Anda di halaman 1dari 232

Computer modelling and simulation of geothermal heat

Title pump and ground-coupled liquid desiccant air


conditioning systems in sub-tropicalregions

Author(s) Lee, Chun-kwong.; .

Citation

Issue Date 2008

URL http://hdl.handle.net/10722/51910

The author retains all proprietary rights, (such as patent


Rights rights) and the right to use in future works.
Abstract of thesis entitled

Computer Modelling and Simulation of Geothermal Heat Pump and


Ground-coupled Liquid Desiccant Air Conditioning Systems in Sub-
tropical Regions
Submitted by

Lee Chun Kwong


for the degree of Doctor of Philosophy
at the University of Hong Kong
in November 2008

A three-dimensional finite difference model was developed for the ground


heat exchanger borefield of geothermal heat pump systems with groundwater
advection. Rectangular coordinate system was employed with the boreholes
approximated by square columns circumscribed by the borehole radius. The
gridding scheme was calibrated against the cylindrical source model with the
inclusion of a load factor equal to 1.047. Variations in the temperature and the
loading along the boreholes were taken into account. The simulated performance for
a single borehole showed that there was little difference when compared with the
analytical model, but the deviation increased with the scale of the borefield.

Using the new model, the effect of groundwater direction on borefield


performance in terms of the percentage variation (PV), was investigated. For square
borefields, PV was negligible for a small groundwater flow with Darcy velocity less
than 10-6m/s, and generally increasing with the groundwater flow magnitude. For
non-square borefields, PV reached a maximum at a particular groundwater flow.
Irrespective of the borefield type, PV increased when the borefield size was larger.
It could reach 5% for a 5x5 borefield, 12% for 5x10 borefield and 10% for 6x9
borefield based on a constant load with a daily operating schedule of 12 hours.
Substantial reduction in the borehole length could thus be made with correct setting
of the borefield orientation.

The new model was employed in thermal response test analysis using
parameter estimation technique to determine the groundwater velocity. It was found
that the estimated groundwater velocity scattered at repeated trials with test data
generated using a low groundwater velocity. At higher groundwater flows exceeding
2x10-7m/s, the groundwater velocity could be estimated with more confidence. For a
longer test period of 30 days, confident results could be obtained with groundwater
velocity down to 10-7m/s. Experimental thermal response tests were performed and
the results were analysed using the present model. Large errors in the estimated
groundwater velocities were found due to several reasons. The range of the
estimated soil thermal conductivity was lower than the maximum value determined
in laboratory soil thermal conductivity measurement test.

A new hybrid system of ground-coupled liquid desiccant air conditioner was


developed in which the fresh air was treated by the liquid desiccant cycle in cooling
mode. Dynamic simulation was carried out for a single-zone sample building at
different occupancy levels based on the weather data for Hong Kong from 1986 to
1995. Comparison of the simulation results with those obtained using a conventional
geothermal heat pump system showed that the borehole length was reduced by
10.11% on average under different groundwater velocities at a low occupancy level
corresponding to a fresh air ratio of 0.066. A larger average reduction of 14.31%
could be reached for a higher occupancy level corresponding to a fresh air ratio of
0.122. For both the new and conventional systems, the energy input to the
compressor was nearly the same. Thus, the new system was as energy-efficient as
the conventional geothermal heat pump system.
Computer Modelling and Simulation of Geothermal Heat Pump and
Ground-coupled Liquid Desiccant Air Conditioning Systems in Sub-
tropical Regions

by

Lee Chun Kwong


B.Sc.(Eng), M.Ph. HKU

A thesis submitted in partial fulfillment of the requirements for


the Degree of Doctor of Philosophy
at the University of Hong Kong

November 2008
Declaration

I declare that this thesis rpresentsmy ovn work, except wherc due
is made,andthat it hasnot beenpreviouslyincludedin a thsis,
acknowledgement
dissertationor rcport suhnittedto this Univrsityor to any otherinstitutionfor a
degree,diplomaor otherMifications.
ii
Acknowledgements

I would like to thank my supervisor, Mr. H. N. Lam, for providing strong


support and good guidance during my three years of study. He provided important
information for my fulfillment of the course requirement, and valuable advice on my
future career development after I graduated.

I would also like to thank Mr. W. K. Leung and Mr. C. C. Chan, the
technicians in Yam Pat Building, for helping me to arrange the laboratory work place
and necessary testing equipment. Their prompt response enabled me to carry out the
experimental work smoothly. My special thank went to Mr. William Lee who helped
me very much in setting up and modification of the test rig.

I would like to express my appreciation for Ms. Candy He who assisted me to


get acquainted with the TRNSYS software. Our cooperation and discussion helped
me to understand the various aspects of geothermal heat pump systems. The help
from Mr. Richard Tan was acknowledged for introducing the principle of liquid
desiccant system and presenting his liquid desiccant dehumidifier model to me.

I would also like to express my great gratitude to Dr. Daniel Pahud of SUPSI
and Professor Jeffrey D. Spitler of Oklahoma State University in providing me the
site thermal response test data. Dr. Pahud strongly supported me in understanding
the application of the TRNVDSTP TRNSYS module for simulation. Professor
Spitler gave valuable advice to me when we met in the conference at Beijing in 2007.

Finally, I would like to thank my parents who remained patient during my


long absence from work. Their tolerance and understanding supported my pursuance
of the study.

iii
Table of Contents

Declaration i
Acknowledgement iii
Table of Contents iv
List of Figures x
List of Tables xxi
List of Abbreviations xxvii
Nomenclature xxviii

Chapter 1 : Introduction 1
1.1 Geothermal Heat Pump Systems 1
1.2 Liquid Desiccant Dehumidification Systems 4
1.3 Objective of This Research 6

Chapter 2 : Literature Review 8


2.1 Geothermal Heat Pump Systems 8
2.1.1 General 8
2.1.2 Modelling of Vertical Ground Heat Exchanger 11
Performance
2.1.2.1 Modelling without Groundwater Effect 13
2.1.2.2 Modelling with Groundwater Effect 21
2.1.3 Thermal Response Tests Analysis 25
2.2 Liquid Desiccant Dehumidification Systems 28
2.2.1 General 28
2.2.2 Modelling of Liquid Desiccant Dehumidification 30
Systems
2.3 Vapour Compression Systems 33

Chapter 3 : Mathematical Formulations for Geothermal and Liquid 36


Desiccant Systems

iv
3.1 Ground Heat Exchanger Borefield 36
3.1.1 Heat Transfer Around Borehole 36
3.1.2 Heat Transfer Inside Borehole 38
3.1.2.1 Flow Process 39
3.1.2.2 No Flow Process 41
3.1.3 Solution Methodology 42
3.1.4 Borefield Performance Formulation 43
3.1.5 Thermal Response Test Analysis 44
3.2 Liquid Desiccant Systems 45
3.2.1 Desiccant Packed Tower 45
3.2.2 Desiccant-to-desiccant Heat Exchanger 47
3.2.3 Desiccant Heater/Cooler 47
3.3 Vapour Compression Air Conditioning Systems 48
3.3.1 Compressor 49
3.3.2 Condenser or Desiccant Heater 49
3.3.3 Expansion Device 52
3.3.4 Evaporator or Desiccant Cooler 52
3.3.5 Air-to-air Heat Exchanger 54
3.3.6 Air Heater 55
3.3.7 Ground-coupled Coil 55
3.3.8 Air Cooler 55
3.3.8.1 Approximate Iterative Method 56
3.3.8.2 Finite Difference Method 57
3.3.9 Single Zone 57

