You are on page 1of 41

Accepted Manuscript

Ground heat storage beneath salt-gradient solar ponds under constant heat
demand

José Amigo, Francisco Suárez

PII: S0360-5442(17)32104-7

DOI: 10.1016/j.energy.2017.12.066

Reference: EGY 12015

To appear in: Energy

Received Date: 05 October 2017

Revised Date: 12 December 2017

Accepted Date: 13 December 2017

Please cite this article as: José Amigo, Francisco Suárez, Ground heat storage beneath salt-
gradient solar ponds under constant heat demand, Energy (2017), doi: 10.1016/j.energy.
2017.12.066

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to
our customers we are providing this early version of the manuscript. The manuscript will undergo
copyediting, typesetting, and review of the resulting proof before it is published in its final form.
Please note that during the production process errors may be discovered which could affect the
content, and all legal disclaimers that apply to the journal pertain.
ACCEPTED MANUSCRIPT

HIGHLIGHTS

 Representation of the thermal dynamics in a solar pond and the ground beneath it.

 Algorithm for removing heat at a constant rate from a solar pond is proposed.

 Water dependent soil thermal properties are defined.

 Temperatures in a solar pond decrease exponentially as the water table is shallower.

 Insulating solar ponds exacerbate temperatures oscillations at the pond’s bottom.


ACCEPTED MANUSCRIPT

1 Ground heat storage beneath salt-gradient solar ponds under constant heat

2 demand

3 José Amigo 1,2,3, Francisco Suárez 1,2,3,4*

6 1 Departamento de Ingeniería Hidráulica y Ambiental, Pontificia Universidad Católica de Chile.

7 Avda. Vicuña Mackenna 4860, Macul, Santiago, Chile.

8 2 Centro de Desarrollo Urbano Sustentable (CEDEUS), CONICYT/FONDAP/15110020, Santiago,

9 RM, Chile.

10 3 Center for Solar Energy Technologies (CSET), Santiago, RM, Chile.

11 4 Centro de Excelencia en Geotermia de los Andes (CEGA), Santiago, RM, Chile.

12

13

14

15

16

17 * Correspondence to: Francisco Suárez, Pontificia Universidad Católica de Chile. Avda. Vicuña

18 Mackenna 4860, Departamento de Ingeniería Hidráulica y Ambiental, Macul, Santiago, Chile.

19 Ph: (+56) 2 2354 5875; Fax: (+56) 2 2354 5876

20

1
ACCEPTED MANUSCRIPT

21 Abstract

22 Salt-gradient solar ponds are energy collectors and storage systems that provide continuous

23 heat supply. Although many studies have investigated the thermal behavior of solar ponds,

24 few researches have investigated how heat lost to the ground beneath a pond can be

25 recovered. Here, a one-dimensional transient model is used to study the thermal interaction

26 between a solar pond with constant heat demand and the ground beneath it. The ground

27 thermal properties were dependent on temperature and moisture. As groundwater depth

28 affects soil moisture distribution, higher thermal conductivities are observed when the

29 groundwater table is shallow. Further, the mean temperatures at the bottom of the pond

30 decrease exponentially as the groundwater depth is shallower. For deep groundwater tables,

31 marginal variations in the groundwater table depth do not have a considerable impact on the

32 pond’s bottom temperatures. The addition of an insulation layer is only beneficial when the

33 water table is shallow. When the water table is deep instead, the ground below the pond acts

34 as an additional heat storage volume, permitting more stable temperatures in the pond

35 throughout the year, making it more suitable for a constant heat demand.

36 Keywords: Solar pond, heat extraction, solar energy, ground heat storage.
37

2
ACCEPTED MANUSCRIPT

38 Nomenclature

ΔE energy change [W] k thermal conductivity [W/m-°C]


CP specific heat capacity [J/kg-°C] Qout outgoing energy flux [W]
Qin incoming energy flux [W] Δz layer thickness [m]
convective heat coefficient between the bottom
A area [m2] h1 of the solar pond and the ground beneath
[W/m2-°C]
convective heat coefficient between the ground
h2 T temperature [°C]
and the groundwater sink [W/m2-°C]
PWP pore water pressure [m] Mw molar mass of water [kg/mol]
convective heat flux from the LCZ to the
QLCZ/s Qs conduction heat flux in the soil [W/m2]
ground beneath [W/m2]
convective heat flux from the ground to the
Qw lateral heat conduction flux [W/m2] Qs/WT
groundwater table [W/m2]
extracted heat with the external heat exchanger extracted heat with the internal heat exchanger
QEHE QIHE
[W/m2] [W/m2]
ṁ mass flow rate [kg/s] U overall heat transfer coefficient [W/m2-°C]
convective heat coefficient of the heat transfer convective heat coefficient of the heat transfer
he hi
fluid outside the external heat exchanger fluid outside the external heat exchanger
external diameter of the internal heat exchanger internal diameter of the internal heat exchanger
de di
pipe [mm] pipe [mm]
QOP heat extraction rate [W/m2] n water retention curve parameter [-]
QD constant heat demand [W/m2] Dv vapor diffusivity [m2/s]
hr relative humidity in the soil [-] ea vapor pressure [Pa]
fw dimensionless flow factor of Campbell [-] ga shape factor of soil particles [-]
es saturated vapor pressure [Pa] R gas constant [J/mol-°C]
parameter of the friction flow factor of
g gravity acceleration [m/s2] q
Campbell [-]
gc shape factor of soil particles [-] Patm atmospheric pressure [Pa]

Greek Symbols

ρ specific mass [kg/m3] ψ water potential [J/kg]


θr residual water content [cm3/cm3] θs residual water content [cm3/cm3]
α water retention curve parameter [1/cm3] Δ slope of the vapor pressure function [Pa/°C]
λ latent heat of vaporization [J/mol] Φ volume fraction [cm3/cm3]
ξ weighting function [-] ρM molar density of air [mol/m3]
water content in which the return flow cuts off convergence criterion parameter for constant
θ0 ε
[cm3/cm3] heat extraction [W/m2]

Subscripts/superscripts

avg yearly average of the air s,N deepest soil layer


WT water table s soil
c concrete SGSP salt-gradient solar pond
LCZ lower convective zone IHE internal heat exchanger
pe polyethylene m mineral
n nth layer in the domain s,1 shallowest soil layer
i ith layer in the domain w wall
ps polystyrene x different components of the soil
w water g gas fraction
b bulk a air
f fluid of the soil

39

40

41

3
ACCEPTED MANUSCRIPT

42 1 Introduction

43 Salt-gradient solar ponds (SGSPs) are low-cost solar collectors, with long-term heat storage

44 capacity that can deliver heat continuously through time [1]. SGSPs have become an

45 attractive technology in locations that have high solar irradiance, excess water and excess

46 salts. An SGSP generally has three characteristic zones (Fig. 1). The upper convective zone

47 (UCZ), located at the top of the pond, is a thin and uniform layer of fresh water or low salt

48 concentration (0–4 % by weight) with a typical thickness ranging between 0.1 and 0.4 m [2].

