performance falls up to w22%. Long term modeling for a factory in Anseong was conducted to
Thermal interference test the applicability of the ATES system. When the pumping/injection rate is 100 m3/day, system
performances during the summer and winter after 3 years of operation are estimated to be w125 kW and
w110 kW, respectively. Therefore, 100 m3/day of the pumping/injection rate satisfies the energy
requirements (w70 kW) for the factory.
Ó 2010 Elsevier Ltd. All rights reserved.

1. Introduction mainly two operating types in the ATES system: cyclic mode and
continuous mode (Fig. 1). In the cyclic mode, pumping and injection
Geothermal systems using heat stored in the underground have wells are switched by the season, while not switched in the
been used for cooling and heating of buildings in many countries continuous mode. The cyclic mode has an advantage of high system
such as the United States [1e4], European countries [5e10], and performance by storing thermal energy in the aquifer, and
other countries [11e14] since 1970s. Recently, thermal energy a disadvantage of the complicated system design. On the other
storage (TES) systems become more popular in the world due to the hand, the continuous mode has a simpler system design but low
depletion of fossil fuel and global warming [15]. The TES systems system performance due to the limited temperature range.
are generally divided into a closed system (e.g., borehole thermal Understanding heat transfer in the aquifer plays a key role in
energy storage: BTES), and an open system (e.g., aquifer thermal designing an ATES system [16,17]. Researches on analytical methods
energy storage: ATES). Due to directly using groundwater with for heat transfer have been conducted by Sauty et al. [18], Uffink [19],
relatively high volumetric heat capacity, the ATES system has the Voigt and Haefner [20], Krarti and Claridge [21], Yang and Yeh [22],
higher system performance than the BTES system and any other and Stopa and Wojnarowski [23]. Kangas and Lund [17] pointed out
system using low temperature geothermal heat. In the ATES that analytical solution is applicable only to the simplest cases or to
system, the contamination and depletion of groundwater can be qualitative estimations. Numerical simulations, therefore, should be
minimal, because the water circulated from underground to a heat used for realistic understanding of heat transfer in complex
exchanger is immediately re-injected though the injection well into geological and hydrological characteristics [24]. The numerical
the aquifer [12]. modeling of the ATES system has been conducted from 1970s [25],
An ATES system usually consists of two or more wells to store and two-dimensional numerical simulations were carried out
warm and/or cold thermal energy in the aquifer, and there are before 1990s [26,27]. Recently, with the advance of modern
computing power, three-dimensional numerical modeling become
widely used such as in Dickinson et al. [8], Lee [13], Tenma et al. [28],
* Corresponding author. Tel.: þ82 42 868 3069; fax: þ82 42 868 3358. and so on. The main objective of the numerical modeling for the
E-mail address: ymlee@kigam.re.kr (Y. Lee). ATES system before 1990s was to predict heat transfer in the model

0360-5442/$ e see front matter Ó 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.energy.2010.08.029
4956 J. Kim et al. / Energy 35 (2010) 4955e4965

Nomenclature S specific storage coefficient (m1)


t time (s)
cf specific heat of fluid (J/kg K) T temperature (K)
Cf volumetric heat capacity of fluid (J/m3 K) T0 uppermost temperature (K)
Cs volumetric heat capacity of the aquifer (J/m3 K) TL lowermost temperature (K)
h hydraulic head (m) Tz temperature at any depth z (K)
Kf hydraulic conductivity (m/s) u velocity of fluid (m/s)
Ks thermal conductivity of the aquifer (W/m K) z depth (m)
L length of vertical section (m) DT temperature drop (K)
P ground thermal load (W)
Pe Peclet number Greek symbols
q flow rate of water (m3/s) h ratio of the distance of the thermal front to the
Qh heat source (W/m3) distance between two boreholes
Qf source/sink mass flux (s1) rf density of fluid (kg/m3)