Chapter 4 : Laboratory Setup for Experimental Thermal Response 59


Tests
4.1 Thermal Response Test Rig 59
4.2 Soil Thermal Conductivity Measurement Setup 61
4.3 Calibrating Equipment 62

Chapter 5 : Methodologies for Simulation and Analysis 66


5.1 Study of Ground Heat Exchanger Borefield Performance 66

v
5.2 Thermal Response Test Analysis 68
5.3 Modelling of Ground-coupled Liquid Desiccant Air 69
Conditioners

Chapter 6 : Results and Discussions 73


6.1 Study of Ground Heat Exchanger Borefield Performance 73
6.1.1 Setting of Ground Grids and Load Factor 73
6.1.2 Performance of Borefield without Groundwater 76
6.1.2.1 Temperature and Loading Profile along 76
Single Borehole
6.1.2.2 Ground Temperature Profiles around a 77
Single Borehole
6.1.2.3 Comparison with the Finite Line Source 78
Model
6.1.2.4 Borefield Thermal Resistance Variation 79
6.1.2.5 Borehole Performance with Time Schedule 80
6.1.3 Performance of a Single Borehole with 81
Groundwater Advection
6.1.3.1 Leaving Fluid Temperature Rise of Single 81
Borehole at Different Darcy Velocities of
Groundwater Flow
6.1.3.2 Temperature and Loading Variation along a 82
Single Borehole at Different Darcy
Velocities of Groundwater Flow
6.1.3.3 Performance of a Single Borehole at 82
Different Loading Profiles
6.1.4 Effect of Groundwater Velocity on Borefield 84
Design Length at Different Load Profiles
6.1.5 Effect of Borehole Length on Performance of 85
Borefield at Different Groundwater Velocities and
Load Profiles
6.1.6 Performance of Borefield at Different Groundwater 88
Directions

vi
6.1.6.1 Variation of PV for Square Borefields 88
6.1.6.2 Variation of PV for Non- square Borefields 89
6.1.6.3 Variation of PV for Different Borefield 90
Scales
6.2 Thermal Response Test Analysis 91
6.2.1 Finalisation of Parameters to be Determined 91
6.2.2 Analysis Based on the Generated TRT data with 93
Specified Groundwater Velocity
6.2.2.1 Analysis Based on a Normal Test Period 94
6.2.2.2 Analysis Based on a Long Test Period 96
6.2.3 Analysis Based on the Generated TRT data 98
Including Groundwater Velocity in the Parameter
Estimation Scheme
6.2.3.1 Analysis Based on a Normal Test Period 98
6.2.3.2 Analysis Based on a Long Test Period 99
6.2.3.3 Effects of Other Parameters on the 100
Confident Minimum of Estimated
Groundwater Velocity
6.2.3.4 Effects of Far Field Temperature and 101
Borefield Load on the Estimated Parameters
6.2.4 Analysis of Experimental Thermal Response Test 103
Results
6.2.4.1 Parameter Estimation Results of 104
Experimental TRT
6.2.4.2 Laboratory Soil Thermal Conductivity 105
Measurement Test Results
6.2.4.3 Source of Experimental Errors 106
6.2.5 Analysis of Site Thermal Response Test Results 108
Provided by Others
6.3 Modelling of Ground-coupled Liquid Desiccant Air 112
Conditioners
6.3.1 Determination of Final Configuration of Liquid 112
Desiccant Dehumidifying Cycle

vii
6.3.1.1 Effect of Desiccant Cooler on Performance 113
of Liquid Desiccant Dehumidifier
6.3.1.2 Effect of Regenerative Air Heater on 113
Performance of Liquid Desiccant
Dehumidifier
6.3.1.3 Effect of Regenerative Air Heat Exchanger 114
on Performance of Liquid Desiccant
Dehumidifier
6.3.2 Performance of Heat-pump-coupled Liquid 115
Desiccant Dehumidifier (HPCLDD)
6.3.3 Analysis of Ground-coupled Liquid Desiccant Air 118
Conditioner (GCLDAC) and Comparison with
Conventional Ground Source Heat Pump (GSHP)
6.3.3.1 Ultimate Configuration of GCLDAC 118
6.3.3.2 Determination of Design Duty of 119
GCLDAC/GSHP
6.3.3.3 Setting of Design Parameters of 121
GCLDAC/GSHP
6.3.3.4 Performance of GCLDAC/GSHP at Design 123
Conditions
6.3.3.5 Dynamic Simulation of GCLDAC/GSHP at 124
No Groundwater Flow
6.3.3.6 Dynamic Simulation of GCLDAC/GSHP 128
with Groundwater Flow

Chapter 7 : Conclusions and Recommendations 130


7.1 Conclusions 130
7.2 Recommendations for Further Research Work 133

Appendices 135
A. Transformation of Thermal Interference Coefficients 135
Between Tubes Inside a Borehole
B. Algorithm for Calculating Condenser Coil Performance 137

viii
C. Determination of Mean Refrigerant Density along 139
Saturated Region in Counter-flow Condenser/Evaporator
D. Algorithm for Calculating Evaporator Coil Performance 141
E. Algorithm for Calculating Air Cooler Performance Based 143
on Approximate Iterative Method
F. Modified Numerical Integration Method Used for FLSM 145
G. Modified Numerical Integration Method Used for LSMGA 146
H. Borehole Thermal Interference Based on FLSMGA 147
I. Calibration Curves for Measuring Instruments Used in 150
Laboratory Tests
J. Detailed Performance Curves at Different Groundwater 152
Directions and Load Profiles for Various Borefields
K. Parameter Estimation Results with Specified Groundwater 163
Flow Using LSMGA on Generated TRT Data for 5 Days at
Different Groundwater Velocities
L. Parameter Estimation Results with Specified Groundwater 166
Flow Using LSMGA on Generated TRT Data for 30 Days
at Different Groundwater Velocities
M. Parameter (Including Groundwater Velocity) Estimation 169
Results Using LSMGA on Generated TRT Data Using
LSMGA for 5 Days at Different Groundwater Velocities
N. Parameter (Including Groundwater Velocity) Estimation 171
Results Using LSMGA on Generated TRT Data Using
LSMGA for 30 Days at Different Groundwater Velocities
O. Logged Temperature Profiles in Experimental Thermal 173
Response Tests
P. Dynamic Simulation Results of GCLDAC/GSHP with 175
Groundwater Flow

