You are on page 1of 31

Retkowski W, Thöming J (2014) Applied Energy (114), 492-503

1 Thermoeconomic optimization of vertical ground-source heat


2 pump systems through nonlinear integer programming
3 Waldemar Retkowski, Jorg Thöming

4 Abstract
5 Vertical ground-source heat pump systems (GSHPSs) use the ground´s undisturbed relative
6 constant temperature as a source for space heating of residential and commercial buildings.

Accepted version
7 The design of GSHPSs is focused in finding the optimal depth and amount of boreholes and
8 also the connected power requirement like the amount and size of heat pumps. In this paper a
9 mixed-integer nonlinear programming (MINLP) approach to solve the design problem of a
10 vertical GSHPS is presented. The resulting mathematical model includes the calculation of the
11 total annual costs (TAC) and the coefficient of performance to obtain estimates of both
12 economic and ecological relevance to design an optimal equipment set-up. For desired
13 constraints the numerically optimal values of the design parameters (borehole depth, mass
14 flow rate, number of boreholes, type and number of the heat pumps) were calculated. Two
15 numerical solution alternatives are investigated, namely Generalized Reduced Gradient
16 (GRG2) and evolutionary algorithm. The GRG2 approach provides a more stable and faster
17 optimal solution. Calculated results are presented through a validation example. The
18 evaluation of the proposed objectives and studied sensitivity effects present the applicability
19 of the model. This method was able to improve the TAC about more than 10%.

20

21

22

23

24

25

26

27

28

29

30

31

32

1
Retkowski W, Thöming J (2014) Applied Energy (114), 492-503

Nomenclature Parameters
Yearly operating hours (h y-1)
Abbreviations
Air temperature (°C)
MINLP Mixed-integer nonlinear programming Electrical efficiency (-)
MILP Mixed-integer linear programming Start-up cycles (h-1)
GSHPS Ground-source heat pump system Geometrical expressions (-)
GSHP Ground-source heat pump
Thermal resistance (m K W-1)
EED Earth energy designer
Empirical based function
GHE Ground-heat exchanger

Accepted version
Full operating time period (h)
Diff. Difference
Heating season time period (h)
Start-up time period (s)
Indices
Ampere (A)
f Fluid
Volt (V)
m Half borehole length
Thermal conductivity (W m-1 K-1)
g Grout
Heat load (kW)
gr Ground
Heat flux (W m-1)
gs Ground surface
b Borehole Radius (m)
geo Geological Half shank space (m)
dwn Down Heat capacity (J kg -1 K -1)
up Up Density (kg m-3)
ln Natural logarithm Viscosity (m²s-1)
s Soil C Costs (€)
∼ Dimensionless variable
Mean value Variables
Massflow rate (kg s-1)
max Maximal
Temperature (K)
min Minimal
Total annual costs (€ y-1)
eva Evaporator
M Manufacturer Investment costs (€)
i Different equipment sections OC Operating costs (€)
Ground heat (kW)
HC Heating circuit
Length (m)
HP Heat pump
SC Soil circuit Integer Number (-)
Q Heat Evaporator heat (kW)
P Power Condenser heat (kW)
tot Total Electrical power start-up (kWh)
Heat coefficients Electrical power operating (kWh)
Power coefficients Coefficient of performance (-)
SPF Seasonal performance factor (-)
p Pipe
Temperature out of SC (K)
dem Demand
Temperature into the SC (K)
load Load
Disturbed Temperature (K)
pi Inner pipe
po Outer pipe Flow rate (m³ h-1)

2
Retkowski W, Thöming J (2014) Applied Energy (114), 492-503

33 1. Introduction
34 One way to exploit sustainably produced electrical energy is in using a vertical ground-source
35 heat pump system (GSHPS) to supply the heating equipment in residential and commercial
36 buildings. The European number of these installed systems is rapidly growing. Extrapolating
37 currently observed growth rates for Europe of 5.4 million heat pump units per year, let expect
38 a number of 70 million installed units in Europe for 2020 [1]. Along with this increasing
39 relevance and impact there is a rising demand for optimal designed vertical GSHPSs.

Accepted version
40 The research and developments of vertical GSHP technology based on various mathematical
41 models and systems is described in detailed reviews [2-4]. Over the years different analytical
42 [2,4], numerical [5-8] and hybrid [2,4] models have been developed especially to calculate
43 the thermal behavior of ground-heat exchangers (GHEs). Therefore is the simulation of these
44 systems an important tool for system design purposes. These approaches focus on the thermal
45 ground behavior and are often time consuming techniques even for experienced users [9].
46 Furthermore is the optimal sizing of GHEs important because of the high drilling costs and
47 the design challenge of sizing an optimal borehole thermal capacity with an optimal capacity
48 of heat pumps [10]. To design competitive GSHPSs involves thorough technical also
49 economic considerations. In these consequences arises the need to consider simultaneously
50 the soil cycle, the heat pump cycle and their economic aspects for optimal design purposes.

51 As one early work Wall [11] analyzed heat pump systems and pointed out that a
52 thermoeconomic optimization is an economic optimization in conjunction with thorough
53 thermodynamic description of the system. In [12] they developed a method to minimize the
54 entropy generated in a heat exchanger, whereupon an optimum u-tube length and diameter
55 was determined. Sayyaadi et al. [13] optimized a vertical GSHP for a given cooling load.
56 Seven temperature differences and one pipe diameter for the ground heat exchanger were
57 chosen as design variables. They minimized a thermodynamic-, a thermoeconomic- and a
58 multi-objective and considered the sensitivities of the interest rate, operating hours and the
59 cost of electricity. In [14] they optimized a vertical GSHP for given heating and cooling
60 loads. They developed a nonlinear optimization model and applied for a thermoeconomic
61 optimization eight temperature design parameters and one nominal pipe design parameter. Li
62 and Lai [15] provided analytical expressions for optimizing flow velocity and borehole length
63 by applying the entropy generation minimization method for GSHPs with a single U-tube.
64 Their analyses indicated the existence of optimum parameters based on pure heat transfer and
65 thermodynamics ground. In [16] they proposed an algorithm for optimization of cooling
66 tower-assisted GSHPs applying 12 decision variables and additional constraints. Their
67 sensitivity consideration of costs showed that the product cost of all regarded systems
68 increased due to an increasing interest rate. Lee et al. [17] used for GSHP optimization an
69 objective function representing the initial system costs divided by the annual energy
70 production as a measure of cost effectiveness, which should be minimized. With an emphasis
71 on building optimization [18] used three objective function criteria, the total cost of the
72 system, the primary energy saving and the CO2 emission costs. They applied different values
73 for different penalty parameters and focused applying a mixed-integer linear program (MILP)
74 on optimization robust building loads. Florides et al. [19] investigated the cost and efficiency

3
Retkowski W, Thöming J (2014) Applied Energy (114), 492-503

75 impact of double U-tubes, single U-tubes and parallel or serial arrangements. They concluded
76 that the building costs of double U-tubes are 22-29% higher than of single U-tubes. While
77 their parallel configuration is more efficient by 26-29%, while the series configuration by 42-
78 59%. The assumption of a maximization of heat pump performance due to a minimization of
79 ground temperature changes was followed by [20]. They achieved a balanced ground cooling
80 with the application of GHE distributed loads. A connection to heat pumps or modeled flow
81 through GHEs had there not a special emphasis. A related work of [21] pointed out that an
82 optimization potential lies in the regulation of the energy extraction of load per BHE
83 individually and a thereby balanced cooling of the ground. An applied seasonally variable

Accepted version
84 heat demand aroused that the maximum temperature in the ground migrated by 10-15%. They
85 concluded that by a priori choosing strategic BHE arrangement the long-term performance of
86 the system could be improved, also without load regulation and connected additional
87 investments. Michopoulos and Kyriakis [22] regarded an overestimation of the electricity
88 consumption of their heat pump of max. 3.8%, due to deviations between measured and
89 predicted temperature values of about ±2 °C.