domain by injected water through a single well [16,18,26], and it is ATES system. Because no analytical solution is available for the
now widening into predicting recovery temperature by heat loss ATES system with pumping/injection rate switched by season, we
[13,16], injection depth in the aquifer [13,28], injection rate [13,28], used the simple case of one-dimensional steady-state conductivee
injection time [28], and regional groundwater flow [17]. convective heat transfer described by Bredehoeft and Papadopulos
The system performance of the ATES system mainly depends on [31] to verify the numerical code. The one-dimensional steady-
the thermal interference between stored warm and cold thermal state equation is
energy in the aquifer [29]. In addition, the degree of the thermal    
interference is primarily determined by the distance between two v2 T=vz2  cf rf u=Ks ðvT=vzÞ ¼ 0 (3)
boreholes, the hydraulic conductivity, and the pumping/injection
rate [30]. However, the thermal interference has not been rigor- The solution to this problem is equation (4)
ously and systematically studied in previous researches. Therefore,
we performed three-dimensional thermo-hydraulic modeling of Tz ¼ T0 þ ðTL  T0 Þf ðb; z=LÞ (4)
a cyclic mode to identify the thermal interference by three where f ðb; z=LÞ ¼ ½expðbz=LÞ  1=½expðbÞ  1, and b ¼ cf rf uL=Ks .
parameters above mentioned, and analyzed the thermal interfer-
ence from quantitative modeling results. The system performance
of the ATES system was then estimated in terms of the degree of the
thermal interference. Finally, we evaluated applicability of the ATES
system for a case study area.

2. ATES system modeling

2.1. Governing equations

Two governing equations, fluid flow and heat transfer equations,


are used for thermo-hydraulic modeling of the ATES system. Fluid
flow in a porous medium is described by Darcy’s law as
h i
Sðvh=vtÞ þ V$  Kf Vh ¼ Qf (1)

Heat transfer equation which includes conduction and convec-


tion can be expressed as
h i
Cs ðvT=vtÞ þ V$  Ks VT þ Cf uT ¼ Qh (2)

In this study, Comsol MultiphysicsÒ software which is


a commercial finite element program was used for simulating the

Fig. 1. Schematic diagram of the ATES system: (a) cyclic mode, and (b) continuous Fig. 2. Comparisons of numerical solutions and analytical solutions for the one-
mode. HEX represents a heat exchanger. dimensional steady-state conductiveeconvective heat transfer.
J. Kim et al. / Energy 35 (2010) 4955e4965 4957

The numerical solutions for u ¼ 107, 108, and 109 m/s cases Table 1
agree well with the analytical solutions (Fig. 2). The maximum Physical and thermal properties of fluid and aquifer for thermo-hydraulic
modeling.
relative error in the numerical solutions is 1.15  104%.
Property Value
2.2. Model domain Porosity 0.3
Density of fluid 1000 kg/m3
Density of aquifer 2000 kg/m3
Model dimension is 160 m  100 m  50 m in the x, y, and z Specific heat of fluid 4200 J/kg K
directions (Fig. 3). The number of elements in the FEM (finite Specific heat of aquifer 800 J/kg K
element model) is 25,940 and the degree of the freedom is 219,870. Thermal conductivity of fluid 0.64 W/m K
Two boreholes for storing thermal energy are set in the model Thermal conductivity of aquifer 1.60 W/m K
Compressibility of fluid 4.4  1010 Pa1
domain. The depth of each borehole is 50 m, and its radius is
Compressibility of aquifer 1.0  108 Pa1
0.035 m. Warm and cold water stored in the aquifer is extracted or
injected at the bottom of wells.
water of 35  C from a heat pump is injected into the aquifer through
the warm well. On the contrary, in the winter (3 months), warm
2.3. Model parameters
water stored in the aquifer during the summer is extracted for
heating the building, and cold water of 5  C from the heat pump is
Physical and thermal properties of fluid and aquifer for thermo-
injected into the aquifer through the cold well. In the spring and the
hydraulic modeling are listed in Table 1. Since the model domain is
autumn (each 3 months), the ATES system is not operated because
assumed to be an alluvial aquifer of which the porosity is relatively
heating and cooling is not required.
high, the thermal conductivity and the density of the aquifer are
smaller than the representative values of crystalline rocks. Fluid
2.6. Model assumptions
properties are assumed to be equal to those of pure water.