49 The intermediate zone, called non-convective zone (NCZ), is formed by a constant salinity

50 gradient that increases salt concentration within the pond as depth increases. Its thickness

51 varies depending on the desired temperature at the bottom of the pond and the heat extraction

52 rate [3]. The bottom zone, called storage zone or lower convective zone (LCZ), is comprised

53 by a high salinity solution (21–26% by weight) and its thickness typically varies between 0.8

54 and 1.2 m [4]. As the salts in the solar pond are completely dissolved, the water is transparent

55 and allows radiation to penetrate into the deepest layer of the pond. The salinity gradient acts

56 as a transparent insulator for the LCZ, allowing the sunlight to travel until the pond’s bottom,

57 but suppressing the natural convection within the pond because the salt gradient counteracts

58 the buoyancy effect of the warmer water below [5–7]. Thus, the only heat losses from the

59 LCZ to the atmosphere occur by conduction, but the relatively low thermal conductivity and

60 high heat capacity of the brine allows collecting heat at the bottom of the pond, which makes

61 the SGSP a long-term heat storage device [8]. One of the most promising application of

62 SGSPs is coupling them with thermal desalination technologies [9–14].

63 The highest temperatures inside the SGSP are achieved in the LCZ, with typical temperatures

64 ranging between 70 and 90 °C [15], which imply that an important part of the stored heat

65 may be lost to the ground [16]. Several investigations have studied ways to improve SGSP

4
ACCEPTED MANUSCRIPT

66 performance [17–25], but few have addressed thermal storage in the ground beneath an SGSP

67 [1,16,26–30]. Sezai and Tasdemiroglu [26] studied the effect of the bottom reflectivity on the

68 ground heat losses. Zhang et al. [27] theoretically studied the heat losses from an SGSP to the

69 ground under different combinations of depth to groundwater, soil thermal properties and

70 heat withdrawal patterns. Among all of these studies, none of them considered the effect of

71 the soil water content on the ground thermal properties, which have been typically assumed

72 to be uniform. This assumption can be unrealistic for shallow groundwater tables, as the soil

73 water content changes more abruptly. According to the de Vries expressions [31], when the

74 volume fraction of water in clay, sand and loam increases from 0.0 to 0.3, the thermal

75 conductivity increases in approximately 100% for clay and loam, and in ~220% for sand,

76 when temperature is ~20 °C.

77 The objective of this work is to study the heat storage capacity of the ground beneath an

78 SGSP and its capacity of improving the operation of this technology when a constant heat is

79 demanded. To achieve this objective, a one-dimensional transient model was used to describe

80 the thermal behavior of an SGPS and the ground beneath it. As a hypothetical study case, the

81 thermal evolution of an SGSP located in Copiapó, Chile, was modeled. Furthermore, an

82 algorithm was developed to satisfy a constant heat demand with heat being extracted from the

83 LCZ. Under this operating condition, we analyzed the main variables, related to the ground,

84 which influence the LCZ temperatures: the soil type and the depth to groundwater. Finally,

85 the impact of an insulation layer at the bottom of the SGSP was evaluated.

5
ACCEPTED MANUSCRIPT

86 2 Materials and methods

87 2.1 Thermal model

88 The temperature evolution in an SGSP and the ground beneath it can be represented by a one-

89 dimensional transient model using the finite difference method [16,32–35]. The developed

90 model uses an energy balance in each node within the domain to describe the thermal

91 dynamics (Fig. 2), as a function of the incoming and outgoing energy fluxes [16,32]:

92 ∆𝐸
𝑡 + ∆𝑇
𝑛 = 𝜌𝑛𝑐𝑃,𝑛∆𝑧𝑛𝐴𝑆𝐺𝑆𝑃 = ∑𝑄𝑖𝑛,𝑛 ‒ ∑𝑄𝑜𝑢𝑡,𝑛 (1)

93 The model assumes that both the UCZ and LCZ are completely mixed, i.e., their temperatures

94 are uniform. The shortwave radiation attenuation through the water column is assumed to be

95 exponential, as proposed by Rabl and Nielsen [36]. Heat loss from the SGSP to the ground,

96 QLCZ/s, is treated as a convective loss according to:

97 𝑄𝐿𝐶𝑍/𝑠 = ℎ1(𝑇𝑠,1 ‒ 𝑇𝐿𝐶𝑍) (2)

98 The convective heat transfer coefficient between the LCZ and the bottom of the SGSP, h1, is

99 considered constant and equal to 78.12 W/m2-°C [37]. Furthermore, heat transport to deeper

100 ground layers occurs only by conduction and is represented using Fourier’s Law:

(
𝑇𝑠,𝑖 ‒ 1 ‒ 𝑇𝑠,𝑖
101 𝑄𝑠,𝑖 = 𝑘𝑠,𝑖
∆𝑧𝑠,𝑖 ) (3)

102 The development of the model, its calibration and validation were performed using an

103 experimental SGSP that was built on Santiago, Chile. The reader is referred to the work of

104 Amigo et al. [16] for further details about the model development and validation. Despite

105 more complex 2D models have been developed to solve the dynamics in SGSPs [38], the

6
ACCEPTED MANUSCRIPT

106 benefits for our study case are not considerable as the discussion is focused in 1D heat fluxes

107 between the SGSP and the ground beneath the pond.

108 2.1.1 Boundary conditions

109 The model’s top boundary condition is given by known heat fluxes from latent and sensible

110 heat and longwave radiation, calculated based on meteorological data [39–41]. Shortwave

111 radiation is treated separately from the heat fluxes across the water surface because it causes

112 internal heating of the fluid [38].

113 The lateral boundary condition is given by a known convective heat flux (Eq. 4) [42]. At a

114 certain lateral distance xs (Fig. 2), heat fluxes coming from the pond can be considered

115 negligible and the temperature in the ground is assumed to be equal to the yearly average of

116 the ambient (Tavg) [32]. An overall convective heat transfer coefficient, Uw, is used to

117 represent the resistance of the ground to heat conduction.

𝑥𝑠
118 𝑄𝑤(𝑧,𝑡) = 𝑈𝑤(𝑇𝑎𝑣𝑔 ‒ 𝑇(𝑧,𝑡)) = (𝑇𝑎𝑣𝑔 ‒ 𝑇(𝑧,𝑡)) (4)
𝑘 𝑠

119 Lastly, the lower boundary condition is given by a convective heat flux between the deepest

120 ground layer and the groundwater table, calculated based on a known water table

121 temperature, TWT (Eq. 5). The convective heat transfer coefficient between the ground and the

122 groundwater sink, h2, is assumed to be constant and equal to 185.8 W/m2/°C [37].