90 Other authors focused on the energy and exergy analysis of GSHPs and promise knowledge
91 about optimization potential. Hepbasli and Akdemir [23] analyzed the energy and exergy of a
92 GSHP system and concluded that the most potential for improvement is in the compressor,
93 followed by the condenser and then the expansion device. A comparison of the
94 irreversibility’s associated with the heat transfer process in the evaporator and GHE showed
95 greater irreversibility’s in the evaporator due to the operation at a lower temperature than the
96 GHE. The losses of an investigated GSHPS were due to the electrical, mechanical and
97 isentropic efficiencies and emphasize the need for paying close attention to the selection of
98 this type of equipment. The second largest irreversibility was due to the condenser, the third
99 the expansion device. The evaporator had the lowest irreversibility on the basis of the heat
100 pump cycle [24]. Wall [11] emphasized that the difference of all incoming and outgoing
101 exergy flows must be minimized. Bi et al. [25] specified the location of maximum exergy loss
102 ratio with the compressor, while the location of minimum exergy efficiency and
103 thermodynamic perfect degree is the ground heat exchanger. Torío et al. [26] distinguished
104 with an extensive exergy study that a holistic view of a GSHP system and especially the
105 interactions and components should be regarded to optimize the system.

106 Only rare optimization approaches applied a mathematical programming considering both
107 crucial cycles. From our knowledge none author up to now has given preferential
108 consideration on a discrete choice analysis approach, as a MINLP framework has enabled to
109 us, which allows system optimization of real existing components. Due to the finding of
110 Shonder et al. [27] is the peak-load method the most consistent one compared with other
111 tested design programs. For the purposes of this analysis, a design heating day should be
112 determined from numerical investigations. Traditional design methods of GSHPSs involve
113 many trials in order to meet the design specifications [9]. This can be avoided through the
114 present design method, which considers the minimization of annual costs, or maximization of
115 ecological benefits or the generation of Pareto-optimal designs while technical constraints are
116 satisfied. The present study demonstrates for the first time an integer-based optimization
117 procedure that involves the entire system, consisting of a ground and also heat pump circuit

4
Retkowski W, Thöming J (2014) Applied Energy (114), 492-503

118 and allows for choosing real units. This approach helps by this to avoid oversized or
119 undersized systems. For this purpose a new mixed-integer nonlinear programming (MINLP)
120 approach has been developed, validated and analyzed.

121 2. Problem statement


122 Given are a huge range of different potential GSHPS configurations, a fixed time interval,
123 fixed and relative investment costs, relevant prices, lower and upper bounds on equipment and
124 physical behavior, heat pump data base, energy balances, some physical properties, a given
125 total heat load and costs associated with the GSPHS operation. The goal of each GSHPS

Accepted version
126 design is to determine the optimal technical configuration with specific components,
127 especially heat pumps, well size, well amount and mass flow rate, fulfilling the physical and
128 technical needs along with the planning decisions that minimize the TAC, or maximize the
129 environmental savings, or determines the Pareto-optimal design.

130 3. Assumptions and mathematical model


131 3.1. Assumptions and requirements

132 The proposed numerical optimization scheme has the following general assumptions:

133  the ground heat conductivity is a mean of all ground layers, which is sufficient [28];
134  the thermal properties of all the materials are means and remain constant as typical for
135 steady state modeling;
136  the design case of GHE fields neglects the thermal interaction between boreholes;
137  special thermal effects in the ground where reasonable simplified or neglected and
138  the fluid flow rate in each GHE tube is equal.
139
140 One general typical strategy is to use ground related simulation results as an input for design
141 programs [27]. This approach is strictly forward and being recommended from us during an
142 application of the new developed design algorithm. Complementary one should execute a
143 specific thermal response test on site in advance to get proper ground properties as the
144 undisturbed temperature, the maximum available heat flux and the heat conductivity of the
145 ground.

146 3.2. Mathematical formulation

147 Mixed-integer nonlinear programming (MINLP) is a numerical optimization approach of


148 mathematically formulated problems to simultaneously optimize structure (using discrete
149 variables) and parameters (using continuous variables) of a system. The mathematical
150 idealization of the thermo-physical and economical programming for generating an optimal
151 solution of a GSHPS configuration design tasks can be expressed as described in the
152 following sections.

153 3.2.1. Heat production subsystem

154 The heat production is taken over by two closed cycles: the soil circuit and the heat pump
155 circuit (Fig.1). The main task for each cycle is to fulfill the energy balance and to generate

5
Retkowski W, Thöming J (2014) Applied Energy (114), 492-503

156 relevant temperature information’s. Both cycles are coupled and this is also being expressed
157 by the linking constraints.

158 3.2.1.1. Soil circuit constraints

159 The soil cycle reaches from the deepest point of the boreholes respective tubes up to the
160 evaporator of the heat pump. Main task is to allocate ground-heat due to estimating the
161 total needed heat exchanger length associated with an estimated maximal possible heat
162 flux from the ground. The formulation is shown in Eq. (1).

Accepted version
163 (1)

164 For simplicity one could determine the discrete chosen number of boreholes with parallel
165 supplied boreholes and a single borehole length , shown in Eq. (2), out of the requested
166 total heat exchanger length.

167 (2)

168 (3)

169 If not other mentioned were here typical values for (analyzed in section
170 5.3.2) and used.

171 (4)
172 The ground-heat through the evaporator is idealized with and carried by an optimal
173 mass flow and specific heat capacity and is given in Eq. (4). For simplicity was
174 here a mean fixed value for our proposed calculations used (Appendix A.3).

175 3.2.1.2. Heat pump circuit constraints

176 The heat pump cycle supplies the required heat to the heating circuit. The functions and
177 represent an individual heat pump and are usually not known explicitly. However, heat
178 supply and power data, at variable fluid temperatures , are often given by the manufacturer
179 as a function of the capacity for a specific heat pump. The details are given in section 3.3 and
180 Appendix A.

181 (5)

182 (6)

183 With Eq. (5) and Eq. (6) one can determine the discrete chosen number of heat pumps
184 the specific produced heat and estimated electrical demand of each heat pump. With the sum
185 of produced thermal heat one could guaranty the requested amount of heat . The

6
Retkowski W, Thöming J (2014) Applied Energy (114), 492-503

186 heat pump start-up process, modeled in Eq. (7) and Eq. (8), requires pre-defined start-up
187 cycles and asks for total additional power to be supplied.