The following conditions are assumed to simplify ATES system


2.4. Boundary and initial conditions
modeling: (1) the model domain is a homogenous, isotropic, and
confined aquifer; (2) the temperatures of fluid and aquifer are
Temperatures at the top and the bottom of the model domain
always in the equilibrium state; (3) the density and viscosity of fluid
are 14  C and 15  C, respectively, and sides are assumed to be
are always constant; (4) the total groundwater extracted for cooling
thermally insulated. The initial temperature in the model domain
or heating of a building is completely re-injected into the aquifer
linearly increases from the top to the bottom to express the thermal
through the injection well; (5) no regional groundwater flow in the
equilibrium state prior to thermal injection. In the hydraulic
model domain is assumed.
domain, the top boundary is water table of which the hydraulic
head is equal to the model elevation and the bottom and sides are
no fluid flux boundary due to an assumption of no regional 3. Results
groundwater flow. The initial head in the model domain is equal to
the model elevation. A total of 72 three-dimensional thermo-hydraulic modeling
scenarios are considered to identify the degree of the thermal
interference in the ATES system with 3 parameters: 1) the distance
2.5. Operation conditions
between two boreholes; 2) the pumping/injection rate; 3) the
hydraulic conductivity as listed in Table 2. To quantify the degree of
In this study, a cyclic mode of the ATES system using two bore-
the thermal interference, the ratio (h) of the length of the thermal
holes is simulated. Warm thermal energy and cold thermal energy,
front to the distance between two boreholes is defined. For instance,
therefore, are stored separately in the aquifer through each borehole,
if the value of h is equal to unity, the thermal front by injected water
and pumping and injection are switched by the operation season to
propagates completely from the injection well to the pumping well.
store separately warm and cold thermal energy in the aquifer.
If the value of h becomes higher, the thermal interference problem
In the summer (3 months), cold water stored in the aquifer
could be severe in the aquifer, and therefore it could be a major
during the winter is extracted for cooling a building, and warm
factor for the system performance decrease. Finally, the changes in
system performance by various h values are evaluated to investigate
effects of the thermal interference on the ATES system.

3.1. Distance of boreholes

Distances of 40 m, 60 m and 80 m between two boreholes are


considered to evaluate the thermal interference by the distance of
two boreholes. Fig. 4 shows temperature distributions in the
aquifer after 3-month injection of warm water (w35  C) through
the warm well (the right well in each diagram), when distances
between two boreholes are (a) 40 m, (b) 60 m and (c) 80 m. The
pumping/injection rate and the hydraulic conductivity are fixed

Table 2
Values of three major parameters for thermo-hydraulic modeling.

Parameter Value
Distance between two boreholes (m) 40, 60, 80
Fig. 3. The three-dimensional finite element model domain for thermo-hydraulic
Hydraulic conductivity (m/s) 103, 104, 105, 106
modeling. Warm water (w35  C) is injected into the right well in the summer, and cold
Production/injection rate (m3/day) 50, 100, 150, 200, 250, 300
water (w5  C) is injected into the left well in the winter.
4958 J. Kim et al. / Energy 35 (2010) 4955e4965

Fig. 4. Simulated temperatures of the ATES system after 3 months of operation showing the effect of distance between two boreholes: (a) 40 m, (b) 60 m, and (c) 80 m. The
hydraulic conductivity and the pumping/injection rate are fixed with 104 m/s, and 150 m3/day, respectively. Arrows represent normalized fluid velocity vector.

with 150 m3/day and 104 m/s, respectively. Fig. 5 presents vari- s and (d) 106 m/s. As shown in Figs. 4 and 5, the propagation of
ations of h values after 3-month injection of warm water (w35  C) the thermal front increases with decreasing the borehole
through the warm well as a function of the distance between two distance, since pressure gradient between two boreholes by
boreholes for various pumping/injection rates in the cases that injected water increases with decreasing the distance between
hydraulic conductivities are (a) 103 m/s, (b) 104 m/s, (c) 105 m/ two boreholes. Fig. 5 indicates that if the distance between two

Fig. 5. Variations of h (the ratio of the distance of the thermal front to the distance between two boreholes) of the ATES system after 3 months of operation as a function of the
distance between two boreholes with various pumping/injection rates. The hydraulic conductivities are (a) 103 m/s, (b) 104 m/s, (c) 105 m/s, and (d) 106 m/s, respectively.
J. Kim et al. / Energy 35 (2010) 4955e4965 4959

Fig. 6. Simulated temperatures of the ATES system after 3 months of operation showing the effect of production/injection rate: (a) 100 m3/day, (b) 200 m3/day, and (c) 300 m3/day.
The hydraulic conductivity and distance between two boreholes are 104 m/s, and 60 m, respectively. Arrows represent normalized fluid velocity vector.