123 𝑄𝑠/𝑊𝑇 = ℎ2(𝑇𝑊𝑇 ‒ 𝑇𝑠,𝑁) (5)

124 2.1.2 Soil water content and thermal properties

125 In this work, the water in the soil is assumed to be in static equilibrium and the pore water

126 pressure at a certain depth is given by:

7
ACCEPTED MANUSCRIPT

127 𝑃𝑊𝑃𝑖 = (𝑧𝑖 ‒ 𝑧𝑊𝑇) (6)

128 The soil water content, θ, is described using the van Genuchten equation [43] as a function of

129 the pore water pressure and of four independent parameters: the residual water content, θr, the

130 saturated water content, θs, and two empirical parameters, α and n:

{
𝜃𝑠 ‒ 𝜃
𝑟
𝜃𝑟 + ,
1
𝑃𝑊𝑃 < 0
131 𝜃(𝑃𝑊𝑃) = 1‒ (7)
[1 + |𝛼𝑃𝑊𝑃|𝑛] 𝑛

𝜃𝑠 , 𝑃𝑊𝑃 ≥ 0

132 The soil volumetric heat capacity (ρsCP,s) and thermal conductivity (ks) are defined using the

133 de Vries expressions [31], as a weighted sum of its components:

𝑛
134 𝜌𝑠𝐶𝑃,𝑠 = ∑𝑥 = 1∅𝑥𝜌𝑥𝐶𝑃,𝑥 (8)

𝑛 ∅𝑥𝜀𝑥𝑘𝑥
135 𝑘𝑠 = ∑𝑥 = 1 ∅ 𝜀 (9)
𝑥 𝑥

136 where the subscript x represent different materials in the soil. Details of the soil thermal

137 properties calculation can be found in Appendix A.

138 2.2 Simulation conditions

139 To understand the role of the soil heat storage on the operation of a solar pond, the thermal

140 evolution of a 20,000-m2 SGSP located in Copiapó, Chile (27°57’32”S, 70°0’36”W) was

141 simulated for a period of 5 years. After the maturation period (six months), heat was removed

142 from the LCZ at a constant rate. It was assumed that algae and turbidity in the SGSP are

143 controlled with hydrochloric acid [44–46] and therefore heat storage in the LCZ is not

144 affected by water quality.

8
ACCEPTED MANUSCRIPT

145 The hourly meteorological conditions of Copiapó were used to estimate the boundary

146 conditions at the top of the SGSP [47]. As a reference, the monthly average climate

147 conditions for a representative year in Copiapó are shown in Table 1. The SGSP was assumed

148 to have 0.10-m concrete walls (kc = 1 W/m-°C, ρc = 880 kg/m3, CP,c = 2,230 J/kg-°C). Three

149 different soil types in which the pond is buried were evaluated: clay, loam and sand, with the

150 water table 35 m below the bottom of the SGSP [48]. Parameters related to the water

151 retention curve (Eq. 7) and thermal properties for each soil type (Eq. 8 and 9) were obtained

152 from [49] and [50], respectively. The values for the different parameters are presented in

153 Table A.1 and Table A.2 (Appendix A). The temperature of the groundwater table was used

154 as the lower boundary condition, which is assumed to be constant and equal to 24.3 °C [51].

155 The UCZ thickness was set at 0.3 m and the LCZ at 1.0 m [52]. The NCZ thickness was 1.2

156 m, and was defined to maximize the LCZ temperatures [16].

157 2.3 Heat extraction

158 Heat was withdrawn from the LCZ using an internal heat exchanger (IHE). A heat transfer

159 fluid is circulated in a closed cycle through the IHE and the transferred thermal energy is then

160 removed using an external heat exchanger (EHE) [53], as shown in Fig. 1. The thermal

161 energy extracted from the EHE was calculated with the following equations [54]:

𝑚𝐶𝑃(𝑇𝑜𝑢𝑡 ‒ 𝑇𝑖𝑛)
162 𝑄𝐸𝐻𝐸 = (10)
𝐴

𝑇𝐿𝐶𝑍 ‒ 𝑇𝑖𝑛
163 𝑇𝑜𝑢𝑡 = 𝑇𝐿𝐶𝑍 ‒ (11)

( )
𝑈𝐼𝐻𝐸𝐴𝐼𝐻𝐸
exp
𝑚𝐶𝑃

9
ACCEPTED MANUSCRIPT

1
164 𝑈𝐼𝐻𝐸 = (12)
𝑑𝑒 1 𝑑𝑒
()
1𝑑𝑒
+ ln +
𝑑𝑖 ℎ𝑖 2𝑘𝑝𝑒 𝑑𝑖 ℎ𝑒

165 The overall heat transfer coefficient of the IHE, UIHE, was calculated with Eq. 12.The IHE

166 design is similar to the one used at the Pyramid Hill’s solar pond [53], with forty 200-m long

167 heat extraction polyethylene tubes (kpe = 0.37 W/m-°C), all connected to two manifold pipes.

168 The convective coefficients of the heat transfer fluid inside, hi, and outside the IHE, he, are

169 5,000 and 1,500 W/m2-°C, respectively.

170 Due to the intermittent nature of the solar radiation, one of the biggest challenges of solar

171 technologies is to continuously supply energy for long-time periods, which implies that heat

172 must be extracted during the night. Because of this issue, as an operation restriction, a

173 constant heat demand, QD, will be satisfied. Further, when temperatures in the LCZ exceed

174 95°C, an additional amount of heat will be removed from the pond in order to avoid the

175 erosion of the salt gradient due to boiling of the LCZ brine. Additionally, a minimum

176 temperature difference of 20°C between the heat transfer fluid and the LCZ temperature will

177 be required for an effective heat removal [32]. To achieve this constant heat extraction rate,

178 the mass flow rate through the heat exchangers must vary in time as a function of the LCZ

179 temperature. A seven-step iteration method is used to determine the mass flow rate for each

180 time step. This method is illustrated in Fig. 3. For a given time and desired heat extraction

181 rate, QOP, if the temperature in the LCZ, TLCZ, is higher than the maximum permitted, Tmax (in

182 our case 95°C), after QOP is removed, an additional amount of heat will be extracted until the

183 temperatures are below Tmax (Step 1-2-3-1). If TLCZ is lower than Tmax instead the model

184 settles an arbitrary value for the mass flow rate to initialize the iteration (Step 4). Using Eq.

185 11, the fluid’s temperature at the outlet of the IHE is calculated (Step 5). Then, Eq. 10 is used

186 to calculate the extracted heat rate at the EHE, QEHE (Step 6). If the difference between

10
ACCEPTED MANUSCRIPT

187 QEHE and QOP is larger than a convergence criterion,  (in our case fixed at 0.01 W/m2), the

188 model recalculates the mass flow rate as a function of QOP and Tout, by reassembling Eq. 10

189 (Step 7), and goes back to Step 5. This loop is maintained until QEHE converges to QOP and

190 the mass flow rate for a certain time step will be the value that meets the convergence

191 criterion. In practice, this algorithm has to be implemented by varying the pump power

192 accordingly to the exchange fluid and the LCZ temperature, to maintain the desired extracted

193 heat. The main assumptions of the model, with their respective impact, are summarized in

194 Table 2.

195 3 Results and discussion

196 3.1 SGSP with no insulation layer

197 The thermal evolution of the LCZ for the different soil types, when no insulation is used and

198 when QD is equal to 50 W/m2, is illustrated in Fig. 4(a). Results show that as the soil texture

199 gets finer, the LCZ temperature oscillations are higher. When quasi-steady state is reached,

200 the annual average LCZ temperatures are 81.0, 80.9 and 80.2 °C for clay, loam and sand,

201 respectively. As the soil texture gets coarser, the soil thermal conductivity increases and the

202 heat losses from the SGSP to the ground increase. According to Fig. 4(b), mean annual heat

203 losses are 45% higher when the soil is sand instead of clay. Further, soils with higher thermal

204 conductivity have more energy recovery during the winter period. For instance, for sand,

205 23.6% of the heat that is lost to the soil during a year is recovered in the winter period (once

206 the quasi-steady state is reached), with fluxes of up to 10.5 W/m2. As more heat is removed in

207 the winter period, when the SGSP is coldest, the ground thermal storage will be more

208 efficient because more heat will be recovered [27].