188 (7)

189 (8)

190 with

Accepted version
191 The shown start-up heat pump power supply calculation considers a three-phase current basis,
192 as typical for the considered heat pumps. A total electrical demand results for a certain
193 operating time interval and can be calculated with Eq. (9).

194 (9)

195 The total chosen heat pumps should contain at least one unit; this is
196 expressed in Eq. (10).

197 (10)

198 It is necessary to calculate the from the ground to be provided heat . The electrical power
199 should be converted into thermal energy; this efficiency can be expressed with the factor .
200 The about the electrical power to heat efficiency factor reduced heat pump electrical power
201 demand should be managed with the heat provided by the heat pumps as shown in
202 Eq. (11). This couples the heat pump circuit with the soil circuit and gives the heat demand
203 to be supplied by the soil circuit.

204 (11)

205 Related to the specific heat pump and to the individual characteristic curve one can
206 define a maximum number of heat the selected heat pump could supply, which is
207 realized in Eq. (12). Also the maximal requested specific electrical power is limited
208 by as shown in Eq. (13).

209 (12)

210 (13)

211 (14)

212 The current mass-flow rate should be converted to flow rate and not fall below
213 a manufacturer´s minimum design flow rate . This ensures unwanted switching of by
214 the controller due to a lack of adequate fluid. These parameter values can be taken from

7
Retkowski W, Thöming J (2014) Applied Energy (114), 492-503

215 manufacturer´s data sheets. To calculate simply one could divide the design variable by
216 the temperature dependent density given in Appendix A.

217 3.2.1.3. Temperature constraints

218 The mean fluid temperature as crucial variable of the soil circuit has to be calculated. A
219 simplified calculation can be realized by building the mean of the soil fluid inlet and soil
220 fluid outlet temperatures, or as shown in Eq. (15).

Accepted version
221 (15)

222 These temperatures depend strongly on the estimated or measured maximal heat flux
223 and the so-called mean borehole resistance , which can be defined as in Eq. (16) and Eq.
224 (17).

225 (16)

226 with , where is assumed with

227 (17)

228 A mean temperature located at the half of a single borehole depth is being calculated
229 with an undisturbed temperature fraction and a depth-dependent fraction related to the
230 specific ground gradient . The value depends strongly on the soil properties. Often used
231 values vary about approx. 0.01 °C m-1 up to 0.05 °C m-1. Typical values for Germany vary
232 between approx. 0.025-0.035 °C m-1. One could calculate the local ground gradient
233 dependency as in [5] upon the depth with

234 (18)

235 It is assumed that the average boundary temperature at the ground surface is approx. 1-2 °C
236 higher than the average annual air temperature .

237 (19)

238 (20)

239 The mean fluid temperature depends in particular on the existing specific thermal borehole
240 resistance. Florides et al. [29] highlighted the great importance of the proper borehole filling.
241 A classical thermal resistance arrangement of a borehole is indicated in Fig. 2. Lamarche et al.
242 [30] have published an overview and evaluation of common methods to calculate the borehole
243 resistance in GSHP systems. They recommended the multipole method proposed by Bennet et
244 al. [31] to solve the problem and mentioned that with the solution in the form of an infinite

8
Retkowski W, Thöming J (2014) Applied Energy (114), 492-503

245 series (multipole expansion) one can compute pipe related steady-state conductive heat flows.
246 The calculation of the thermal resistance is given by Eqs. (21)-(23).

247 (21)

248 ; ; (22)

249 (23)

Accepted version
250 Lamarche et al. [30] pointed out that the first term in Eq. (23) is known as line-source formula
251 and the second term was proposed as first-order multipole correction [30-32] and that this
252 expression is also used in the EED design software [33] and GLHEPRO 4.0 [18]. The range
253 of values for is approximately between 0.01-0.8 K m-1 W-1 with typical values for
254 Germany between 0.10-0.35 K m-1 W-1. For several test cases Lamarche et al. [34] have
255 calculated a maximal error between analytical and numerical solutions of only 0.9%. The
256 maximal error of the other tested methods varied between 18% and 47%. The overall estimate
257 for all tested methods is between 43% and 150%.

258 To avoid frozen layers close to the GHE and damage of equipment German engineer
259 standards [35] restrict as shown in Eq. (24) the returning heat carrier fluid for a peak load with
260 and for a constant load with .

261 (24)

262 Additionally one should restrict the lowest fluid temperature depending on the binary mixture,
263 in our case approx. -10 °C [44]. Ensuring a proper heat transfer through the evaporator one
264 should set a lower bound of approx. 2 °C and an upper bound of approx. 7 °C
265 as shown in Eq. (25) which are derived from experience.

266 (25)

267 3.2.1.3. Linking constraints

268 A crucial linking constraint is formulated in Eq. (26). With this constraint should the heat
269 balance from the soil up to the heat pumps been guaranteed.

270 (26)

271 This ensures with the given heat extraction capacity shown in Eq. (27) the fulfillment of the
272 heat demand derived from the parallel executed heat pumps and soil capacity. The second
273 crucial equality constraint guarantees the proper generated heat provided by the optimal
274 heat pump configuration.

275 (27)

9
Retkowski W, Thöming J (2014) Applied Energy (114), 492-503

276 3.2.1.4. Performance indicators

277 The efficiency and therefore the environmental impact under certain conditions might be
278 expressed as coefficient of performance (COP) as shown in Eq. (28).

279 (28)

280 To get an approx. for a specific year of operation one could apply after an optimization
281 run the annual average heat flux, calculated as shown in Eq. (29) or directly derived from
282 measurements, to get according model outputs with as the hours of an specific year. This

Accepted version
283 approach is equal to a simplified averaged COP, where in Europe often an averaged COP is
284 called SPF. Here is COP used as for one operating point valid, regarding the input it is
285 especially the peak heat flux, instead is the SPF connected to a longer term, expressed as
286 averaged heat flux.

287 (29)

288 3.2.2. Economic constraints

289 The economic sub-model is based on two simplified main economic factors which are the
290 investment cost (IC) and the operating cost (OC). The sum of these components should be
291 minimized. IC includes the cost of heat pumps, heat exchangers and an average amount for
292 connecting these components, namely additional drilling cost. For this value one can
293 assess in Germany between approx. 25%-35% of the borehole drilling costs and gets the
294 additional investment installation costs (includes PE pipes, filling material, etc.). Mainly
295 the electricity required on site is occupied to determine the OC. The economical sub-model is
296 expressed as follows in Eqs. (30)-(33):

297 (30)

298 (31)

299 (32)

300 (33)

301 3.3. Heat pump data basis

302 The from manufacturer´s provided technical data sheets are containing information’s about
303 characteristic curves of specific heating capacity and required electrical power. The through
304 regression resulting functions are related to the fluid temperature. Note that depending on the
305 type of the characteristic curve it may fit better to apply the model based fluid temperature .
306 And due to specific restrictions on measuring system may some additional restrictions be
307 required. Out of this empirical data were coefficients (Appendix A.1) and functions in the
308 general form of Eq. (34)-(35) determined. The used data bases contained three or four
309 different heat pumps, each valid for a heating circuit temperature of 35 °C. The temperature

10
Retkowski W, Thöming J (2014) Applied Energy (114), 492-503

310 of 35 °C is a typical water temperature level used in floor heating systems which promise a
311 good efficiency for heat pump systems. These coefficients produce a bridge to the real
312 problems a designer is being faced: the selection of a proper heat pump and circuit pump
313 fulfilling the specific constraints and boundary conditions.