boreholes is 40 m, the pumping/injection rate is 300 m3/day and interference by various pumping/injection rates. Fig. 6 shows
the hydraulic conductivity is over 106 m/s, the thermal front by temperature distributions in the aquifer after 3-month injection of
injected water propagates completely to the pumping well; the warm water (w35  C) through the warm well (the right well in
value of h is equal to unity. Therefore, in this case, separately each diagram), in the case that the pumping/injection rates are (a)
storing warm and cold thermal energy in the ATES system is 100 m3/day, (b) 200 m3/day and (c) 300 m3/day. The distance
probably disturbed by severe the thermal interference, and the between two boreholes and the hydraulic conductivity are fixed
ATES system performance could be fall down significantly. In with 60 m and 104 m/s, respectively. In this scenario, values of h
conclusion, it should be required for enough distance between are determined to be (a) 0.4 for 100 m3/day, (b) 0.6 for 200 m3/day
two boreholes to store thermal energy separately through warm and (c) 0.67 for 300 m3/day, respectively, which implies that the
and cold wells in the ATES system. thermal front propagates more with increasing the pumping/
injection rate. It is because pressure gradient between two bore-
3.2. Pumping/injection rate holes increases with increasing fluid source, i.e., injection flow rate.
Therefore, in the design of the ATES system, the determination of
The range of pumping/injection rates is set up from 50 m3/day the optimal pumping/injection rate is important for meeting the
to 300 m3/day at intervals of 50 m3/day to estimate the thermal requirement of thermal energy load.

Fig. 7. Simulated temperatures of the ATES system after 3 months of operation showing the effect of hydraulic conductivities: (a) 103 m/s, (b) 104 m/s, and (c) 105 m/s. The
distance between two boreholes and the pumping/injection rate are fixed with 60 m and 150 m3/day, respectively. Arrows represent normalized fluid velocity vector.
Fig. 8. Variations of h of the ATES system after 3 months of operation as a function of the hydraulic conductivity with various pumping/injection rates. The distances between two
boreholes are (a) 40 m, (b) 60 m, and (c) 80 m, respectively.

Fig. 9. Variations of fluid velocity of the ATES system as a function of the hydraulic conductivity with various pumping/injection rates. The distances between two boreholes are (a)
40 m, (b) 60 m, and (c) 80 m, respectively.
J. Kim et al. / Energy 35 (2010) 4955e4965 4961

Fig. 10. Variations of Peclet number (Pe) after 3 months of operation as a function of the hydraulic conductivity with various pumping/injection rates. The distances between two
boreholes are (a) 40 m, (b) 60 m, and (c) 80 m, respectively.

3.3. Hydraulic conductivity hydraulic conductivities are (a) 103 m/s, (b) 104 m/s and (c)
105 m/s. The distance between two boreholes and the pumping/
To investigate the thermal interference by transmissibility of injection rate are fixed with 60 m and 150 m3/day, respectively.
the aquifer, the hydraulic conductivities of 103 m/s, 104 m/s and From modeling results, values of h are (a) 0.5 for 103 m/s, (b) 0.47
105 m/s are considered. Fig. 7 shows temperature distributions in for 104 m/s and (c) 0.38 for 105 m/s, respectively. Fig. 8 presents
the aquifer after 3-month injection of warm water (w35  C) variations of h after 3-month injection of warm water (w35  C)
through the warm well (the right well in each diagram), when the through the warm well as a function of the hydraulic conductivity
for various pumping/injection rates, in the cases that distances
between two boreholes are (a) 40 m, (b) 60 m and (c) 80 m. As
illustrated in Figs. 7 and 8, the propagation of the thermal front
increases with increasing the hydraulic conductivity because the
hydraulic conductivity controls fluid velocity in the aquifer. Fig. 9
shows variations of fluid velocity as a function of the hydraulic
conductivity for various pumping/injection rates when distances
between two boreholes are (a) 40 m, (b) 60 m and (c) 80 m. Fluid
velocity between two boreholes increases with increasing the

Table 3
Comparative analysis of the system performance for case I.