11
ACCEPTED MANUSCRIPT

209 When the soil thermal properties are defined as a function of the water content, the

210 groundwater table depth (also known as water table depth, WTD) influences the distribution

211 of moisture in the soil profile. Fig. 5 shows the soil moisture profile when the WTDs are 7

212 and 20 m with their resultant thermal conductivity and temperature envelopes. As the WTD is

213 shallower, the soil beneath the SGSP has a higher water content, which will increase the soil

214 effective thermal conductivity. Therefore, heat losses to the ground will be higher. As a

215 comparison, the model was also run with constant soil thermal properties. Fig. 6(a) shows the

216 heat fluxes between the LCZ and the soil beneath, which for this example case is loam. The

217 black line represents the fluxes when the soil thermal properties are water dependent (Eqs. 8

218 and 9). The blue, red and purple lines instead, represent the fluxes for moist ( = s), dry ( =

219 r) and standard loam respectively, where constant thermal properties are used (Table A.2,

220 Appendix A). Standard loam thermal properties are defined as the average of the dry and

221 moist loam thermal properties. Results show that the heat fluxes oscillation at the bottom of

222 the SGSP is higher for moist soils, which is expected due to their higher thermal

223 conductivity. It can be observed that the loam with standard thermal properties can be a good

224 representation of a soil with water dependent thermal properties, for the specific case when

225 the WTD is 35 m.

226 To understand the impact of the WTD on the SGSP temperature and the soil thermal

227 properties, a sensitivity analysis was made. Fig. 6(b) shows the variation of the LCZ mean

228 temperatures as the WTD changes. The origin represents the base scenario for a loam soil

229 when the water table depth is 35 m and the soil thermal properties are water dependent. As

230 the WTD increases, the LCZ temperature is mildly affected: when the WTD increases in

231 +50%, i.e., WTD = 52.5 m, the temperature increases only in a 0.2%. A similar conclusion

232 was made by [55], where it was demonstrated that below a certain WTD, further depression

12
ACCEPTED MANUSCRIPT

233 does not have a significant impact on the thermal performance of the SGSP. Instead, when

234 the WTD is shallower, the temperature in the LCZ decreases more abruptly.

235 According to [27], the heat storage in the soil beneath a SGSP can stabilize the temperatures

236 in these systems. Nevertheless, the soil will act as an additional volume of heat storage as

237 long as a significant part of the heat that is lost to the ground can be recovered during the

238 winter period. Fig. 7 shows the heat fluxes from the LCZ to the soil beneath for different

239 WTDs and soil types. For each soil type, there is a minimum WTD for heat to be recovered

240 during the winter, under the specified operation condition. For instance, for sand with WTDs

241 shallower than 7 m, the heat will continuously be lost to the ground. According to [28], when

242 the soil beneath an SGSP has high moisture or a shallow water table, an insulation layer may

243 be needed.

244 3.2 SGSP with insulation layer

245 To understand the impact of an insulation layer in the thermal dynamics of an SGSP, we

246 compared the thermal behavior between an SGSP with and without a 20-cm polystyrene layer

247 (kps = 0.03 W/m-°C, ρps = 35 kg/m3, CP,ps = 1400 J/kg-°C) at its bottom. For this comparison,

248 the soil surrounding the SGSP is loam and different scenarios for heat demand and WTD

249 were considered. Figs. 8(a)-(f) shows the temperature evolution in the LCZ and in the

250 ground, 1 meter below the SGSP bottom, for the different scenarios. When the temperatures

251 in the ground are higher than those at the LCZ, heat is being recovered from the ground to the

252 SGSP. Results show that when the WTD is 35 m (Figs. 8(a)-(c)), the different scenarios for

253 heat demand can be satisfied with or without the insulation layer. Moreover, when QD is 30

254 and 50 W/m2, the average temperature in the LCZ once the steady-state is reached is higher

255 for the cases with no insulation layer. As QD increases, more oscillations in the LCZ

256 temperatures are observed and are higher in the cases with insulation layer. When QD is 50

13
ACCEPTED MANUSCRIPT

257 W/m2, the lowest temperature in the LCZ when an insulation layer is used is 10% lower than

258 that observed for the case without an insulation layer. This can be explained because when an

259 insulation layer is added, the heat storage volume is reduced to a smaller volume –only the

260 LCZ. Therefore, the thermal evolution of the LCZ is more sensible to atmospheric changes,

261 which implies more variability in temperatures throughout the year. The scenario with no

262 insulation layer instead, uses the soil as an additional volume of heat storage, maintaining

263 more stable LCZ temperatures throughout the year, making the heat extraction in the winter

264 period more suitable. A better perspective of the impact of the insulation layer on the

265 temperature field is seen in Fig. 9, where the thermal evolution in the complete domain,

266 SGSP and ground, is shown.

267 When the WTD is 4 m instead, heat losses to the ground are larger and the soil is no longer a

268 suitable volume for heat storage (Figs. 8(d)-(f)). For instance, when QD is 30 W/m2, heat

269 losses to the ground are 352% higher than those observed when the WTD is 35 m. Despite

270 the heat demand is satisfied for all the cases without an insulation layer, satisfying demands

271 higher than 50 W/m2 would not be possible, because the temperature difference between the

272 LCZ and the heat transfer fluid would be lower than 20°C.

273 4 Conclusions

274 Results showed that as the soil beneath the SGSP is coarser, more heat is lost to ground.

275 When the WTD is 35 m, heat losses to ground are 45% higher when the soil is sand instead of

276 clay. Further, the soil thermal conductivity is influenced by the WTD: the shallower the

277 WTD, the larger the soil thermal conductivity. For WTD deeper than 35 m, marginal

278 variations in the WTD do not have a considerable impact on the LCZ temperatures. For

279 instance, when the WTD increases in 50% (from 35 to 52.5 m), the soil thermal conductivity

280 changes in 0.2%. Further, heat storage in the ground beneath an SGSP can have a positive

14
ACCEPTED MANUSCRIPT

281 impact on the operation of an SGSP when a constant heat demand has to be satisfied. For

282 example, for the studied scenarios with a 35-m WTD and QD=50 W/m2, the LCZ

283 temperatures during the winter period were 10% higher when no insulation layer was used. If

284 the water table is deep enough, the soil beneath the SGSP acts as an additional volume for

285 heat storage, increasing the capacity of the SGSP to store heat during the summer period,

286 stabilizing the LCZ temperatures throughout the year and making heat extraction in the

287 winter periods more suitable.

288 Nevertheless, the soil may act as an additional volume of heat storage only if the WTD is

289 deep enough. When the SGSP was surrounded with sand, the minimum WTD to recover heat

290 from the soil was 7 m. When the WTD was 4 m instead, all heat is lost to ground and the

291 insulation layer was justified as the heat losses to ground increased abruptly (352% in

292 comparison to a 35-m WTD, for the case where QD=30 W/m2). Given the large and negative

293 impact that a shallow WTD has on the SGSP thermal performance, it is recommended to

294 carefully characterize the soil before constructing an SGSP.