314 (34)

315 (35)

Accepted version
316 The specific upper bounds are respectively given by the maximum thermal and electrical
317 properties and attend as inequality constraints as shown in Eqs. (12)-(14). It should be taken
318 care of the constraints that the maximal or minimal known allowed heap pump specific values
319 are not violated. Note that related on the type of characteristic curve and resp. heat pump one
320 might restrict the temperature spread in Eq. (25) as well for a constant level of 3 K, which
321 dependents on the preconditions of the specific heat pump.

322 3.4. Objective functions

323 Three different objective functions were applied to illustrate the capability of the new method
324 to generate powerful solutions. The main focus was taken on the investigation of the technical
325 design by a minimization of the costs in the first year, which is realized with Eq. (37). With
326 the presented formulation of Eq. (38) it might be possible to get a compromise between a
327 thermo-economic design and a focus on the crucial ecological impact. And with the Eq. (39) it
328 is possible to maximize the crucial performance indicator COP. All optimizations were
329 undertaken with the constraints given in Eqs. (1)-(14), Eqs. (16)-(28) and Eqs. (30)-(36).

330 3.4.1. Thermo economic performance

331 The total annual cost (TAC) of the first year of a specific geothermal system is given by Eq.
332 (36). The TAC function should be minimized due to variation of the formulated design
333 variables by the chosen solving method.

334 (36)

335 The process variables and are modeled as non-negative stationary design variables.
336 The integer design variable is representing the chosen specific heat pump provided in a
337 database (Appendix A). The integer variable is responsible for a proper amount of wells.

338 (37)

339

340 3.4.3. Pareto optimal performance

341 As objective function to calculate Pareto optimal solutions were the already shown Eq. (36)
342 and Eq. (28) taken. The division results in Eq. (38).

343

11
Retkowski W, Thöming J (2014) Applied Energy (114), 492-503

344 (38)

345

346 3.4.2. Environmental performance

347 Eq. (28) can be taken as optimization criteria and supplemented as objective function as
348 shown in Eq. (39).

349 (39)

Accepted version
350

351 4. Solution algorithms


352 The new GSHP design method was solved through application of a Generalized-Reduced-
353 Gradient-2 algorithm (GRG2) and an Evolutionary algorithm (EA). Both solvers are
354 implemented in the basic Microsoft Excel 2010 environment.

355 4.1. Generalized reduced gradient 2 (GRG2)

356 The Microsoft Excel 2010 Solver employs the GRG2 Algorithm [36] for solving nonlinear
357 problems [37-38]. The method extended first-order reduced gradients with second-order
358 information based on reduced Hessian. This enables solutions also of constrained nonlinear
359 problems. Integer constrained nonlinear problems are solved by a branch and bound algorithm
360 which starts an optimization process by solving the relaxed problem using GRG2 [39].
361 Iteratively solved sub-problems update the best bounds with the best objectives until a by the
362 user fixed tolerance is satisfied as shown in Eq. (40). The number of sub-problems may
363 grow exponentially [40].

364 (40)

365 4.2. Evolutionary algorithm (EA)

366 The applied method follows a nondeterministic approach and is a hybrid, based on the
367 principles of genetic algorithms and evolutionary algorithms. In a genetic algorithm the
368 problem is often encoded in a series of bit strings that are manipulated by the algorithm. In an
369 evolutionary algorithm the decision variables and problem functions are used directly. It
370 generates many trial points and uses “constraint repair” methods to satisfy the integer
371 constraints. The constraint repair methods include classical methods, genetic algorithm
372 methods, and integer heuristics from the local search literature. This approach cannot
373 guarantee the optimality [40].

374 5. Application of the new GSHP optimization model


375 The provided model evaluation includes a validation case where a comparison between model
376 results and measurements (Fig. 3) under equal conditions is presented. A constitutive
377 comparison of optimal results applying the proposed objectives is shown in Tab. 2.

12
Retkowski W, Thöming J (2014) Applied Energy (114), 492-503

378 Furthermore follows in section 5.3. an extensive analysis of the optimization model with three
379 hypothetical comparable case studies, where every point entails a different optimal GSHP
380 design.

381 5.1. Validation case

382 Design and measurement data were taken from [41] and compared with results of the crucial
383 variables generated by this method. The origin construction of the associated values was
384 equipped with two ground-coupled heat pumps, each with a max. capacity of approx. 82 kW.
385 This heat pump type is included in the provided data base and referred to as Type 3 (Tab.

Accepted version
386 A.5). For validation purpose was only one single heat pump considered and related results for
387 the electrical power demand can be taken from Fig.3C. A ground-source heat exchanger field
388 with a total length of 2400 m and 16 boreholes was installed on site. Half of the pipes were up
389 to 8 m insulated, whereupon for validation only the separate available data with no influence
390 of the insulation was used. Regarding the distributed annual temperatures contained a free-
391 cooling and a heating period. The average fluid inlet temperature became for this period 8.8
392 °C and the outlet temperature 6.8 °C into the ground. The plant was designed with a average
393 heat flux of 35.6 W m-1. This value could nearly be confirmed with 34 W m-1 measured. In
394 October of the first heating period the maximal heat flux was 39.3 W m-1, what is used as
395 criterion for the calculated design cases shown in Tab. 2. As an annual average was 13.02
396 W/m calculated. The heat conduction of the submerged borehole filling material was
397 estimated with 1.45 W m-1 K-1 and the PE-pipe leg distance with 0.035 m. The crucial rare
398 available data is tightly arranged in Tab. 1. Further in this section not given values can be
399 taken from and Tab. 6.

400 Table 1
401 Design conditions and measurements taken from [41].
Parameter Value Unit
Maximal heat load 111.2 kW
Maximal heat flux 39.3 W m-1
Mean soil conduction 2.4 W m-1 K-1
Full operation hours 4111 h y-1
Average air temperature 8 °C
Heating circuit temperature 35 °C
Borehole diameter 0.115 m
Pipe diameter 0.032 m
Mass flow rate 0.16 kg/s
402

403 For one heating season are compared results between measurements and modeled data shown
404 in Fig.3. The comparison emphasizes the mean fluid temperature given in Eq. (15) and the
405 electrical power demand given in Eq. (9) inclusive all related equations. These values
406 represent the main target values of operating conditions. For the calculation of the mean fluid
407 temperature were the measured single borehole depth, mass flow rate, heat conductions, the
408 pipe radius, the borehole radius and the leg distance of PE-pipes, the heat flux and borehole
409 temperatures used as inputs. According to the plant operation were the measured heat flux and
410 borehole temperatures varied, whereupon the other inputs were left constant. Depending
411 values can be taken from Fig.3A-B. The heat pump operated from 300 up to 691 h a month.