Production/injection Hydraulic Distance h Relative system


rate (m3/day) conductivity between two performance
(m/s) boreholes (m)
Fig. 11. Thermal power as a function of elapsed time with various h values for case I.
200 104 40 1.0 7%
The hydraulic conductivity and the pumping/injection rate are 104 m/s, and 200 m3/
200 104 60 0.6 0.1%
day, and distances between two boreholes are 40 m, 60 m, and 80 m, respectively. In
200 104 80 0.4 Base case
this case, the values of h are estimated as 1, 0.6, and 0.4, respectively.
4962 J. Kim et al. / Energy 35 (2010) 4955e4965

The system performance increases with the decrease of h and the


increase of operation time. The increase of system performance
with operation time is due to the accumulation of thermal energy
in the aquifer as the operation continues. To analyze system
performance for case I, it was assumed that h of 0.4 is the base
case. If the value of h is unity, the system performance is w7%
lower than the base case, and if the value of h is 0.6, the system
performance is w0.1% lower than the base case after 5 years of
continuous operation (Table 3). For case II, the hydraulic conduc-
tivity of 104 m/s, the pumping/injection rate of 300 m3/day, and
the distances of 40 m, 60 m and 80 m between two boreholes
were considered. In this case, values of h are estimated as 1 for
40 m, 0.7 for 60 m and 0.5 for 80 m, respectively. Fig. 12 shows the
system performance as a function of time for various h values.
The system performance, as the same as case I, increases with the
Fig. 12. Thermal power as a function of elapsed time with various h values for case II.
The hydraulic conductivity and the pumping/injection rate are 104 m/s, and 300 m3/ decrease of h and the increase of operation time. In the analysis of
day, and distances between two boreholes are 40 m, 60 m, and 80 m, respectively. In case II, h of 0.5 was assumed to be the base case. If the value of h is
this case, the values of h are estimated as 1, 0.7, and 0.5, respectively. unity, the system performance is w22% lower than the base case,
and if the value of h is 0.7, the system performance is w1% lower
than the base case after 5 years of continuous operation (Table 4).
hydraulic conductivity (Fig. 9). As a result, the propagation of the
As shown in Figs. 11 and 12, the system performance during the
thermal front increases with increasing fluid velocity.
summer period is higher than during the winter period, because
difference between injection water temperature and pumping
3.4. Peclet number water temperature during the summer period (DT ¼ w20  C) is
higher than that during the winter period (DT ¼ w10  C). Finally,
To evaluate the heat transfer mechanism in the model domain, two system performance analyses suggest that if the thermal front
Pe analysis is performed. Pe means the ratio of heat transfer by propagates completely from the injection well to the pumping
convection to heat transfer by conduction. If Pe is higher than unity well (h ¼ w1), the system performance by the thermal interfer-
in the model domain, convection is more dominant than conduc- ence decreases up to 22%. In contrast, if the thermal front does not
tion. Fig. 10 shows Pe as a function of the hydraulic conductivity for propagate completely to pumping well (h < 1), the system
various pumping/injection rates when distances between two performance would not dropped down significantly. Therefore, the
boreholes are (a) 40 m, (b) 60 m and (c) 80 m. Since Pe ranges from value of h should be always lower than unity for the optimal ATES
w4 to w45, convection is obviously dominant against conduction system design.
for the heat transfer in the ATES system.

3.5. System performance

In this study, the system performance of the ATES system is


defined as the ground thermal load which is a function of mainly
temperature difference between injected and extracted fluids, and
fluid flow rate. Therefore the system performance analyses to
evaluate the effect of the thermal interference are carried out by
estimating the ground thermal load. The ground thermal load is
calculated by equation (5) as

P ¼ qrf cf DT (5)
The analysis of two cases for the system performance is per-
formed in terms of the thermal interference. For case I, the model
parameters are chosen as follows: the hydraulic conductivity and
the pumping/injection rate are fixed with 104 m/s and 200 m3/
day, respectively; distances between two boreholes are 40 m,
60 m, and 80 m. From the modeling, values of h are estimated as 1
for 40 m, 0.6 for 60 m, and 0.4 for 80 m, respectively. Fig. 11 shows
the system performance as a function of time for various h values.

Table 4
Comparative analysis of the system performance for case II.

Production/injection Hydraulic Distance h Relative system


rate (m3/day) conductivity between two performance
(m/s) boreholes (m)
300 104 40 1.0 22%
300 104 60 0.7 1%
300 104 80 0.5 Base case
Fig. 13. Location map of the study area (Anseong) and the synoptic station (Suwon).
J. Kim et al. / Energy 35 (2010) 4955e4965 4963

Table 5
Chemical analyses of the groundwater sampled in the study area.