References 177

ix
List of Figures

Figure 1.1 Schematic diagram of typical geothermal heat pump 2


system
Figure 1.2 Layout of a simple open type borefield configuration 3
Figure 1.3 Details of vertical ground heat exchanger borefield 4
Figure 1.4 Basic arrangement of liquid desiccant dehumidifier 5
Figure 2.1 Transformation of load profile to series of step loads 12
Figure 2.2 Electrical analogy of heat transfer between borehole 13
circulating fluid and surrounding ground
Figure 3.1 Discretisation scheme for the ground surrounding a 38
borehole
Figure 3.2 Discretisation scheme inside borehole 38
Figure 3.3 Discretisation scheme for a packed tower 45
Figure 3.4 Notations used for vapour compression cycle modelling 48
Figure 3.5 Change of state of refrigerant and condensing fluid 50
across condenser
Figure 3.6 Change of state of refrigerant and evaporating fluid 52
across evaporator
Figure 3.7 Discretisation scheme for air-to-air heat exchanger 54
Figure 3.8 Change of state of refrigerant and air across air cooler 56
Figure 3.9 Single zone with air conditioning 58
Figure 4.1 Schematic diagram of thermal response test rig 60
Figure 4.2 Layout of thermal response test setup 60
Figure 4.3 Details of experimental setup for measuring soil thermal 61
conductivity
Figure 4.4 Layout of experimental setup for soil conductivity 62
measurement
Figure 4.5 Temperature calibrator 63
Figure 4.6 Schematic diagram of calibration setup of voltmeter and 64
ammeter
Figure 4.7 Layout of calibration setup for voltmeter and ammeter: 64

x
(a) front view, (b) rear view
Figure 6.1 Temperature rise along single borehole without 77
groundwater flow at different times
Figure 6.2 Loading along single borehole without groundwater 77
flow at different times
Figure 6.3 Comparison of fluid and borehole temperature rise in 77
single borehole without groundwater flow after 1 year
Figure 6.4 Ground temperature profiles around single borehole 78
without groundwater flow after 5 years
Figure 6.5 Ground temperature profiles around single borehole 78
without groundwater flow after 10 years
Figure 6.6 Variation of borefield load during a daily operation 81
schedule of 12 hours for single borehole without
groundwater flow
Figure 6.7 Variation of leaving fluid temperature rise of single 81
borehole at different Darcy velocities of groundwater flow
Figure 6.8 Temperature rise along single borehole after 5 years at 82
various groundwater velocities
Figure 6.9 Temperature rise along single borehole after 10 years at 82
various groundwater velocities
Figure 6.10 Loading along single borehole after 5 years at various 82
groundwater velocities
Figure 6.11 Loading along single borehole after10 years at various 82
groundwater velocities
Figure 6.12 Performance of single borehole at no groundwater flow 83
for different load profiles
Figure 6.13 Performance of single borehole at groundwater velocity 83
-7
10 m/s for different load profiles
Figure 6.14 Performance of single borehole at groundwater velocity 83
2x10-7m/s for different load profiles
Figure 6.15 Performance of single borehole after 1 year at high 83
groundwater velocity for different load profiles
Figure 6.16 Variation of borehole length ratio with groundwater 84

xi
velocity at different load profiles with fixed leaving fluid
temperature rise for single borehole
Figure 6.17 Variation of borehole capacity ratio with groundwater 84
velocity at different load profiles with fixed leaving fluid
temperature rise for single borehole
Figure 6.18 Variation of borehole length ratio with groundwater 85
velocity at different load profiles with fixed leaving fluid
temperature rise for2x2 borefield
Figure 6.19 Variation of borehole capacity ratio with groundwater 85
velocity at different load profiles with fixed leaving fluid
temperature rise for 2x2 borefield
Figure 6.20 Variation of leaving fluid temperature rise ratio of 86
single borehole with short borehole length at different
groundwater velocities and load profiles
Figure 6.21 Variation of leaving fluid temperature rise ratio of 86
single borehole with long borehole length at different
groundwater velocities and load profiles
Figure 6.22 Variation of leaving fluid temperature rise ratio of 2x2 86
borefield with short borehole length at different groundwater
velocities and load profiles
Figure 6.23 Variation of leaving fluid temperature rise ratio of 2x2 86
borefield with long borehole length at different groundwater
velocities and load profiles
Figure 6.24 Variation of leaving fluid temperature rise ratio with 87
short borehole length at different groundwater velocities and
borefields under load profile 3
Figure 6.25 Variation of leaving fluid temperature rise ratio with 87
long borehole length at different groundwater velocities and
borefields under load profile 3
Figure 6.26 Variation of leaving fluid temperature rise ratio of 2x2 88
borefield with short borehole length at different groundwater
velocities and load profiles based on fixed leaving fluid
temperature rise at default borehole length

xii
Figure 6.27 Variation of leaving fluid temperature rise ratio of 2x2 88
borefield with long borehole length at different groundwater
velocities and load profiles based on fixed leaving fluid
temperature rise at default borehole length
Figure 6.28 Variation of PV with groundwater velocity for 2x2 89
borefield
Figure 6.29 Variation of PV with groundwater velocity for 3x3 89
borefield
Figure 6.30 Variation of PV with groundwater velocity for 1x2 89
borefield
Figure 6.31 Variation of PV with groundwater velocity for 1x3 89
borefield
Figure 6.32 PV of various borefields with groundwater velocity of 90
10-5m/s at constant load with a daily schedule
Figure 6.33 Sensitivity of ground volumetric heat capacity on 92
borehole performance at different groundwater velocity in
TRT
Figure 6.34 Sensitivity of ground thermal conductivity on borehole 92
performance at different groundwater velocity in TRT
Figure 6.35 Sensitivity of borehole thermal resistance on borehole 92
performance at different groundwater velocity in TRT
Figure 6.36 Sensitivity of groundwater volumetric heat capacity on 92
borehole performance at different groundwater velocity in
TRT
Figure 6.37 Performance of single borehole in TRT at different 93
groundwater velocity
Figure 6.38 Variation of applied load with temperature difference 106
across soil in laboratory soil thermal conductivity
measurement test
Figure 6.39 Schematic of advanced liquid desiccant dehumidifier 112
Figure 6.40 Schematic diagram of heat-pump-coupled liquid 115
desiccant dehumidifier
Figure 6.41 Outdoor temperature profile for 1986 116

xiii
Figure 6.42 Performance of ACFAU for 1986 116
Figure 6.43 Performance of HPCLDD with dp=0.02 for 1986 117
Figure 6.44 Performance of HPCLDD with dp=0.015 for 1986 117
Figure 6.45 Schematic diagram of ground-coupled liquid desiccant 119
air conditioner (cooling mode)
Figure 6.46 Building load profile for Case 1 120
Figure 6.46 Building load profile for Case 2 120
Figure 6.47 Performance of GSHP at no groundwater flow for Case 125
1 within 1986~1995
Figure 6.48 Performance of GCLDAC at no groundwater flow for 126
Case 1 within 1986~1995
Figure 6.49 Performance of GSHP at no groundwater flow for Case 128
2 within 1986~1995
Figure 6.50 Required borehole lengths of GCLDAC/GSHP at 129
different groundwater velocities for Case 1
Figure 6.51 Required borehole lengths of GCLDAC/GSHP at 129
different groundwater velocities for Case 2
Figure F.1 Variation of integrand of Eq. (2.33) with s 145
Figure I.1 Calibration curve for borehole temperature logger 150
Figure I.2 Calibration curves for temperature log buttons 150
Figure I.3 Calibration of thermocouples for soil thermal 151
conductivity measurement
Figure I.4 Calibration curve for voltmeter 151
Figure I.5 Calibration curve for ammeter 151
Figure J.1 Performance of 2x2 borefield with groundwater velocity 152
10-7m/s at different groundwater directions under load
profile 1
Figure J.2 Performance of 2x2 borefield with groundwater velocity 152
-7
2x10 m/s at different groundwater directions under load
profile 1
Figure J.3 Performance of 2x2 borefield with groundwater velocity 152
5x10-7m/s at different groundwater directions under load
profile 1