13
Retkowski W, Thöming J (2014) Applied Energy (114), 492-503

412 Note that each measured monthly heat flux was converted to 740 h to get averaged ground
413 related values for each month. To compare proper the borehole temperatures the measured
414 points were reduced by an average of 0.3 °C, due to slightly differences between the modeled
415 (borehole edge) and measured (inside a borehole) temperature position. To compare the
416 electrical power demand as shown in Fig.3C the measured mean fluid inlet temperature and
417 the heat pump operation hours were time dependent applied, as well as a fixed heat pump
418 specific operating caused conversion factor of 0.85, which was derived from data. The
419 relative error between the measured and modeled data was in these cases under 1% (Fig.1C-
420 D). The heat pumps can be assumed to be validated due to publication of empirical based

Accepted version
421 characteristic curves. The R-squared value was for every in this paper applied heat pump
422 above 0.99.

423 5.2 Evaluation of the objectives

424 Further a constitutive comparison of optimal results applying the proposed objectives and a
425 not optimized base case are shown in Tab. 2. The calculated optimal data were generated by
426 application of the objective functions given in Eqs. (37)-(39) and the constraints given in Eqs.
427 (1)-(14), Eqs. (16)-(28) and Eqs. (30)-(36). The data base of considered heat pumps can be
428 taken form Table A.5. The data base includes also the heat pump type used on site, namely
429 Type 4. For comparability similar model values were inserted, they are given in section 5.1,
430 Table 1 and Table A.6. The calculated optimal results provided by GRG2 and EA can be
431 taken from Table 2. The GRG2 considered a solution accuracy of 10-4, used automatic
432 scaling, forward differences, and zero as lower bound for not constraint variables. The EA
433 considered a tolerance of 10-2, a mutation rate of 0.5, population of 100, automatic scaling and
434 zero as lower bound for not constraint variables. The optimality criterion for integer design
435 variables was in both cases zero.

436 Table 2
437 Base case values and optimal results obtained by the proposed algorithm solved with the
438 GRG2 and EA. The optimal number of heat pumps is aligned with Tab. 5 and arranged
439 as follows: Type 1/ Type 2/ Type 3/ Type 4. All given differences are related to the base case.

Units - € - m m - °C
Base case 0/0/2/0 267516 16 150 2400 5.32 5.51 7.7
Eq.(37) – GRG2 0/0/0/6 239670 8 295 2358 5.73 5.92 9.5
Diff. [%] -10.4 -50 96.7 -1.8 7.7 7.4 23.4
Eq.(37) – EA 1/1/0/0 254603 15 122 1831 4.52 4.65 7.4
Diff. [%] -4.8 -6.3 -18.7 -23.7 -15.0 -15.6 -3.9
Eq.(38) – GRG2 0/0/0/6 239670 8 295 2358 5.73 5.92 9.5
Diff. [%] -10.4 -50 96.7 -1.8 7.7 7.4 23.4
Eq.(38) – EA 1/1/1/2 279421 5 277 1387 4.97 5.12 9.3
Diff. [%] 4.5 -69 84.7 -42.2 -6.6 7.1 20.8
Eq.(39) – GRG2 0/0/0/9 249001 7 303 2121 5.74 5.93 9.7
Diff. [%] -6.9 -56 102 -11.6 7.9 7.6 26.0
Eq.(39) – EA 0/0/0/9 249156 7 304 2122 5.74 5.93 9.7
Diff. [%] -6.9 -56 103 -11.6 7.9 7.6 26.0

14
Retkowski W, Thöming J (2014) Applied Energy (114), 492-503

440 The in Table 2 shown differences between the calculations of the two solving methods and the
441 base case have a causal connection to a change of the single well amount, well length, heat
442 pump capacity and arrangement. The system with the about 10.4% improved costs delivers
443 parallel an improved efficiency of about 7.7%. The best performance of the individual optimal
444 GSHPS provided by the EA and the GRG2 is almost 8% higher than the base case. It may be
445 useful to compare the calculated COP, as the algorithm considers operating points. Looking
446 for a longer term in literature was the annual SPF given with 4.6 of the regarded heating
447 season. Related to this value exists an optimization potential up to the best numerical annual
448 SPF with 5.93 of 28.9%. This could be assessed on the basis of additional optimized control

Accepted version
449 and conversion strategies, which are not subject matter here. Along with the objective related
450 increased optimal performance decreased the TAC about 18.515 € which is equivalent to
451 almost 7%. In accordance to the assumptions the optimal fluid temperature and total well
452 length seemed to be sufficient estimates. As expected the total optimal length decreased along
453 with a lower efficiency and increased along with a higher efficiency. The required optimal
454 start-up energy varied between 419 and 628 kWh/a.

455 5.3. Three case study series

456 In the following sensitivity study three separate hypothetical design series are considered. In
457 these case studies optimal values for three different case series were calculated. In the first
458 case series was the input parameter heat load varied, in the second case the heat flux and in
459 the third the mean outside temperature. All the other input values were left constant. The used
460 values are given in Table 3 and Table A.6. The three heat pumps in the data basis are
461 providing a heating system temperature of 35 °C and the associated parameter values are
462 given in Table A.4.

463 5.3.1. Model set-up

464 The heat load was varied with an interval of 10 kW between 60 kW and 350 kW. This range
465 can be assigned to the group of new industrial size GSHPSs. In the other two series the heat
466 load was fixed with 100 kW. The range for heat flux values was chosen between 31 W m -1
467 and 59 W m-1 with an interval of 1 W m-1. For the third series the average outside temperature
468 varies between 5.5 °C and 14.5 °C with an interval of 0.5 °C. Further input values were taken
469 for the ground properties, hydraulic properties, operating conditions and economical
470 properties and can be taken from Table 6. The related results are shown in Fig.4-6, for heat
471 flux variations they are indicated with A, for heat load variations they are indicated with B
472 and for temperature they are indicated with C.

473 Table 3
474 Parameter values and -ranges used as input variation.
Heat load 60-350 100 100 kW
Heat flux 50 31-59 50 W m-1
Average annual temperature 10 10 5.5-14.5 °C y-1
475
476

477

15
Retkowski W, Thöming J (2014) Applied Energy (114), 492-503

478 5.3.2. Results and discussion

479 Optimization programming for the sensitivity study was solved with the GRG2 algorithm and
480 calibrated as shown in section 5.2. Additional used values are presented in Tab. 6.

481 5.3.2.1. Optimal investment and operational costs

482 The optimal values for investment and operational costs are shown in Fig. 4. With an
483 increasing heat flux are the investment costs falling, but the operating costs remain almost the
484 same. The integer caused variations indicate the dependent relationship between the

Accepted version
485 investment and operational costs. This seems to be caused by a connected change in the
486 optimal heat pump configuration. Even though discrete variables exist both cost components
487 are with an increasing heat demand almost parallel rising. Surprisingly investment and
488 operational costs related to a variation of heat load is almost smooth and linear. The rising
489 average outside temperature tends to result in lower costs. The impact of the discretized
490 solution domain shows compared to the other ones the main significance.