Sample K Na Ca Mg Cl SO4 HCO3 F NO3 NH4 Fe Li


PW-1 (08/02/27) 2.90 18.5 18.3 3.44 15.8 1.86 115.9 0.09 26.8 0.26 0.01 0.02
PW-1 (08/04/01) 1.92 15.9 16.9 2.91 16.5 1.77 100.7 0.01 16.9 0.42 0.01 0.02
PW-1 (09/05/14) 1.38 24.0 12.2 2.03 8.5 1.32 79.3 0.03 14.1 0.03 0.44 0.01

Table 6 4. ATES system in Anseong city


Physical and thermal properties of fluid and aquifer for thermo-hydraulic
modeling of the ATES system in Anseong city.
The first ATES system in Korea was installed in Anseong city in
Property Value 2009. Anseong city is located in about 70 km south of Seoul, Korea
Porositya 0.3 (Fig. 13). The study area consists of a w50 m deep alluvial aquifer,
Density of fluid 1000 kg/m3 and granite as the basement rock below the aquifer. The alluvial
Density of aquifer 2000 kg/m3 aquifer is composed of unconsolidated clay and sand, and weath-
Specific heat of fluid 4200 J/kg K
Specific heat of aquifer 800 J/kg K
ered granite. Annual mean ground surface temperature observed in
Thermal conductivity of fluid 0.64 W/m K the Suwon synoptic station which is the nearest station from the
Thermal conductivity of aquifera 1.60 W/m K study area is w13.26  C [32]. Table 5 shows the result of the
Compressibility of fluid 4.4  1010 Pa1 chemical analyses of groundwater sampled in the study area.
Compressibility of aquifer 1.0  108 Pa1
The groundwater sampled in the study area is Na and HCO3
a
Measured values. dominant types, and is likely to be the typical characteristic of the

Fig. 14. Simulated temperatures of the ATES system during 15 years operation time (Anseong site). The distance of 80 m between two boreholes, the hydraulic conductivity of
2  104, and the pumping/injection rate of 100 m3/day are considered. Arrows represent normalized fluid velocity vector.
4964 J. Kim et al. / Energy 35 (2010) 4955e4965

rate is higher than 100 m3/day, the energy demand for the factory
(w70 kW) can be satisfied in this site.
As shown in Figs. 14 and 15, the system performance of the ATES
system dose not reach the maximum state during the early years of
operation time, because thermal energy is not fully accumulated in
the aquifer. Therefore, increasing the pumping/injection rate
during the early stage of the operation is recommended to enhance
the system performance.