xiv
Figure J.4 Performance of 2x2 borefield with high groundwater 152
velocity after 2 years at different groundwater directions
under load profile 1
Figure J.5 Performance of 2x2 borefield with groundwater velocity 152
-7
10 m/s at different groundwater directions under load
profile 2
Figure J.6 Performance of 2x2 borefield with groundwater velocity 152
2x10-7m/s at different groundwater directions under load
profile 2
Figure J.7 Performance of 2x2 borefield with groundwater velocity 153
-7
5x10 m/s at different groundwater directions under load
profile 2
Figure J.8 Performance of 2x2 borefield with high groundwater 153
velocity after 2 years at different groundwater directions
under load profile 2
Figure J.9 Performance of 2x2 borefield with groundwater velocity 153
-7
10 m/s at different groundwater directions under load
profile 3
Figure J.10 Performance of 2x2 borefield with groundwater velocity 153
2x10-7m/s at different groundwater directions under load
profile 3
Figure J.11 Performance of 2x2 borefield with groundwater velocity 153
-7
5x10 m/s at different groundwater directions under load
profile 3
Figure J.12 Performance of 2x2 borefield with high groundwater 153
velocity after 2 years at different groundwater directions
under load profile 3
Figure J.13 Performance of 2x2 borefield with groundwater velocity 154
-7
10 m/s at different groundwater directions under load
profile 4
Figure J.14 Performance of 2x2 borefield with groundwater velocity 154
2x10-7m/s at different groundwater directions under load
profile 4

xv
Figure J.15 Performance of 2x2 borefield with groundwater velocity 154
5x10-7m/s at different groundwater directions under load
profile 4
Figure J.16 Performance of 2x2 borefield with high groundwater 154
velocity after 2 years at different groundwater directions
under load profile 4
Figure J.17 Performance of 3x3 borefield with groundwater velocity 154
10-7m/s at different groundwater directions under load
profile 1
Figure J.18 Performance of 3x3 borefield with groundwater velocity 154
-7
2x10 m/s at different groundwater directions under load
profile 1
Figure J.19 Performance of 3x3 borefield with groundwater velocity 155
5x10-7m/s at different groundwater directions under load
profile 1
Figure J.20 Performance of 3x3 borefield with high groundwater 155
velocity after 2 years at different groundwater directions
under load profile 1
Figure J.21 Performance of 3x3 borefield with groundwater velocity 155
10-7m/s at different groundwater directions under load
profile 2
Figure J.22 Performance of 3x3 borefield with groundwater velocity 155
-7
2x10 m/s at different groundwater directions under load
profile 2
Figure J.23 Performance of 3x3 borefield with groundwater velocity 155
5x10-7m/s at different groundwater directions under load
profile 2
Figure J.24 Performance of 3x3 borefield with high groundwater 155
velocity after 2 years at different groundwater directions
under load profile 2
Figure J.25 Performance of 3x3 borefield with groundwater velocity 156
10-7m/s at different groundwater directions under load
profile 3

xvi
Figure J.26 Performance of 3x3 borefield with groundwater velocity 156
2x10-7m/s at different groundwater directions under load
profile 3
Figure J.27 Performance of 3x3 borefield with groundwater velocity 156
-7
5x10 m/s at different groundwater directions under load
profile 3
Figure J.28 Performance of 3x3 borefield with high groundwater 156
velocity after 2 years at different groundwater directions
under load profile 3
Figure J.29 Performance of 3x3 borefield with groundwater velocity 156
-7
10 m/s at different groundwater directions under load
profile 4
Figure J.30 Performance of 3x3 borefield with groundwater velocity 156
2x10-7m/s at different groundwater directions under load
profile 4
Figure J.31 Performance of 3x3 borefield with groundwater velocity 157
-7
5x10 m/s at different groundwater directions under load
profile 4
Figure J.32 Performance of 3x3 borefield with high groundwater 157
velocity after 2 years at different groundwater directions
under load profile 4
Figure J.33 Performance of 1x2 borefield with groundwater velocity 157
-7
10 m/s at different groundwater directions under load
profile 1
Figure J.34 Performance of 1x2 borefield with groundwater velocity 157
2x10-7m/s at different groundwater directions under load
profile 1
Figure J.35 Performance of 1x2 borefield with groundwater velocity 157
-7
5x10 m/s at different groundwater directions under load
profile 1
Figure J.36 Performance of 1x2 borefield with high groundwater 157
velocity after 2 years at different groundwater directions
under load profile 1

xvii
Figure J.37 Performance of 1x2 borefield with groundwater velocity 158
10-7m/s at different groundwater directions under load
profile 2
Figure J.38 Performance of 1x2 borefield with groundwater velocity 158
-7
2x10 m/s at different groundwater directions under load
profile 2
Figure J.39 Performance of 1x2 borefield with groundwater velocity 158
5x10-7m/s at different groundwater directions under load
profile 2
Figure J.40 Performance of 1x2 borefield with high groundwater 158
velocity after 2 years at different groundwater directions
under load profile 2
Figure J.41 Performance of 1x2 borefield with groundwater velocity 158
10-7m/s at different groundwater directions under load
profile 3
Figure J.42 Performance of 1x2 borefield with groundwater velocity 158
-7
2x10 m/s at different groundwater directions under load
profile 3
Figure J.43 Performance of 1x2 borefield with groundwater velocity 159
5x10-7m/s at different groundwater directions under load
profile 3
Figure J.44 Performance of 1x2 borefield with high groundwater 159
velocity after 2 years at different groundwater directions
under load profile 3
Figure J.45 Performance of 1x2 borefield with groundwater velocity 159
10-7m/s at different groundwater directions under load
profile 4
Figure J.46 Performance of 1x2 borefield with groundwater velocity 159
-7
2x10 m/s at different groundwater directions under load
profile 4
Figure J.47 Performance of 1x2 borefield with groundwater velocity 159
5x10-7m/s at different groundwater directions under load
profile 4

xviii
Figure J.48 Performance of 1x2 borefield with high groundwater 159
velocity after 2 years at different groundwater directions
under load profile 4
Figure J.49 Performance of 1x3 borefield with groundwater velocity 160
-7
10 m/s at different groundwater directions under load
profile 1
Figure J.50 Performance of 1x3 borefield with groundwater velocity 160
2x10-7m/s at different groundwater directions under load
profile 1
Figure J.51 Performance of 1x3 borefield with groundwater velocity 160
-7
5x10 m/s at different groundwater directions under load
profile 1
Figure J.52 Performance of 1x3 borefield with high groundwater 160
velocity after 2 years at different groundwater directions
under load profile 1
Figure J.53 Performance of 1x3 borefield with groundwater velocity 160
-7
10 m/s at different groundwater directions under load
profile 2
Figure J.54 Performance of 1x3 borefield with groundwater velocity 160
2x10-7m/s at different groundwater directions under load
profile 2
Figure J.55 Performance of 1x3 borefield with groundwater velocity 161
-7
5x10 m/s at different groundwater directions under load
profile 2
Figure J.56 Performance of 1x3 borefield with high groundwater 161
velocity after 2 years at different groundwater directions
under load profile 2
Figure J.57 Performance of 1x3 borefield with groundwater velocity 161
-7
10 m/s at different groundwater directions under load
profile 3
Figure J.58 Performance of 1x3 borefield with groundwater velocity 161
2x10-7m/s at different groundwater directions under load
profile 3