491 5.3.2.2. Optimal heat pump configuration

492 Information about the optimal type and number of heat pumps for each design case provided
493 by solutions of the optimization model can be taken from Fig. 5. A huge range of optimal
494 points related to a change of the heat flux remain almost constant with a tendency to an
495 increased total number of heat pumps along with a higher capability of the ground. The
496 second case study series shown in Fig. 5B with solutions of optimal heat pumps created
497 through a varying heat load corresponds to the progress of investment and operational costs.
498 With a higher heat load the total amount of heat pumps increases almost linear and stepwise.
499 The heat pump type tendency gradually switches with higher heat loads to a higher number of
500 heat pump types, in our case study number three, as shown in Table A.4, this is the major heat
501 pump in the provided database. The left heat load gap is filled with a combination of the other
502 fitting and cheaper heat pumps. The third case study series shown in Fig. 5C shows a step at
503 10 °C and this is significant correlated with the costs. A change in the average outside
504 temperature causes an only slightly lower number of necessary heat pumps.

505 5.3.2.3. Optimal well configuration

506 The optimal borehole configuration can be taken from Fig. 6. The almost equal number of
507 heat pumps shown in Fig. 5 can be reached due to a decreasing amount of wells within a
508 developing higher heat flux. For higher heat loads there is a need of a higher number of wells,
509 whereupon the total well length linearly rises. A higher average outside temperature causes
510 only a slightly higher stepwise demand of a total well length.

511 5.3.3. Optimal COP

512 The optimal COP value is for a change of the heat flux for the investigated data range 7.31
513 with a standard deviation of approx. 0.061. For a change of the heat load amounts the value to
514 7.26 with and standard deviation of 0.083. And for a change of the average ground surface
515 temperature is the average COP value 7.04 and the standard deviation 0.581. For the variation
516 of the heat load is the COP increasing with a higher heat load (Fig. 7). This integer constraint

16
Retkowski W, Thöming J (2014) Applied Energy (114), 492-503

517 solution is comparative shown with calculated relaxation solutions and can be taken from Fig.
518 7. A bound was applied as shown in section 3.2.1.1 for a maximal single well depth and
519 created three different algorithm cases with the values 300 m, 900 m and 1500 m. For each
520 one were optimal design points related to a varying heat load calculated. The average COP
521 shown in Fig. 7A is related to Fig. 4B, 5B and 6B. There were 300 m applied, in Fig.7B and
522 Fig.7C 600 m and respectively 900 m applied. The GRG2 found best relaxation based
523 solutions near the bound of 300 m with increasing heat loads. For this reason also the COP
524 remains at a value of approx. 7.37 with a bound at 300 m. With a change of the upper bound
525 to a value of 900 m, the COP increases significant. The solver uses the higher solution domain

Accepted version
526 and finds solutions which vary close to an almost constant linear distribution around a COP of
527 8.75. Changing the maximal single length bound to a “virtual” value of 1500 m, the effect of a
528 higher variation in the solutions increases only slightly with a variation of heat load. The
529 optimal COP results for the heat load variation problem with integer constraints, without
530 integer constraints (due to relaxation) and with a variation of a single bound can be taken
531 from Fig. 7.

532 6. Conclusion and future work


533 A new method for design optimization of GSHPSs is presented. The proposed method
534 combines financial and thermodynamic aspects with the selection problem of real components
535 and delivers an optimal designed system.

536 It became obvious that a huge range of starting points within the design space led to converge
537 to a local minimum instead of the global minimum. However, the GRG2 found often
538 sufficient solutions in a time span under 1s. The EA found often solutions in reasonable time
539 below 360s when used as hybrid with the GRG2 applied in advance. Note that it might be
540 convenient to adjust narrow bounds for each design variable using the EA. In general requires
541 the GRG2 less computational effort than the EA. A comparison of different objectives
542 showed that different systems might be optimal, where in the studied case the straight
543 minimization of costs seemed to be as well ecological sufficient. The sensitivity of the costs,
544 well and heat pump arrangement related to a variation of three input parameter values (heat
545 load, heat flux and the average temperature) were analyzed. Results show that the investment
546 costs drop with increased thermal properties. Heat pump variations depend on the size and
547 quality of pumps included in the data base and the algorithm prefers for an optimal design of
548 industrial size problems lower units of heat pumps with a higher capacity. The heat load has a
549 strong impact especially on the configuration of heat pumps and wells. Comparing this
550 influence with the influence of the variation of the average ground temperature and heat flux
551 these variables have a less impact on the selection of heat pumps and wells. The peak-load
552 method seems to provide sufficient design solutions. All optimal thermo-economic
553 configurations had beneficial environmental balance. The heat load dependency related to
554 investment costs, operational costs, total well length and number of total heat pumps is even
555 under the constraints of integer values almost linear. Optimal results have shown that required
556 heat pump start-up energy can be a crucial design criterion. This can lead to different optimal
557 heat pump arrangements. Due to a change of local average ground surface temperatures it is
558 more suitable to change the conditions of the well arrangement in the ground than the

17
Retkowski W, Thöming J (2014) Applied Energy (114), 492-503

559 arrangement of heat pumps. Altogether delivers the proposed optimization model good
560 estimates for quick design investigations of real components to design an optimal vertical
561 GSHPS. Already Dickinson [42] pointed out that of course the optimization of GSHPSs
562 remains only part of the required methodology to reduce economic and CO2 cost of using
563 energy in buildings. Additionally should the building designer weight up the most cost
564 effective solution to result in the desired benefits.

565 Further investigations could extend the model with a heating circuit, the latest heat pump
566 technology or study the impact of a huge heat pump data base in detail. With an extended
567

Accepted version
time depended heat flux one could investigate possible limits of the related operational time.
568 As one next step the author will investigate and provide a quantitative global sensitivity
569 analysis to get a deep insight of the new optimization model and the related processes and
570 potentials.

571

572

18
Retkowski W, Thöming J (2014) Applied Energy (114), 492-503

573 Acknowledgements
574 The authors wish to acknowledge support from the Wirtschaftsförderung Bremen GmbH
575 (WFB Bremen – Economic Development) and the EUROPEAN UNION: Investing in your
576 future – European Regional Development Fund.

577

Accepted version

19
Retkowski W, Thöming J (2014) Applied Energy (114), 492-503

578 Appendix A
579 For completeness are the used mathematical coefficients and some further parameter values
580 characterizing investigated heat pumps provided. Also the model set-up parameter values
581 used for specific calculations are presented.

582 A.1. Heat pump data base functions and coefficients


583
584 The provided coefficients and parameter values given in Table A.4 and A.5 were determined
585 from manufacturer´s data sheets with the creation of linear regression lines for each specific

Accepted version
586 heat pump. The correlation coefficient was for each function above 0.99. The parameter
587 used in Eq. (8) was for all cases 0.8. The minimal flow rate in the soil circuit could be
588 estimated with 20% of the nominal flow rate.
589
590 Table 4
591 Properties for heat pumps 1-3 with a close spread capacity selection, valid for a mean heating
592 circuit inlet temperature of 35 °C.

Unit - - - - A m³ h-1
heat pump 1 0.261 7.900 -0.0013 1.63 14.8 5658 17.00 0.38
heat pump 2 0.344 10.907 0.0000 2.20 19.8 6212 20.00 0.52
heat pump 3 0.489 13.914 -0.0025 2.80 26.8 6693 23.00 0.68
593
594 Table 5
595 Properties for heat pumps 1-4 with a wide spread capacity selection, valid for a mean heating
596 circuit inlet temperature of 35 °C.