295 Acknowledgements

296 The authors thank the Comisión Nacional de Investigación Científica y Tecnológica

297 (CONICYT), Chile, for funding project Fondecyt N°1170850, the Centro de Desarrollo

298 Urbano Sustentable (CEDEUS – CONICYT/FONDAP/15110020) , the Centro de Excelencia

299 en Geotermia de los Andes (CEGA – CONICYT/FONDAP/15090013), and the Center for

300 Solar Energy Technologies (CSET—CORFO 13CEI2-21803), which supported this

301 investigation.

302

303 Appendix A. Soil thermal properties

15
ACCEPTED MANUSCRIPT

304 A.1. Soil volumetric heat capacity

305 The soil is assumed to be composed by minerals and water, and calculated as:

306 𝜌𝑔𝐶𝑃,𝑔 = ∅𝑚𝜌 𝐶𝑃,𝑚 + 𝜃𝜌𝑤𝐶𝑃,𝑤 (𝐴.1)


𝑚

307 The water density, ρw, is calculated as a function of the temperature T as [32]:

308 𝜌𝑤 = 998 ‒ 0.4 ∗ (𝑇 ‒ 20) (𝐴.2)

309 The volumetric water fraction, θ, is given by the soil water retention curve and the mineral

310 fraction, m, is calculated as it follows:

𝜌𝑏
311 ∅𝑚 = (𝐴.3)
𝜌𝑚

312 A.2. Soil thermal conductivity

313 The soil thermal conductivity was calculated as a weighted sum of the thermal conductivities

314 of its components.

𝜃𝜉𝑤𝑘𝑤 + ∅𝑔𝜉𝑔𝑘𝑔 + ∅𝑚𝜉𝑚𝑘𝑚


315 𝑘𝑠 = (𝐴.4)
∅𝑤𝜉𝑤 + ∅𝑔𝜉𝑔 + ∅𝑚𝜉𝑚

316 The volumetric fraction of the gaseous phase, g, is obtained from:

317 ∅𝑔 = 1 ‒ ∅𝑚 ‒ 𝜃 (𝐴.5)

318 The thermal conductivities of the gaseous and liquid (water) phases are calculated as follows:

319 𝑘𝑤 = 0.56 + 0.0018 ∗ 𝑇 (𝐴.6)

16
ACCEPTED MANUSCRIPT

𝜆𝛥ℎ𝑟𝑓𝑤𝜌𝑀𝐷𝑣
320 𝑘𝑔 = 𝑘𝑎 + (𝐴.7)
𝑃𝑎𝑡𝑚 ‒ 𝑒𝑎

321 where the latent heat of vaporization of water, λ, and the vapor diffusivity of the gases in air,

322 Dv, are temperature dependent [50]. The air thermal conductivity (ka), the dimensionless flow

323 factor of Campbell (fw), the vapor pressure (ea), the slope of the vapor pressure function (Δ)

324 and the relative humidity in the soil (hr) are calculated using Eqs. (A.8)-(A.13):

325 𝑘𝑎 = 0.024 + 0.00007 ∗ 𝑇 (𝐴.8)

1
326 𝑓𝑤 = (𝐴.9)
()
𝜃 ‒𝑞
1+
𝜃0

327 𝑒𝑎 = ℎ𝑟𝑒𝑠 (𝐴.10)

328 𝑒𝑠 = 2.718 ∗ 10
10
(
∗ exp ‒
4,157
𝑇 + 239.24 ) (𝐴.11)

(
exp 23.238 ‒
3,841
)
𝑇 + 228.15
329 ∆ = 3,841 ∗ (𝐴.12)
(𝑇 + 228.15)2

(
𝑀𝑤𝜓
330 ℎ𝑟 = exp
𝜌𝑤𝑅(𝑇 + 273.15) ) (𝐴.13)

331 The water potential, ψw, is calculated as follows:

332 𝜓𝑤 = 𝜌𝑤𝑔𝑃𝑊𝑃 (𝐴.14)

333 The weighting functions for gas (ξg), water (ξw) and minerals (ξm) are calculated using Eqs.

334 (A.15)-(A.17):

17
ACCEPTED MANUSCRIPT

2 1
335 𝜉𝑔 = + (𝐴.15)

( ( )) ( ( ))
3 1 + 𝑔𝑎
𝑘𝑔
𝑘𝑓
‒1 3 1 + 𝑔𝑐
𝑘𝑔
𝑘𝑓
‒1

2 1
336 𝜉𝑤 = + (𝐴.16)

( ( )) ( ( ))
3 1 + 𝑔𝑎
𝑘𝑤
𝑘𝑓
‒1 3 1 + 𝑔𝑐
𝑘𝑤
𝑘𝑓
‒1

2 1
337 𝜉𝑚 = + (𝐴.17)

( ( )) ( ( ))
3 1 + 𝑔𝑎
𝑘𝑚
𝑘𝑓
‒1 3 1 + 𝑔𝑐
𝑘𝑚
𝑘𝑓
‒1

338 𝑔𝑐 = 1 ‒ 2𝑔𝑎 (𝐴.18)

339 𝑘𝑓 = 𝑘𝑔 + 𝑓𝑤(𝑘𝑤 ‒ 𝑘𝑔) (𝐴.19)

340 All the parameters used in Appendix A are presented in Table A.1 with their respective

341 values.

342 References

343 [1] Prasad R, Rao D. Feasibility studies on the enhancement of energy storage in the

344 ground beneath solar ponds. Sol Energy 1993;50:135–44.

345 [2] El-Sebaii AA, Ramadan MRI, Aboul-Enein S, Khallaf AM. History of the solar ponds:

346 A review study. Renew Sustain Energy Rev 2011;15:3319–25.

347 doi:10.1016/j.rser.2011.04.008.

348 [3] JR H, Nielsen C, Golding P. Salinity-gradient solar ponds. Boca Raton, Florida: CRC

349 Press; 1989.

350 [4] Garrido F, Soto R, Vergara J, Walczak M, Kanehl P, Nel R, et al. Solar pond

18
ACCEPTED MANUSCRIPT

351 technology for large-scale heat processing in a Chilean mine. J Renew Sustain Energy

352 2012;4:053115. doi:10.1063/1.4757627.

353 [5] Saleh A, Qudeiri JA, Al-Nimr MA. Performance investigation of a salt gradient solar

354 pond coupled with desalination facility near the Dead Sea. Energy 2011;36:922–31.

355 doi:10.1016/j.energy.2010.12.018.

356 [6] Suárez F, Ruskowitz JA, Childress AE, Tyler SW. Understanding the expected

357 performance of large-scale solar ponds from laboratory-scale observations and

358 numerical modeling. Appl Energy 2014;117:1–10.

359 doi:http://dx.doi.org/10.1016/j.apenergy.2013.12.005.

360 [7] Abdel-Salam HEA, Probert SD. Solar ponds: Designs and prospects. Appl Energy

361 1986;24:91–126. doi:10.1016/0306-2619(86)90064-4.

362 [8] Lu H, Walton JC, Swift AHP. Desalination coupled with salinity-gradient solar ponds.