xix
Figure J.59 Performance of 1x3 borefield with groundwater velocity 161
5x10-7m/s at different groundwater directions under load
profile 3
Figure J.60 Performance of 1x3 borefield with high groundwater 161
velocity after 2 years at different groundwater directions
under load profile 3
Figure J.61 Performance of 1x3 borefield with groundwater velocity 162
10-7m/s at different groundwater directions under load
profile 4
Figure J.62 Performance of 1x3 borefield with groundwater velocity 162
-7
2x10 m/s at different groundwater directions under load
profile 4
Figure J.63 Performance of 1x3 borefield with groundwater velocity 162
5x10-7m/s at different groundwater directions under load
profile 4
Figure J.64 Performance of 1x3 borefield with high groundwater 162
velocity after 2 years at different groundwater directions
under load profile 4
Figure O.1 Logged temperature profiles for Test A in experimental 173
TRT
Figure O.2 Logged temperature profiles for Test B in experimental 173
TRT
Figure O.3 Logged temperature profiles for Test C in experimental 173
TRT
Figure O.4 Logged temperature profiles for Test D in experimental 174
TRT
Figure O.5 Logged temperature profiles for Test E in experimental 174
TRT
Figure O.6 Logged temperature profiles for Test F in experimental 174
TRT

xx
List of Tables

Table 5.1 Default values of parameters used for simulation of 67


borefield performance
Table 5.2 Default parameter values for determination of final 70
configuration of liquid desiccant dehumidifier
Table 5.3 Additional parameter values used for analysis of heat- 71
pump-coupled liquid desiccant dehumidifier
Table 5.4 Details of single zone building for dynamic simulation 72
Table 6.1 Comparison of present model with CSM 74
Table 6.2 Comparison of present model with FLSM 75
Table 6.3 Comparison of present model with LSMGA 75
Table 6.4 Comparison of present model with FLSMGA 76
Table 6.5 Comparison of average ground temperature rise between 78
present model and FLSM after 10 years
Table 6.6 Comparison of average borehole temperature rise between 79
present model and FLSM after 10 years for various
borefields
Table 6.7 Variation of borefield thermal resistances with time for 80
various borefields without groundwater flow
Table 6.8 Variation of peak borefield load with fixed leaving fluid 84
temperature rise at different load profiles for single and 2x2
borefield without groundwater flow
Table 6.9 Variation of peak borefield load at different load profiles 87
and groundwater flow for 2x2 borefield based on fixed
leaving fluid temperature rise at default borehole length
Table 6.10 Parameter estimation results with specified groundwater 94
flow using LSMGA on generated TRT data for 5 days
without groundwater
Table 6.11 Parameter estimation results with specified groundwater 95
flow using LSMGA on generated TRT data for 5 days with
groundwater velocity 2x10-7m/s

xxi
Table 6.12 Parameter estimation results with specified groundwater 96
flow using LSMGA on generated TRT data for 5 days with
groundwater velocity 10-6m/s
Table 6.13 Parameter estimation results with specified groundwater 97
flow using LSMGA on generated TRT data for 30 days with
groundwater velocity 2x10-8m/s
Table 6.14 Parameter estimation results with specified groundwater 97
flow using LSMGA on generated TRT data for 30 days with
groundwater velocity 5x10-8m/s
Table 6.15 Parameter (including groundwater velocity) estimation 98
results using LSMGA on generated TRT data for 5 days
with groundwater velocity 5x10-8m/s
Table 6.16 Parameter (including groundwater velocity) estimation 99
results using LSMGA on generated data for 5 days with
groundwater velocity 10-7m/s
Table 6.17 Parameter (including groundwater velocity) estimation 99
results using LSMGA on generated data for 5 days with
groundwater velocity 2x10-7m/s
Table 6.18 Parameter (including groundwater velocity) estimation 100
results using LSMGA on generated data for 30 days with
groundwater velocity 5x10-8m/s
Table 6.19 Parameter (including groundwater velocity) estimation 100
results using LSMGA on generated data for 30 days with
groundwater velocity 10-7m/s
Table 6.20 Estimated groundwater velocities using LSMGA based 101
on generated data with various ground thermal
conductivities and borehole thermal resistances
Table 6.21 Effects of Qbf and T0 on parameter estimation results 102
using LSMGA on generated data with groundwater velocity
5x10-7m/s
Table 6.22 Effects of Qbf and T0 on parameter estimation results 102
using LSMGA on generated data with groundwater velocity
10-6m/s

xxii
Table 6.23 Applied conditions for different experimental TRTs 104
Table 6.24 Analysis results of experimental TRT data 105
Table 6.25 Percentage error of estimated groundwater flow based on 105
experimental TRT data
Table 6.26 Temperature differences across soil at different applied 106
loads in laboratory soil thermal conductivity measurement
test
Table 6.27 Parameter estimation results of TRT data at north 108
borehole of Luzern using different models
Table 6.28 Parameter estimation results of TRT data conducted on 110
th
18 September 1998 for Maxey School at Lincoln using
different models
Table 6.29 Parameter estimation results of TRT data conducted on 110
30th June 1999 for Maxey School at Lincoln using different
models
Table 6.30 Parameter estimation results of TRT data conducted on 111
th
15 September 1998 for Campbell School at Lincoln using
different models
Table 6.31 Parameter estimation results of TRT data conducted on 111
8th October 1998 for Campbell School at Lincoln using
different models
Table 6.32 Effect of desiccant cooler on performance of liquid 113
desiccant dehumidifier
Table 6.33 Effect of regenerative air heater on performance of 114
liquid desiccant dehumidifier
Table 6.34 Effect of regenerative air heat exchanger on performance 114
of liquid desiccant dehumidifier
Table 6.35 Design requirement of GCLDAC/GSHP 121
Table 6.36 Parameters used for GSHP 121
Table 6.37 Parameters used for GCLDAC 122
Table 6.38 Performance of GSHP/GLCDAC at design conditions 124
for Case 1
Table 6.39 Performance of GSHP/GLCDAC at design conditions 124