Unit - - - - A m³ h-1
heat pump 1 1.83 63.45 0.03 19.27 88.00 25359 63.00 0.78
heat pump 2 3.15 108.8 0.100 26.00 160.00 39451 80.00 1.36
heat pump 3 1.74 55.7 0.002 12.96 82.00 26785 64.00 0.70
heat pump 4 0.48 16.3 -0.002 3.65 24.50 5537 30.00 0.76
597
598 A.2. Model Set-up parameters
599
600 Further values for case studies shown in section 5.3. and for the cases 5.1 and 5.2 which are
601 not in the sections or Tab. 1 given are shown here in Table A.6.
602
603 Table 6
604 Input conditions.
Ground properties
Mean soil conductivity 2.0 W m-1 K-1
Borehole diameter 0.15 m
Grout thermal conductivity 1.0 W m-1 K-1
Geothermal gradient 0.06 W m-2
Hydraulic properties
Specific heat capacity 3800 J kg-1 K-1

20
Retkowski W, Thöming J (2014) Applied Energy (114), 492-503

PE pipe diameter 0.032 m


Thickness PE-pipe material 0.0037 m
Operating conditions
Full operation hours 2300 h y-1
Seasonal operation hours 4500 h y-1
HP work cycle 3 Unit h-1
Electrical to heat efficiency 0.85 -
Economic properties
Energy price 0.25 € kWh-1
Borehole drilling cost 60 € m-1

Accepted version
Additional connection drilling costs 30 % of IC
605

606 A.3. Fluid density


607
608 The fluid density depends strongly on the average fluid temperature as shown in Eq. (41). The
609 density for a binary water-propylenglycol mixture of 25% propylene glycol and 75% water is
610 approximated from manufacturer´s data sheets and can be calculated with Eq. (42) taken from
611 [43].
612
613 (41)
614
615 (42)
616

21
Retkowski W, Thöming J (2014) Applied Energy (114), 492-503

617
618 References
619
[1] Bayer P, Saner D, Bolay S, Rybach L, Blum P. Greenhouse gas emossion savings of
ground source heat pump systems in Europe. Renew Sust Energ Rev 2012; 16: 1256-
1267.
[2] Yang H, Cui P, Fang Z. Vertical-borehole ground-coupled heat pumps: A review of
models and systems. Appl Energy 2010; 87: 16-27.
[3] Do SL, Haberl JS. A review of ground coupled heat pump models used in whole-
building computer simulation programs. Proceedings 17th Symposium for improving

Accepted version
building systems in hot and humid climates. Energy Systems Laboratory, Austin Texas
2010.
[4] Javed S, Fahlén P, Claesson J. Vertical ground heat exchangers: A review of heat flow
models. Effstock 2009- Stockholm, Sweden, 2009-06-14--17. Proceedings vol. CD-
proceedings.
[5] Eskilson Per, Claesson J. Simulation model for thermally interacting heat extraction
boreholes, Numerical Heat Transfer, Part: A Applications 1988; (13-2): 149-165.
[6] Hellstrom G, Sanner B. PC-programs and modeling for borehole heat exchanger
design. In: Proc. intern. Geothermal days Germany, Gtv, Geeste, Supplement; 2001:
35-44.
[7] Yavuzturk C. Modelling of vertical ground loop heat exchangers for ground source
heat pump systems. Doctoral Thesis, Oklahoma State University, USA; 1999: 1-231.
[8] Zeng HY, Diao NR, Fang ZH. A finite line-source model for boreholes in geothermal
heat exchangers. Heat Transfer Asian Res 2002; 31(7):558–67.
[9] Lamarche L, Dupré G, Kajl S. A new design approach for Ground Source Heat Pumps
based on hourly load simulations, International Conference On Renewable Energies
And Power Quality (ICREPQ); 2007.
[10] Rusevljan M, Soldo V, Curko T. Optimal sizing of borehole heat exchangers.
Strojarstvo 2009; 51(5): 473-480.
[11] Wall G. Thermoeconomic optimisation of a heat pump system. Chalmers University of
Technology and University of Göteborg 1985:1-17.
[12] Marzbanrad J, Sharifzadegan A, Kahrobaeian A. Thermodynamic Optimization of
GSHPS Heat Exchangers. Int. J. of Thermodynamics 2007; 10 (3): 107-111.
[13] Sayyaadi H, Hadaddi E, Amidpour M. Multi-objective optimization of a vertical
ground source heat pump using evolutionary algorithm. Energy Convers Manage 2009;
(50): 2035-2046.
[14] Sanaye S, Niroomand B. Thermal-economic modeling and optimization of vertical
ground-coupled heat pump. Energy Convers Manage 2009; (50): 1136-1147.
[15] Li M, Lai ACK. Thermodynamic optimization of ground heat exchangers with single
U-tube by entropy generation minimization method. Energy Convers Manage 2013; 65:
133-139.
[16] Sayyadi H, Nejotolahi M. Thermodynamic and thermoeconomic optimization of a
cooling tower-assisted ground source heat pump. Geothermics 2011; 40: 221-232.
[17] Lee K-H, Lee D-W, Baek N-C, Kwon H-M, Lee C-J. Preliminary determination of
optimal size for renewable energy resource in buildings using RETScreen. Energy
2012; 47: 83-96.
[18] Rezvan AT, Gharneh NS, Gharehpetian GB. Robust optimization of distributed
generation investment in building. Energy 2012; 48: 455-463.
[19] Florides GA, Christodoulides P, Pouloupatis P, Single and double U-tube ground heat
exchangers in multiple-layer substrates. Appl Energy 2013; 102: 364-373.

22
Retkowski W, Thöming J (2014) Applied Energy (114), 492-503

[20] de Paly M, Hecht-Méndez J, Beck M, Blum P, Zell A, Bayer P. Optimization of energy


extraction for closed shallow geothermal systems using linear programming.
Geothermics; 43: 57-65.
[21] Beck M, Bayer P, de Paly M, Hecht-Méndez J, Zell A. Geometric arrangement and
operation mode adjustment in low-enthalpy geometrical borehole fields for heating.
Energy 2013; 49: 434-443.
[22] Michopoulos A, Kyriakis N. Predicting the fluid temperature at the exit of the vertical
heat exchangers. Appl Energy 2009; 86: 2065-2070.
[23] Hepbasli A, Akdemir O. Energy and exergy analysis of a ground source (geothermal)
heat pump syste. Energy & Conversion Manage 2004; 45: 737-753.