5. Conclusions

The thermal interference of an ATES system that affects the


system performance primarily depends on the distance between
two boreholes, the hydraulic conductivity of an aquifer, and the
production/injection rate. In this study, three-dimensional thermo-
hydraulic modeling was conducted to examine the degree of the
Fig. 15. Thermal power as a function of elapsed time with the summer and the winter thermal interference by three major parameters and to estimate the
(Anseong site). system performance changes due to the thermal interference. The
modeling results indicate that the thermal interference grows as
shallow groundwater. The measured average pH, electrical the distance between two boreholes decreases, as the hydraulic
conductivity (EC) and dissolved oxygen (DO) of the groundwater in conductivity increases, and as the pumping/injection rate increases.
the study area are w7, w215 mS/cm, and w4.2 mg/L, respectively. In the actual ATES system design, one may not have a control on the
An ATES system is installed in a fertilizer factory, and the total area distance between boreholes and the hydraulic conductivity at all
for cooling and heating is about 500 m2. Thermal energy demand of times due to location conditions, but the pumping/injection rates.
cooling and heating for the factory is estimated to be w70 kW. The Therefore, before installation of an ATES system, the optimal
study site is equipped with the temperature monitoring system to pumping/injection rate could be determined by thermo-hydraulic
observe temperature variations in the aquifer during the operation modeling, which can give the initial cost reduction and the system
of the ATES system. Five observation wells were drilled between the performance maximization.
injection and the pumping wells, and twenty T-type thermocouples Comparative analysis of the system performance was performed
were installed in observation wells. The water quality analysis is in terms of the thermal interference. If the value of h is equal to
being conducted once a month to detect a potential scaling unity, the system performance can decrease up to w22%. If the
problem due to chemical changes of water by extracting and value of h, however, is smaller than unity, the system performance
injecting water. is not significantly affected. Therefore, the optimal distance
The following hydraulic and thermal properties of the study site between two boreholes and the pumping/injection rate should be
were measured or estimated. The hydraulic conductivity estimated determined to maintain the value of h lower than unity for the high
by a pumping test is w2  104 m/s, and the average porosity system performance of the ATES system.
estimated by the electrical conductivity logging is w0.3. The Long term three-dimensional thermo-hydraulic modeling for
average thermal conductivity of the aquifer measured by a needle a fertilizer factory, located in Anseong city, was conducted to test
probe is w1.6 W/m K, and the underground temperature measured the applicability of the ATES system. When the pumping/injection
by a temperature logging is w14e15  C. rate is 100 m3/day, system performances of the ATES system after 3
To predict the system performance of the ATES system installed years of operation are estimated to be w125 kW in the summer and
in the study site, long term three-dimensional thermo-hydraulic w110 kW in the winter, respectively. Therefore, in the study area, if
modeling was conducted. The boundary and initial conditions, and the pumping/injection rate is higher than 100 m3/day, the thermal
assumptions for the modeling are the same as the conceptual energy demand (w70 kW) for the factory seems to be satisfied.
modeling mentioned earlier. Physical and thermal properties of
fluid and aquifer for modeling are listed in Table 6. The distance Acknowledgement
between two boreholes and the hydraulic conductivity are deter-
mined to be 80 m and 2  104 m/s, respectively, and the pumping/ This research was supported by Korea Energy Management
injection rate is assumed to be 100 m3/day. The depth of pumping/ Corporation through grant number NP2007-037 and supported by
injection wells is 50 m, and warm water and cold water are injected Basic Research Program(GP2009-016) of Korea Institute of Geo-
and/or extracted at the bottom of wells. Fig. 14 shows temperature science and Mineral Resources (KIGAM). Chemical analysis of
distributions in the aquifer during 15 years of continuous operation. groundwater was conducted by Dr. Byong Wook Cho of KIGAM. Dr.
Warm thermal energy is stored through the right well during the Seho Hwang of KIGAM conducted geophysical loggings. We thank
summer period, and cold thermal energy, on the contrary, is stored Mr. Sun Joon Ahn, and Mr. Wan Soo Kim of nexGeo INC. for offering
through the left well during the winter period. The value of h after hydraulic properties.
15 years of continuous operation is estimated to be w0.4. There-
fore, the system performance would not be significantly affected by References
the thermal interference. Fig. 15 presents thermal power as
a function of time for the summer and the winter. For the first 3 [1] Meyer CF, Todd DK. Heat storage wells. Water Well J 1973;27(10):35e41.
[2] Molz FJ, Warman JC, Jones TE. Aquifer storage of heated water: Part 1 e A field
years, the system performance considerably increases, and after 3
experiment. Ground Water 1978;16(4):234e41.
years, the system performance reaches the maximum state. It [3] Papadopulos SS, Larson SP. Aquifer storage of heated water: Part 2 e
implies that thermal energy is almost completely stored in the Numerical simulation of field results. Ground Water 1978;16(4):242e8.
aquifer after the first 3 years. At this period, system performances in [4] Parr DA, Molz FJ, Melville JG. Field determination of aquifer thermal energy
storage parameters. Ground Water 1983;21(1):22e35.
the summer and the winter are estimated to be w125 kW, and [5] Andersson O, Hellstrom G, Nordell B. Heating and cooling with UTES in
w110 kW, respectively. Therefore, when the pumping/injection Sweden-current situation and potential market development. In: Proceedings
J. Kim et al. / Energy 35 (2010) 4955e4965 4965