363 Desalination 2001;136:13–23. doi:10.1016/S0011-9164(01)00160-6.

364 [9] González D, Amigo J, Suárez F. Membrane distillation: Perspectives for sustainable

365 and improved desalination. Renew Sustain Energy Rev 2017;80:238–59.

366 doi:10.1016/j.rser.2017.05.078.

367 [10] Suárez F, Urtubia R. Tackling the water-energy nexus: an assessment of membrane

368 distillation driven by salt-gradient solar ponds. Clean Technol Environ Policy

369 2016;18:1697-1712.

370 [11] Suárez F, Ruskowitz JA, Tyler SW, Childress AE. Renewable water: Direct contact

371 membrane distillation coupled with solar ponds. Appl Energy 2015;158:532–9.

372 doi:http://dx.doi.org/10.1016/j.apenergy.2015.08.110.

19
ACCEPTED MANUSCRIPT

373 [12] Salata F, Coppi M. A first approach study on the desalination of sea water using heat

374 transformers powered by solar ponds. Appl Energy 2014;136:611–8.

375 doi:http://dx.doi.org/10.1016/j.apenergy.2014.09.079.

376 [13] Lu H, C. Walton J, H.P. Swift A. Desalination coupled with salinity-gradient solar

377 ponds. Desalination 2001;136:13–23. doi:http://dx.doi.org/10.1016/S0011-

378 9164(01)00160-6.

379 [14] Szacsvay T, Hofer-Noser P, Posnansky M. Technical and economic aspects of small-

380 scale solar-pond-powered seawater desalination systems. Desalination 1999;122:185–

381 93. doi:10.1016/S0011-9164(99)00040-5.

382 [15] Abdullah AA, Lindsay KA, AbdelGawad AF. Construction of sustainable heat

383 extraction system and a new scheme of temperature measurement in an experimental

384 solar pond for performance enhancement. Sol Energy 2016;130:10–24.

385 doi:10.1016/j.solener.2016.02.005.

386 [16] Amigo J, Meza F, Suárez F. A transient model for temperature prediction in a salt-

387 gradient solar pond and the ground beneath it. Energy 2017;132:257–68.

388 doi:10.1016/j.energy.2017.05.063.

389 [17] Suárez F, Aravena JE, Hausner MB, Childress AE, Tyler SW. Assessment of a vertical

390 high-resolution distributed-temperature-sensing system in a shallow thermohaline

391 environment. Hydrol Earth Syst Sci 2011;15:1081–93. doi:10.5194/hess-15-1081-

392 2011.

393 [18] Suárez F, Childress AE, Tyler SW. Temperature evolution of an experimental salt-

394 gradient solar pond. J Water Clim Chang 2010;1:246–50.

20
ACCEPTED MANUSCRIPT

395 [19] Silva C, Suárez F. An experimental and numerical study of evaporation reduction in a

396 salt-gradient solar pond using floating discs. Sol Energy 2017;142:204–14.

397 doi:http://dx.doi.org/10.1016/j.solener.2016.12.036.

398 [20] Ruskowitz JA, Suárez F, Tyler SW, Childress AE. Evaporation suppression and solar

399 energy collection in a salt-gradient solar pond. Sol Energy 2014;99:36–46.

400 doi:http://dx.doi.org/10.1016/j.solener.2013.10.035.

401 [21] Husain M, Sharma G, Samdarshi SK. Innovative design of non-convective zone of salt

402 gradient solar pond for optimum thermal performance and stability. Appl Energy

403 2012;93:357–63. doi:10.1016/j.apenergy.2011.12.042.

404 [22] Bezir NÇ, Dönmez O, Kayali R, Özek N. Numerical and experimental analysis of a

405 salt gradient solar pond performance with or without reflective covered surface. Appl

406 Energy 2008;85:1102–12. doi:10.1016/j.apenergy.2008.02.015.

407 [23] Sarathkumar P, Sivaram AR, Rajavel R, Praveen Kumar R, Krishnakumar SK.

408 Experimental Investigations on the Performance of A Solar Pond by using

409 Encapsulated Pcm with Nanoparticles. Mater. Today Proc. 2017;4:2314–22.

410 doi:10.1016/j.matpr.2017.02.080.

411 [24] Shi Y, Yin F, Shi L, Wence S, Li N, Liu H. Effects of porous media on thermal and

412 salt diffusion of solar pond. Appl Energy 2011;88:2445–53.

413 doi:10.1016/j.apenergy.2011.01.033.

414 [25] Alcaraz A, Valderrama C, Cortina JL, Akbarzadeh A, Farran A. Enhancing the

415 efficiency of solar pond heat extraction by using both lateral and bottom heat

416 exchangers. Sol Energy 2016;134:82–94. doi:10.1016/j.solener.2016.04.025.

21
ACCEPTED MANUSCRIPT

417 [26] Sezai I, Taşdemiroǧlu E. Effect of bottom reflectivity on ground heat losses for solar

418 ponds. Sol Energy 1995;55:311-319. doi:10.1016/0038-092X(95)00054-U.

419 [27] Zhang ZM, Wang YF. A study on the thermal storage of the ground beneath solar

420 ponds by computer simulation. Sol Energy 1990;44:243–8. doi:10.1016/0038-

421 092X(90)90052-E.

422 [28] Kanan S, Dewsbury J, Lane-Serff GF, Asim M. The Effect of Ground Conditions

423 under a Solar Pond on the Performance of a Solar Air-conditioning System. Energy

424 Procedia 2016;91:777–84. doi:10.1016/j.egypro.2016.06.243.

425 [29] Ganguly S, Date A, Akbarzadeh A. Heat recovery from ground below the solar pond.

426 Sol Energy 2017;155:1254–60. doi:10.1016/j.solener.2017.07.068.

427 [30] Singh TP, Singh AK, Kaushika ND. Investigations of thermohydrodynamic

428 instabilities and ground storage in a solar pond by simulation model. Heat Recover

429 Syst CHP 1994;14:401–7. doi:10.1016/0890-4332(94)90043-4.

430 [31] De Vries D. Thermal properties of soil. In: Van Wick W, editor. Phys. plant Environ.,

431 Amsterdam: North Holland; 1963, p. 210–35.

432 [32] Date A, Yaakob Y, Date A, Krishnapillai S, Akbarzadeh A. Heat extraction from Non-

433 Convective and Lower Convective Zones of the solar pond: A transient study. Sol

434 Energy 2013;97:517–28. doi:10.1016/j.solener.2013.09.013.

435 [33] Sayer AH, Al-Hussaini H, Campbell AN. New theoretical modelling of heat transfer in

436 solar ponds. Sol Energy 2016;125:207–18. doi:10.1016/j.solener.2015.12.015.

437 [34] Wang YF, Akbarzadeh A. A study on the transient behaviour of solar ponds. Energy

438 1982;7:1005–17. doi:10.1016/0360-5442(82)90084-6.

22
ACCEPTED MANUSCRIPT

439 [35] Kurt H, Ozkaymak M, Binark AK. Experimental and numerical analysis of sodium-

440 carbonate salt gradient solar-pond performance under simulated solar-radiation. Appl

441 Energy 2006;83:324–42. doi:10.1016/j.apenergy.2005.03.001.

442 [36] Rabl A, Nielsen CE. Solar ponds for space heating. Sol Energy 1975;17:1–12.