Accepted version
[24] Hepbasli A. A key review on exergetic analysis and assesment of renewable energy
resources for a sustainable future. Renewable and sustainable energy reviews 2008; 12:
593-661.
[25] Bi Y, Wang X, Liu Y, Zhang H, Chen L. Comprehensive exergy analysis of a
ground/source heat pump system for both building heating and cooling modes. Appl
Energy 2009; 86:2560-2565.
[26] Torío H, Angelotti A, Schmidt D. Exergy analysis of renewable energy-based
climatisation systems for buildings: A critical review. Energy and Buildings 2009; 41:
248-271.
[27] Shonder JD, Martin M, Hughes P, Thornton J. Geothermal Heat Pumps in K-12
Schools – A case study of Lincoln, Nebraska, Schools. ORNL/TM-2000/80, Oak Ridge
National Laboratory; 2000.
[28] Lee CK. Effects of multiple ground layers on thermal response test analysis and
ground/source heat pump simulation. Appl Energy 2011; 88: 4405-4410.
[29] Florides GA, Christodoulides P, Pouloupatis P. An analyis of heat flow through a
borehole heat exchanger valideted model. Appl Energy 2012; 92: 523-533.
[30] Lamarche L, Kajl S, Beauchamp B. A review of methods to evaluate borehole thermal
resistances in geothermal heat-pump systems. Geothermics 2010; 39: 187-200.
[31] Bennet J, Claesson J, Hellstrom G. Multipole Method to Compute the Conductive Heat
Transfer to and between Pipes in a Composite Cylinder. Notes on Heat Transfer 3-
1987. Department of Building Physics, Lund Institute of Technology, Lund, Sweden
1987: 1-42.
[32] Liu X, Hellstrom G. Enhancements of an integrated simulation tool for ground-source
heat pump system design and energy analysis. Proceedings of Ecostock 2006, the 10th
International Conference on thermal energy storage, the Richard Stockton College of
New Jersey.
[33] Hellstrom G, Sanner B. EED – Earth Energy Designer, Version 1.0, User’s Manual.
Prof. Dr. Knoblich & Partner GmbH, Wetzlar, Germany, 1997: 43.
[34] Spitler JD. GLHEPRO – A design tool for commercial building ground loop heat
exchanger. In: Proceedings of the 4th International Conference on Heat Pumps in Cold
Climates, 17–18 August, Aylmer, QC, Canada; 2000: 1–16.
[35] VDI 4640 Blatt 2 - Verein Deutscher Ingenieure (VDI) [Hrsg.]. Thermische Nutzung
des Untergrundes – Erdgekoppelte Wärmepumpenanlagen. Düsseldorf, 2001: 12.
[36] Lasdon L, Waren A. GRG2 User's Guide. Department of Management Science and
Information Systems, The University of Texas at Austin. 2 October 1997: 1-50.
[37] Bernard O, Alata O, Francaux M. On the modelling of breath-by-breath oxygen uptake
kinetics at the onset of high-intensity exercises: simulated anneling vs. GRG2 method, J
Appl Physiol 2005; 100: 1049-1058.
[38] Hyde KM, Maier HR. Distance-based and stochastic uncertainty analysis for multi-
criteria decision analysis in Excel using Visual Basic for Applications, Environ Modell
& Soft 2006; 21: 1695-1710.

23
Retkowski W, Thöming J (2014) Applied Energy (114), 492-503

[39] Fylstra D, Lasdon L, Watson J. Waren, A., Design and use of the Microsoft Excel
Solver. Interfaces, Vol. 28, No 5, Sept-Oct. 1998: 29-55.
[40] Frontline Systems, Frontline Solvers – User Guide. Version 11.5. For use with Excel
2003-2010. Frontline Systems, Inc., Incline Village, Nevada, 2011.
[41] Eberhard M. Erdwärmesondenfeld Aarau. Heizen und Kühlen eines grossen
Bürogebäudes mit teilweise wärmeisolierten Erdwärmesonden. Bundesamt für Energie.
Schlussbericht Oktober 2005. Aarau, 2005.
[42] Dickinson J, Jackson T, Matthews M, Cripps A. The economic and environmental
optimisation of integrating source energy system into buildings. Energy 2009; 34:
2215-2222.

Accepted version
[43] Glück B. Simulationsmodell Erdwärmesonden zur wärmetechnischen Beurteilung von
Wärmequellen, Wärmesenken und Wärme-/Kältespeichern. Bericht der RUD. OTTO
MEYER – Umwelt – Stiftung, Hamburg, 2007.
[44] Eskilson P, Cleasson J, Blomberg T, Sanner B. Earth Energy Designer. Version 2.0
(Dec 19,1999).
620
621
622
623

24
Retkowski W, Thöming J (2014) Applied Energy (114), 492-503

624

625
626
627
compressor
628
629
Tair soil rotary pump
630 Tf2,i

Accepted version
631 Qhp.i
Qsc,2
632 GSHE heat pump
633 cycle evaporator cycle
634
Qhc.load
Qsc, 3
635 Tgs binary mixture Tf1,i
636
637
expansion
638
Qsc,1 ground-scource valve
639
heat exchanger
640
Tbi
Li

641
642
643
644
645
646
647
648 Fig.1. GSHP system flow with indicated modeled action.
649
650
651
652
653
654
655
656
657
658
659
660
661

25
Retkowski W, Thöming J (2014) Applied Energy (114), 492-503

662
663
664
665
666 Tb

667 R1 R2 xc

668 Tup Tdn

669 R12 rb rp

Accepted version
670
671
672 Fig.2. Thermal resistance circuit with one U-tube (left) and Parameters (right) related to Eqs. (21)-(23).
673
674
675
676
677
678
679
680
681
682
683
684
685
686
687
688
689
690
691
692
693
694
695
696
697
698
699

26
737
736
735
734
733
732
731
730
729
728
727
726
725
724
723
722
721
720
719
718
717
716
715
714
713
712
711
710
709
708
707
706
705
704
703
702
701
700
Retkowski W, Thöming J (2014)

Fig. 3. Measured versus modeled mean fluid temperature Tf (D) and total electrical energy demand Ptot (C)
considered under equal input conditions (A-B). Relative error is in both cases (C-D) less than 1%.

27
Applied Energy (114), 492-503

Accepted version
Retkowski W, Thöming J (2014) Applied Energy (114), 492-503

738
739
740
741
742
743

Fig. 4. Investment and operating costs of different GSHP design points due to variations of heat flux, heat load and the
744
745

Accepted version
746
747
748
749
750
751
752
753
754
755
756
757
758
759
760
761
average annual temperature.

762
763
764
765
766
767
768
769
770
771
772
773
774
775

28
Retkowski W, Thöming J (2014) Applied Energy (114), 492-503

776
777
778
779
780
781

Fig. 5. Optimal heat pump configurations of different GSHP design points due to variations of heat flux, heat load and
782
783

Accepted version
784
785
786
787
788
789
790
791
792
793
794
795
796
797
798
799
800
the average annual temperature.

801
802
803
804
805
806
807
808
809
810
811
812
813

29
Retkowski W, Thöming J (2014) Applied Energy (114), 492-503

814
815
816
817
818
819
820

Fig. 6. Optimal well configurations of different GSHP design points due to variations of heat flux, heat load and the average
821

Accepted version
822
823
824
825
826
827
828
829
830
831
832
833
834
835
836
837
838
839
840
841
842
843
annual temperature.

844
845
846
847
848
849
850
851

30
887
886
885
884
883
882
881
880
879
878
877
876
875
874
873
872
871
870
869
868
867
866
865
864
863
862
861
860
859
858
857
856
855
854
853
852
Retkowski W, Thöming J (2014)

31
Fig. 7. Optimal COP design points of varied heat load conditions with integer constraints and without integer
constraints (relaxation). A: Single well length was bounded with max. 300 m. B: Single well length was
bounded with max. 900 m. C: Single well length was bounded with 1500 m.
Applied Energy (114), 492-503

Accepted version

You might also like