of the 9th international conference on thermal energy storage, Warsaw, [19] Uffink GJM. Dampening of fluctuations in groundwater temperature by heat
Poland, vol. 1; 2003. p. 359e66. exchange between the aquifer and the adjacent layers. J Hydrol 1983;60
[6] Sanner B, Karytsas C, Mendrinos D, Rybach L. Current status of ground source (1e4):311e28.
heat pumps and underground thermal energy storage in Europe. Geothermics [20] Voigt HD, Haefner F. Heat transfer in aquifers with finite caprock thickness
2003;32(4e6):579e88. during a thermal injection process. Water Resour Res 1987;23(12):
[7] Paksoy HO, Andersson O, Abaci S, Evliya H, Turgut B. Heating and cooling of 2286e92.
a hospital using solar energy coupled with seasonal thermal energy storage in [21] Krarti M, Claridge DE. Two-dimensional heat transfer from earth-sheltered
an aquifer. Renew Energy 2000;19(1e2):117e22. buildings. J Sol Energy Eng 1990;112(1):43e50.
[8] Dickinson JS, Buik N, Matthews MC, Snijders A. Aquifer thermal energy: [22] Yang SY, Yeh HD. Solution for flow rates across the wellbore in a two-zone
theoretical and operational analysis. Geotechnique 2009;59(3):249e60. confined aquifer. J Hydraul Eng 2002;128(2):175e83.
[9] Novo VA, Bayon RJ, Castro-Fresno D, Rodriguez-Hernandez R. Review of [23] Stopa J, Wojnarowski P. Analytical model of cold water front movement in
seasonal heat storage in large basins: water tanks and gravel-water pits. Appl a geothermal reservoir. Geothermics 2006;35(1):59e69.
Energy 2010;87(2):390e7. [24] Tsang CF. Aquifer simulation-in theory and in practice. In: Proceedings of the
[10] Preene M, Powrie W. Ground energy systems: delivering the potential. Energy international conference on subsurface heat storage in theory and practice,
2009;162(2):77e84. Stockholm, Sweden; 1983. p. 116e25.
[11] Umemiya H, Satoh Y. A cogeneration system for a heavy-snow fall zone based [25] Kazmann RG. Exotic uses of aquifers. J Irrig Drain 1971;97(3):515e22.
on aquifer thermal energy storage. Jpn Soc Mech Eng 1990;33(4):757e65. [26] Dwyer TE, Eckstein Y. Finite-element simulation of low-temperature, heat--
[12] Gao Q, Li M, Yu M, Spitler JD, Yan YY. Review of development from GSHP to pump-coupled, aquifer thermal energy storage. J Hydrol 1987;95(1e2):
UTES in China and other countries. Renew Sustain Energy Rev 2009;13 19e38.
(6e7):1383e94. [27] Rouve G, Daniels H, Pelka W, Hahne E, Fisch N, Giebe R. Finite element
[13] Lee KS. Performance of open borehole thermal energy storage system under simulation of the hydrothermal behavior of an artificial aquifer for seasonal
cyclic flow regime. Geosci J 2008;12(2):169e75. energy storage. Adv Water Resour 1988;11(3):127e32.
[14] Fan R, Jiang Y, Yao Y, Shiming D, Ma Z. A study on the performance of [28] Tenma N, Yasukawa K, Zyvoloski G. Model study of the thermal storage
a geothermal heat exchanger under coupled heat conduction and ground- system by FEHM code. Geothermics 2003;32(4e6):603e7.
water advection. Energy 2007;32(11):2199e209. [29] Yotov I. Assessment of natural resources of thermal aquifers-formulation, test
[15] Rosen MA. Second-law analysis of aquifer thermal energy storage systems. data, results. Geol Balcanica 2008;37(1e2):79e84.
Energy 1999;24(2):167e82. [30] Hall SH, Raymond JR. Geohydrologic characterization for aquifer thermal
[16] Molson JW, Frind EO, Palmer CD. Thermal energy storage in an unconfined energy storage. In: 27th intersociety energy conversion engineer conference,
aquifer 2. Model development, validation, and application. Water Resour Res San Diego, CA, USA, vol. 4; 1992. p. 101e7.
1992;28(10):2857e67. [31] Bredehoeft JD, Papadopulos IS. Rates of vertical groundwater movement
[17] Kangas MT, Lund PD. Modeling and simulation of aquifer storage energy estimated from the earth’s thermal profile. Water Resour Res 1965;1(2):
systems. Sol Energy 1994;53(3):237e47. 325e8.
[18] Sauty JP, Gringarten AC, Menjoz A, Landel PA. Sensible energy storage in [32] Koo MH, Song Y, Lee JH. Analyzing spatial and temporal variation of ground
aquifers 1. Theoretical study. Water Resour Res 1982;18(2):245e52. surface temperature in Korea. Econ Environ Geol 2006;39(3):255e68.

You might also like