443 doi:10.1016/0038-092X(75)90011-0.

444 [37] Sodha M, Nayak J, Kaushik S. Physics of shallow solar pond water heater. Energy Res

445 1980;4:323–37.

446 [38] Suárez F, Tyler SW, Childress AE. A fully coupled, transient double-diffusive

447 convective model for salt-gradient solar ponds. Int J Heat Mass Transf 2010;53:1718–

448 30. doi:http://dx.doi.org/10.1016/j.ijheatmasstransfer.2010.01.017.

449 [39] Adams EE, Cosler DJ, Helfrich KR. Evaporation from heated water bodies: Predicting

450 combined forced plus free convection. Water Resour Res 1990;26:425–35.

451 doi:10.1029/WR026i003p00425.

452 [40] Henderson-Sellers B. Calculating the surface energy balance for lake and reservoir

453 modeling: a review. Rev Geophys 1986;24:625–49.

454 [41] Losordo TM, Piedrahita RH. Modelling temperature variation and thermal

455 stratification in shallow aquaculture ponds. Ecol Modell 1991;54:189–226.

456 doi:10.1016/0304-3800(91)90076-D.

457 [42] González D, Amigo J, Lorente S, Bejan A, Suárez F. Constructal design of salt-

458 gradient solar pond fields. Int J Energy Res 2016;40:1428-46. doi:10.1002/er.3539.

459 [43] Simunek J, Van Genuchten MT, Sejna M. The HYDRUS Software Package for

460 Simulating the Two- and Three-Dimensions Movement of Water, Heat, and

23
ACCEPTED MANUSCRIPT

461 Multiple Solutes in Variably-Saturated Media. 2nd ed. Prague, Czech Republic, PC-

462 Progress; 2011. doi:10.1007/SpringerReference_28001.

463 [44] Jaefarzadeh MR, Akbarzadeh A. Towards the design of low maintenance salinity

464 gradient solar ponds. Sol Energy 2002;73:375–84. doi:10.1016/S0038-

465 092X(02)00114-7.

466 [45] Akbarzadeh A, Andrews J, Golding P. Solar pond technologies: a review and future

467 directions. Adv Sol Energy 2005;16:233–94.

468 [46] Valderrama C, Gibert O, Arcal J, Solano P, Akbarzadeh A, Larrotcha E, et al. Solar

469 energy storage by salinity gradient solar pond: Pilot plant construction and gradient

470 control. Desalination 2011;279:445–50. doi:10.1016/j.desal.2011.06.035.

471 [47] Red Meteorológica del Instituto de Investigaciones Agropecuarias de Chile. Estación

472 Amolana, Tierra Amarilla n.d. http://agromet.inia.cl/estaciones.php (accessed July 20,

473 2017).

474 [48] DGA. Cuenca del río Copiapó: diagnóstico y clasificación de los cursos y cuerpos de

475 agua según objetivos de calidad. 2004.

476 [49] Leji FJ, Alves WJ, Van Genuchten MT, Williams JR. The UNSODA unsaturated

477 hydraulic database. Cincinnati, OH: 1996.

478 [50] Campbell G, Norman J. An introduction to environmental biophysics. Second. New

479 York: Springer International Publishing; 1998.

480 [51] Soto M. Hidrogeología e hidroquímica de aguas subterráneas en el distrito de Inca de

481 Oro, Región de Atacama: procesos de interacción agua-roca y dispersión geoquímica.

482 Universidad de Chile, 2010.

24
ACCEPTED MANUSCRIPT

483 [52] Garrido F, Vergara J. Design of solar pond for water preheating used in the copper

484 cathodes washing at a mining operation at Sierra Gorda, Chile. J Renew Sustain

485 Energy 2013;5:043103. doi:10.1063/1.4812652.

486 [53] Leblanc J, Akbarzadeh A, Andrews J, Lu H, Golding P. Heat extraction methods from

487 salinity-gradient solar ponds and introduction of a novel system of heat extraction for

488 improved efficiency. Sol Energy 2011;85:3103–42. doi:10.1016/j.solener.2010.06.005.

489 [54] Jaefarzadeh MR. Heat extraction from a salinity-gradient solar pond using in pond heat

490 exchanger. Appl Therm Eng 2006;26:1858–65.

491 doi:10.1016/j.applthermaleng.2006.01.022.

492 [55] Saxena AK, Sugandhi S, Husain M. Significant depth of ground water table for

493 thermal performance of salt gradient solar pond. Renew Energy 2009;34:790–3.

494 doi:10.1016/j.renene.2008.04.040.

495 Table Captions

496 Table 1. Monthly average climate conditions of Copiapó, Chile [47].

497 Table 2. Summary of the model assumptions and implications.

498 Table A.1. General parameters used in the model.

499 Table A. 2. Parameters for the different soils investigated in this study.

500 Figure Captions

501 Fig. 1. Characteristic zones in an SGSP and its heat extraction system.

502 Fig. 2. Heat fluxes representation in the transient thermal model.

503 Fig. 3. Seven-step algorithm to determine the mass flow rate to satisfy a constant heat demand.

504 Fig. 4. (a) Thermal evolution of the LCZ for different soil types; (b) Heat fluxes from the bottom of the
505 SGSP to the soil beneath, for different type of soils. Positive fluxes are directed upwards and negative
506 fluxes are directed downwards.

25
ACCEPTED MANUSCRIPT

507 Fig. 5. Soil thermal properties for water table depths (WTDs) at 7 and 20 m: (a) Soil moisture profile; (b)
508 Thermal conductivity envelopes; (c) Temperature envelopes.

509 Fig. 6. Comparison of the model results when water-dependent and constant soil thermal properties are
510 used: (a) Heat fluxes from the LCZ to the ground beneath it. Positive fluxes are directed upwards and
511 negative fluxes are directed downwards; (b) Sensitivity analysis of the water table depth (WTD).

512 Fig. 7. Heat fluxes from the LCZ to the ground beneath it for different water table depths (WTDs) and
513 different soil types. Positive fluxes are directed upwards and negative fluxes are directed downwards.

514 Fig. 8. Thermal evolution of the LCZ and the ground 1 m below the SGSP for different WTD and QD: (a)
515 WTD = 35 m, QD=20 W/m2, (b) WTD = 35 m, QD=30 W/m2, (c) WTD = 35 m, QD=50 W/m2, (d) WTD = 4
516 m, QD=20 W/m2, (e) WTD = 4 m, QD=30 W/m2, (f) WTD = 4 m, QD=50 W/m2.

517 Fig. 9. Thermal evolution of the SGSP (z ≥ 0 m) and the soil beneath it (z < 0 m): (a) without an insulation
518 layer; and b) with a 20-cm polystyrene insulation layer.

26
Solar Heat loss
radiation to atmosphere

UCZ Heat loss


to walls

NCZ Useful
Pump heat (QEHE)

EHE
Heat removal (QIHE)
LCZ
Tout IHE Tin

Heat loss
to ground
Solar Heat loss
radiation to ambient
xs xs

UCZ

Conductive
Radiation
fluxes

fluxes
NCZ

LCZ QLCZ/s
T=TLCZ
T=Tavg Lateral heat T=Tavg
loss (Qw)
T=Ts,1

Qs,i-1
Ground
Qs,i Layer i Δzi

T=Ts,N Qs/WT

Water table T=TWT


t = t+1 Step 3
(t) = f(Tout,QEHE,...)