xxiii
for Case 2
Table 6.40 Simulation results of GSHP with borehole length 250m 125
at no groundwater flow for Case 1 within 1986~1995
Table 6.41 Simulation results of GCLDAC with borehole length 126
250m at no groundwater flow for Case 1
Table 6.42 Simulation results of GSHP/GCLDAC at no 127
groundwater flow for Case 2
Table K.1 Parameter estimation results with specified groundwater 163
flow using LSMGA on generated TRT data for 5 days with
groundwater velocity 10-8m/s
Table K.2 Parameter estimation results with specified groundwater 163
flow using LSMGA on generated TRT data for 5 days with
groundwater velocity 2x10-8m/s
Table K.3 Parameter estimation results with specified groundwater 164
flow using LSMGA on generated TRT data for 5 days with
groundwater velocity 5x10-8m/s
Table K.4 Parameter estimation results with specified groundwater 164
flow using LSMGA on generated TRT data for 5 days with
groundwater velocity 10-7m/s
Table K.5 Parameter estimation results with specified groundwater 165
flow using LSMGA on generated TRT data for 5 days with
groundwater velocity 5x10-7m/s
Table L.1 Parameter estimation results with specified groundwater 166
flow using LSMGA on generated TRT data for 30 days
without groundwater flow
Table L.2 Parameter estimation results with specified groundwater 166
flow using LSMGA on generated TRT data for 30 days with
groundwater velocity 10-8m/s
Table L.3 Parameter estimation results with specified groundwater 167
flow using LSMGA on generated TRT data for 30 days with
groundwater velocity 10-7m/s
Table L.4 Parameter estimation results with specified groundwater 167
flow using LSMGA on generated TRT data for 30 days with

xxiv
groundwater velocity 2x10-7m/s
Table L.5 Parameter estimation results with specified groundwater 168
flow using LSMGA on generated TRT data for 30 days with
groundwater velocity 5x10-7m/s
Table L.6 Parameter estimation results with specified groundwater 168
flow using LSMGA on generated TRT data for 30 days with
groundwater velocity 10-6m/s
Table M.1 Parameter (including groundwater velocity) estimation 169
results using LSMGA on generated TRT data for 5 days
without groundwater
Table M.2 Parameter (including groundwater velocity) estimation 169
results using LSMGA on generated TRT data for 5 days
with groundwater velocity 10-8m/s
Table M.3 Parameter (including groundwater velocity) estimation 169
results using LSMGA on generated TRT data for 5 days
with groundwater velocity 2x10-8m/s
Table M.4 Parameter (including groundwater velocity) estimation 170
results using LSMGA on generated TRT data for 5 days
with groundwater velocity 5x10-7m/s
Table M.5 Parameter (including groundwater velocity) estimation 170
results using LSMGA on generated TRT data for 5 days
with groundwater velocity 10-6m/s
Table M.6 Parameter (including groundwater velocity) estimation 170
results using LSMGA on generated TRT data for 5 days
with groundwater velocity 2x10-6m/s
Table N.1 Parameter (including groundwater velocity) estimation 171
results using LSMGA on generated TRT data for 30 days
without groundwater
Table N.2 Parameter (including groundwater velocity) estimation 171
results using LSMGA on generated TRT data for 30 days
with groundwater velocity 10-8m/s
Table N.3 Parameter (including groundwater velocity) estimation 171
results using LSMGA on generated TRT data for 30 days

xxv
with groundwater velocity 2x10-8m/s
Table N.4 Parameter (including groundwater velocity) estimation 172
results using LSMGA on generated TRT data for 30 days
with groundwater velocity 2x10-7m/s
Table N.5 Parameter (including groundwater velocity) estimation 172
results using LSMGA on generated TRT data for 30 days
with groundwater velocity 5x10-7m/s
Table N.6 Parameter (including groundwater velocity) estimation 172
results using LSMGA on generated TRT data for 30 days
with groundwater velocity 10-6m/s
Table P.1 Simulation results of GSHP/GCLDAC with Darcy 175
velocity 10-7m/s of groundwater flow for Case 1
Table P.2 Simulation results of GSHP/GCLDAC with Darcy 175
velocity 5x10-7m/s of groundwater flow for Case 1
Table P.3 Simulation results of GSHP/GCLDAC with Darcy 175
velocity 10-6m/s of groundwater flow for Case 1
Table P.4 Simulation results of GSHP/GCLDAC with Darcy 176
velocity 10-7m/s of groundwater flow for Case 2
Table P.5 Simulation results of GSHP/GCLDAC with Darcy 176
velocity 5x10-7m/s of groundwater flow for Case 2
Table P.6 Simulation results of GSHP/GCLDAC with Darcy 176
velocity 10-6m/s of groundwater flow for Case 2

xxvi
List of Abbreviations

ACFAU Air cooled fresh air unit


CSM Cylindrical source model
FLSM Finite line source model
FLSMGA Finite line source model with ground water advection
HPCLDD Heat pump coupled liquid desiccant dehumidifier
GCLDAC Ground coupled liquid desiccant air conditioner
GSHP Ground source heat pump
LSM Line source model
LSMGA Line source model with ground water advection
TRT Thermal response test

xxvii
Nomenclature

A Dimensionless function according to Eq. (2.15)


a Thermal diffusivity (m2/s)
a Specific interfacial surface area of packing (m2/m3)
B Dimensionless function according to Eq. (2.16)
b Centre to centre distance between tubes inside
borehole (m)
C Capacity rate (kW/K)
CAP Capacity of geothermal heat pump (W)
Cap Heat capacity (kJ/K)
Cf = rpi2 f c f / t according to Eq. (3.12) (kW/mK)

COP Coefficient of performance of geothermal heat pump


c Specific heat capacity (kJ/kgK)
d Insulated length of borehole (m)
d = d / H in Eq. (2.45)
DB Dry bulb temperature (oC)
dc Packed column diameter (m)
Deq Equivalent diameter of U-tubes inside borehole (m)
deq Equivalent diameter of packing (m)
di Distance of tube centre to borehole centre (m)
dHG Enthalpy rate change in gas (kW)
dp Diameter of packing (m)
DSH Degree of superheat at compressor suction (oC or K)
Dw Diffusion coefficient of water (m2/s)
E Heat transferred by groundwater (kJ)
erfc Complementary error function
F Overall mass transfer coefficient (kmol/m2s)
Fo = V ' 2 t / a g according to Eq. (2.38)

G Time dependent function according to Eqs. (2.24&25)


G Superficial air mass flowrate (kg/m2s)

xxviii
g Time dependent function according to Eq. (2.2)
H Effective borehole length (m)
h Specific enthalpy (kJ/kg)
h = H gw c gwV gw / 2k g according to Eq. (2.40)

Hac Enthalpy change rate in air cooler (kW)


hc Convective heat transfer coefficient (kW/m2K)
hc Convective heat transfer coefficient corrected for
simultaneous heat and mass transfer (kW/m2K)
hd Specific heat of dilution (kJ/kg)
J0, J1 Bessel function of the first kind of order 0, 1
j1 Summation index in Eqs. (2.27~29)
K0, K1 Modified Bessel function of the second kind of order
0, 1
k Thermal conductivity (W/mK)

L Superficial desiccant mass flowrate (kg/m2s)
M Mass (kg)
m Mass flowrate (kg/s)
m1 Fluid flow direction inside tube in Eq. (3.3)
MSD Mean square difference according to Eq. (3.22) (oC2 or
K2)
Mw Molecular weight of water (kg/kmol)
N Summation index in Eqs. (2.11), (2.21) and (2.22)
n Summation index in Eq. (2.29)
naahxr Number of discretisation segments of air-to-air heat
exchanger at each air stream
nac Number of discretisation segments of air cooler
nbore Number of boreholes
ndata Number of data set in TRT analysis
Nra,he Number of discretisation segment of regenerative air
heat exchanger
nsa Number of discretisation segment of supply air coil
nt Number of tubes inside borehole
NTU = UA / C min according to Eq. (3.38)

xxix
Nu Average Nusselt number over borehole surface
NZ Number of discretisation segment of desiccant packed
tower
nzbore Number of ground grid points along borehole length
P Pressure (kPa)
p = r / rb according to Eq. (2.24)

Pr Prandtl number according to Eqs. (3.37) and (A.9)


PV Percentage variation according to Eq. (3.21)
Q Loading (W)
Q Loading per unit volume (W/m3)

q Loading per unit length (W/m)


q Average heat transfer rate on borehole surface (W/m)
q~ Weighted average of loading per unit length according
to Eq. (3.8) (W/m)
qt Tube loading per unit length (W/m)
R Thermal resistance (mK/W)
~ Weighted average of thermal resistance according to
R
Eq. (3.7) (mK/W)
Thermal interference coefficient according to Eqs.
R
(A.1~3)
Thermal interference coefficient according to Eq. (3.3)
R
r Radial distance from borehole centre (m)
R1 = V / r / a g according to Eq. (2.38)

r1 Mid centre to centre distance between boreholes in


thermal store (m)
Re Reynolds number according to Eqs. (3.36) and (A.8)
Rp Thermal resistance between fluid inside tube and
grouting (mK/W)
rpi Inner radius of tube (m)
rpo Outer radius of tube (m)
RF Run fraction of geothermal heat pump
RH Relative humidity (%)

xxx
S Cross sectional area (m2)
s Specific entropy (kJ/kgK)
s Integrating variable in Eqs. (2.33), (2.34), (2.39),
(2.41) and (2.43)
Sc Schmidt number according to Eq. (3.35)
T Temperature (oC or K)
~ Weighted average of temperature according to Eq.
T
(3.8) (oC or K)
t Time (s)
t Integration variable in Eqs. (2.35)
UA Overall heat transfer coefficient (kW/K)
URF Under relaxation factor according to Eq. (3.16)
V Parameter according to Eqs. (2.27&29)
V = gw c gwV gw / g c g according to Eq. (2.35) (m/s)

V& Volume flowrate (m3/s)


VG Velocity of gas in packed column (m/s)
Vgw Groundwater velocity (m/s)
VOL Volume (m3)
W Mechanical power input (kW)
X = r / 4a g t according to Eq. (2.6)

x Distance in x-direction (m)


x Portion of coil at different refrigerant regions
Y = r 2 / 4a g t according to Eq. (2.9)

y Distance in y-direction (m)


Y0, Y1 Bessel function of the second kind of order 0, 1
Z Desiccant packed tower height (m)
z Distance in z-direction (m)

Subscript
aahxr Air-to-air heat exchanger
AC Air conditioning
ac Air cooler

xxxi
aci Inlet of air cooler
ac1 Intermediate state of refrigerant in air cooler according
to Figure 3.8
adp Air dewpoint
avg Average
b Borehole
bf Borefield
co Condenser outlet
comp Compressor
cond Condenser
conv Convection by groundwate
cool Cooling
deh Dehumidifier tower
dc Desiccant cooler
dci Desiccant cooler inlet
dc1 Intermediate state of refrigerant in desiccant cooler
according to Figure 3.6
dh Desiccant heater
dhi Desiccant heater inlet
dh1, dh2 Intermediate states of refrigerant in desiccant heater
according to Figure 3.5
dh1 Alternative state of dh1
di Dehumidifier tower inlet
dis Compressor discharge
do Dehumidifier tower outlet
dsa Dry supply air
dzone Dry zone air
ea Exhaust air stream in air-to-air heat exchanger
eao Exhaust air discharge
ei Evaporator inlet
evap Evaporator
f Fluid
f Saturated liquid refrigerant

xxxii
fl Fluid at middle of control volume of pipe inside
borehole
flsm Finite line source model
flsmga Finite line source model with groundwater advection
G Gas
g Ground
g Saturated gaseous refrigerant
gc Ground coupled coil
gw Ground water
he Heat exchanger
heat Heating
i Discretisation step of ground and borehole in z-
direction
ia Intake air stream in air-to-air heat exchanger
iao Intake air discharge
in Inlet
inf Thermal interference
inst Instantaneous
j Discretisation step of ground in x-direction
L Liquid desiccant
lat Latent
ll Liquid line
m Discretisation step of ground in y-direction
ma Mixed air
max Maximum
min Minimum
nf No fluid flow inside borehole
ngw No groundwater flow
out Outlet
p Discritisation step of intake air stream in air-to-air heat
exchanger
psga Point source with groundwater advection
q Discritisation step of exhaust air stream in air-to-air

xxxiii
heat exchanger
qfact Load factor
r Refrigerant
ra Regenerative air
rahi Regenerative air heater inlet
rai Regenerative air intake
rao Regenerative air discharge
reg Regenerator tower
ri Regenerator tower inlet
ro Regenerator tower outlet
rs Refrigerant in saturated region
rta Return air
s Heat source
sa Supply air
sat Saturated region
sen Sensible
sc Sub-cooled region
sh Superheated region
store Ground heat store
suc Compressor suction
t Liquid desiccant system
total Total effect including groundwater advection
u,v Tube designation inside borehole
w Water
wet Wet
wG Water in gas phase
wL Water in liquid desiccant phase
x+ From downstream grid point in x-direction
x- From upstream grid point in x-direction
y+ From downstream grid point in y-direction
y- From upstream grid point in y-direction
z+ From downstream grid point in z-direction
z- From upstream grid point in z-direction

xxxiv
zone Building zone
0 Undisturbed condition
Far field

Superscript
n Discretisation step in time

Symbol
Parameter according to Eqs. (2.27&28)
Integrating variable in Eqs. (2.5), (2.6), (2.8) and
(2.25)
= (Tr T f ,in )C f / mr h f g according to Eq. (C.4)

Quality of refrigerant
h f g Specific enthalpy difference between saturated
gaseous and liquid refrigerant (kJ/kg)
f g Density difference between saturated liquid and
gaseous refrigerant (kg/m3)
Heat transfer effectiveness
Euler constant ( = 0.57721)
= g + f g ( in + e NTU ) according to Eq. (C.7)

(kg/m3)
Specific latent heat of evaporation of water (kJ/kg)
l Characteristic length across coil
Dynamic viscosity (kg/ms)
= 2k g / qb according to Eq. (2.38)

Mean dimensionless temperature rise according to Eq.