Step 1 YES Step 2


TL(t) > Tmax QEHE = f(TL(t+1)=Tmax, TL(t),...)

NO

Step 4 Step 5 Step 6


m(t) = Tout= f(m(t),...) QEHE= f(m(t),Tout,...)

t = t+1

Step 7
m(t) = f(Tout,QOP,...)

NO

YES |QOP- QEHE| < ε?


100

80 Summer
T (°C)

60

Winter Clay
40
Loam
Sand

20
1 2 3 Year 4 5
(a)

10

Winter
Q(W/m2)

-10

-20
Clay
-30 Summer Loam
Sand

-40
1 2 3 4 5
Year
(b)
SGSP
0 0 0

WTD = 20 m

WTD
-1 -1 -1

7m=

m
WTD = 20 m

WTD

20
z (m)
z (m)

-2 -2

z (m)
-2

D=
Tmin Tmax

=7m

WT
m
7
=
TD
-3

W
-3 -3 Tmin Tmax
Tmin Tmax

Tmin Tmax
-4 -4 -4

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 20 30 40 50 60 70 80 90
θ (cm3/cm3) K (W/m/°C) T (°C)
(a) (b) (c)
10

0
Q (W/m2)

-10

-20
Dry
-30 Moist
Standard
-40 Water dependent

1 2 3 Year 4 5
(a)

-5
ΔTLCZ (%)

-10
WTD = 20 m
WTD = 9 m

WTD = 35 m

WTD = 49 m
-15

-20

-25
-80 -60 -40 -20 0 20 40
ΔWTD (%)
(b)
10 10

0 0
Q (W/m2)

Q (W/m2)
-10 -10

-20 -20

WTD = 7 m WTD = 7 m
-30 -30
WTD = 9 m WTD = 9 m
WTD = 20 m WTD = 20 m
-40 Clay WTD = 35 m -40 Loam WTD = 35 m

1 2 3 4 5 1 2 3 4 5
Year Year
(a) (b)

10

0
Q (W/m2)

-10

-20

-30 WTD = 7 m
WTD = 9 m
WTD = 20 m
-40 Sand WTD = 35 m

1 2 3 4 5
Year
(c)
100 100

80 80

T (°C)
T (°C)

60 60
QD = 20 W/m2
WTD = 35 m QD = 20 W/m2

40 40 WTD = 4 m
LCZ, with insulation Ground (z = -1 m), with insulation
LCZ, without insulation Ground (z = -1 m), without insulation

20 20
1 2 3 Year 4 5 1 2 3 Year 4 5
(a) (d)

100 100

80 80

T (°C)
T (°C)

60 60

40 QD = 30 W/m2 40 QD = 30 W/m2
WTD = 35 m WTD = 4 m

20 20
1 2 3 Year 4 5 1 2 3 Year 4 5
(b) (e)
100 100

80 80
T (°C)

T (°C)

60 60

40 QD = 50 W/m2 40 QD = 50 W/m2
WTD = 35 m WTD = 4 m

20 20
1 2 3 Year 4 5 1 2 3 4 5
Year
(c) (f)
ACCEPTED MANUSCRIPT

1 Tables

2 Table 1. Monthly average climate conditions of Copiapó, Chile [47].

Month Solar radiation Air temperature Wind speed Relative


(W/m2) (°C) (m/s) humidity (%)
January 329.3 22.0 2.2 39.4
February 298.6 21.0 2.2 40.6
March 249.5 19.3 1.8 47.5
April 226.0 16.9 1.4 49.0
May 168.0 14.5 1.3 40.8
June 154.6 14.2 1.5 23.1
July 159.4 14.4 1.5 22.6
August 195.3 16.1 1.5 28.1
September 245.5 16.4 1.6 30.4
October 291.8 18.0 1.7 33.3
November 337.3 18.8 1.9 34.2
December 347.8 20.9 2.1 36.0
3

10

11

12

13

14

15

16

17

18

19

1
ACCEPTED MANUSCRIPT

20 Table 2. Summary of the model assumptions and implications.

Assumption Model implication References


Heat fluxes in the SGSP, in the x and 1D thermal model. [35]
z-direction, are small in comparison to
those in the y-direction, so they can be
neglected.
UCZ and LCZ are completely mixed Temperature is uniform and it can be [19,33,41]
and homogeneous regions. treated as a single cell in the model.
Fresh water is flushed at the UCZ and The UCZ and LCZ thicknesses are constant [19,33,44,46]
a salt charger is used at the LCZ to in time.
maintain a saturated solution.
Groundwater temperature fluctuations Groundwater temperature is constant and [33,34]
throughout the year can be neglected. equal to the yearly average.
A minimum temperature difference of Convergence in the mass flow rate of the [32]
20°C between the LCZ and the heat heat exchange fluid required to satisfy a
exchange fluid is required for an specific heat demand (loop in the steps 5-6-
effective heat removal. 7, Fig. 3).
Turbidity in the SGSP is controlled SGSP thermal properties are not affected by [44-46]
effectively with hydrochloric acid. turbidity.
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
2
ACCEPTED MANUSCRIPT

40 Table A.1. General parameters used in the model.

Parameter (units) Symbol Value


Specific heat of water (J/kg-°C) CP,w 4,180
Specific heat of mineral (J/kg-°C) CP,w 870
Molar density of air (mol/m3) ρM 41.4
Atmospheric pressure (Pa) Patm 101,300
Shape factor (-) ga 0.1
Molecular mass of water (kg/mol) Mw 0.018 kg/mol
41
42
43
44
45
46
47

48

49

50

51

52

53

54

55

56

57

58
3
ACCEPTED MANUSCRIPT

59 Table A. 2. Parameters for the different soils investigated in this study.

Parameter (units) Symbol Value


Clay Loam Sand
Bulk density (kg/m3) ρb 1,100 1,400 1,600
Mineral density (kg/m ) 3 ρm 2,837 2,700 2,655
Thermal conductivity of the mineral km 2.3 2.0 1.7
fraction (W/m-°C)
Volumetric water content content in which θ0 0.25 0.15 0.05
the return flow cuts off (-)
Parameter that determines how quickly the q 6 4 2
cutoff occurs (-)
Residual water content (cm3/cm3) θr 0.068 0.078 0.045
Saturated water content (cm3/cm3) θs 0.38 0.43 0.43
Parameter in the water retention function α 0.008 0.036 0.145
(1/cm)
Exponent in the soil water retention n 1.09 1.56 2.68
function (-)
Moist thermal conductivity (W/m-°C) kmoist 1.1 1.25 2.0
Dry thermal conductivity (W/m-°C) kdry 0.2 0.5 0.5
Moist volumetric heat capacity (MJ/m3-°C) CP,moist 3.0 3.2 3.4
Dry volumetric heat capacity (MJ/m3-°C) CP,dry 1.0 1.2 1.4
60

61

62

63

64

65

66

67

68

69

70
4

You might also like