(2.37)
Ground temperature rise (oC or K)
q Angular displacement in ground from borehole centre
measured counter-clockwise from positive x-axis (rad)
Ground water direction measured counter-clockwise
from positive x-axis (rad)

xxxv
Liquid desiccant concentration
Integrand of Eq. (2.35)
Integrand of Eq. (2.37)
Density (kg/m3)
Mean density (kg/m3)
= (k b k g ) /(k b + k g ) according to Eq. (A.4)

= 4a g t / H 2 according to Eq. (2.40)

= t t according to Eq. (G.1) (s)


* Value of for maximum (s)

Humidity ratio of moist air


= a g t / rb2 according to Eq. (2.24)

Integrating variable in Eqs. (2.36&37)


* Value of for maximum

xxxvi
Chapter 1 : Introduction

1.1 Geothermal Heat Pump Systems

With the increasing concern for global warming nowadays, the reduction of
release of greenhouse gases, especially carbon dioxide into the atmosphere has
become an important task worldwide. The higher usage of renewable energy sources
including solar, wind, tidal, hydrothermal, etc., provides one way of solution.
However, the cost of renewable energy plants used to be higher than that of
conventional power plants. In addition, renewable energy supplies rely on the
environmental conditions, which tend to be unstable. A better utilisation of energy
offers another key direction. Besides transportation, buildings are the major energy
consumers. Further investigation indicates that air conditioning systems constitute
the highest proportion of the total energy demand. Clearly, the development of a
more energy-efficient air conditioning system helps reduce the release of carbon
dioxide and consequently the energy cost significantly.

Air conditioning systems are installed in most buildings. The controlled


indoor conditions in terms of temperature and relative humidity provide a comfort
environment for living and efficient working. In some situations, strict control of the
indoor conditions is necessary for the proper functioning of process lines, computer
rooms, hospitals, etc. Most air conditioning systems adopt the conventional vapour
compression cycle where the heat dissipated from the condenser is transferred to the
surrounding by the circulating water or air. Water-cooled systems offer a higher
energy efficiency than air-cooled systems due to the higher thermal capacity and
usually lower temperature of the condensing water as compared with ambient air.
Direct water-cooled systems which offer the best efficiency require abundant supply
of condensing water, the availability of which depends on the site location. Indirect
water-cooled systems employing cooling towers may induce the risk of
Legionnaires disease. Hence, air-cooled systems are most commonly used despite
their lower energy efficiency.

1
Geothermal heat pump systems which are sometimes referred to as ground
source heat pump systems use the ground as the medium of heat exchange with the
surrounding rather than water or air. Being lower than the ambient in summer and
warmer in winter, the more stable and favourable temperature of the ground
combined with its higher thermal capacity as compared with air makes the systems
more energy-efficient than air-cooled systems. Besides, the release of condensing
heat to the ground instead of the ambient helps relieve the heat island effect. Figure
1.1 shows the schematic diagram of a typical geothermal heat pump system.

Fresh air
Supply Evaporator
air fan

Return air

Condenser

Compressor

Expansion valve

Borefieid circulating pump

Borefield

Figure 1.1 Schematic diagram of typical geothermal heat pump system

The pipework connecting the ground and the condenser may operate as open
or closed systems. For open systems, the underground water is directly pumped from
one well and drained back to the ground through another well, as shown in Figure 1.2.
Open systems offer a better heat transfer performance, but to sustain long term
operation, the hydraulic conductivity of the ground must be high. Moreover,
groundwater may corrode the condenser coil. The application of open systems is
thus limited.

2
Condenser

Ground surface

Figure 1.2 Layout of a simple open type borefield configuration

Closed systems are most common. A fluid which is usually water is


circulated around the condenser and ground coils in a borefield. The ground coils
may be installed vertically or horizontally in the ground. Horizontal ground coils are
usually laid within a shallow depth below the ground surface, and the installation
cost is relatively lower. However, a large area of land is needed, the availability of
which is dependent upon the given site condition. Vertical ground coils, usually in
the form of long U-tubes, are installed inside deep boreholes. The land requirement
for this arrangement is less constrained, which leads to their much wider acceptance.
Figure 1.3 indicates some details of vertical ground heat exchangers. The U-tubes
are usually made of HDPE, and thermally enhanced backfill is added inside the
boreholes to improve the heat transfer rate and to prevent groundwater from running
into the boreholes. Each borehole may contain more than one U-tube, depending on
the design. Sometimes, a tube-in-tube pipe may be used.

Drilling of the boreholes is expensive, especially if very deep boreholes are


required. To reduce cost, several short boreholes may be connected in series in order
to achieve the same fluid temperature change. Besides the drilling cost,
consideration should also be given to the hydrostatic pressure of the circulating fluid
inside the U-tubes at the bottom of the boreholes. For very deep boreholes, the
hydrostatic pressure will be very high, the magnitude of which being calculated at an
approximate rate of 1 bar per 10m of water column. High pressure pipe will

3
therefore be required, which is costly and the thick pipe wall reduces the heat transfer
rate of the boreholes.

Borehole
Tube

Grouting
Ground surface

Tout

Tin

Borefield

Figure 1.3 Details of vertical ground heat exchanger borefield

The application of geothermal heat pump systems in sub-tropical regions like


Hong Kong with cooling-dominated loading profiles will result in accumulation of
heat in the ground. The fluid temperature leaving the borefield will rise gradually
which will reduce the capacity of the air conditioning system. To improve the
situation, longer boreholes should be used. This will incur a high installation cost
due to the drilling of longer boreholes, thus lengthening the payback period
substantially. Groundwater flow helps relieve the ground temperature change, thus
enabling the borehole length to be shortened. Precise modelling of the performance
of ground heat exchanger borefield including groundwater effect is therefore
essential for achieving an optimal system design.

1.2 Liquid Desiccant Dehumidification Systems

Liquid desiccant dehumidifiers have been widely used in industry to provide


a low humidity environment in particular processes with better energy efficiency as
compared with machines employing the conventional vapour compression cycle.
Traditional liquid desiccant dehumidifiers remove the moisture in the process

4
(supply) air by circulating strong liquid desiccant solution in a dehumidifier packed
tower in contact with the air as shown in Figure 1.4. The diluted desiccant solution
is then reheated before entering a regenerator packed tower where it releases
moisture to the regenerative air stream. Usually, the regenerative heat comes from a
low-grade heat source such as waste heat from a power plant or an industrial process,
or even from a solar system.

Desiccant heater Regenerative


air fan

Desiccant
heat
Dehumidifier exchanger Regenerator
Process
tower tower
air fan

Weak Strong
desiccant desiccant
pump pump

Figure 1.4 Basic arrangement of liquid desiccant dehumidifier

The application of liquid desiccant dehumidifiers in HVAC systems is


increasing in response to the rising demand for a better indoor environment having a
lower moisture content without use of excessive energy. They are also excellent in
treating the fresh air, especially in sub-tropical regions like Hong Kong where the
outdoor air is both hot and humid in summer. Large latent loads can be handled in
this way with a relatively lower energy consumption. However, conventional liquid
desiccant dehumidifiers cannot provide sensible cooling. Therefore, extra equipment
is needed to handle the sensible heat in order to meet the required air conditioning
load.

The addition of heat pumps to liquid desiccant dehumidifiers may solve the
problem. Heat dissipated in the condenser is used to regenerate the liquid desiccant
before entering the regenerator tower, and the evaporator is used to cool the liquid
desiccant before entering the dehumidifier tower, thus providing a certain amount of

5
sensible cooling. Still, such design can only handle the fresh air with satisfactory
performance, and auxiliary facility is needed to deal with the return air.

1.3 Objective of This Research

Most of the available analytical models do not consider the groundwater


effect. In dealing with more than one borehole, the superposition method usually
requires approximation of the thermal interference effect from the adjacent boreholes.
Even for common numerical methods, the schemes are either two dimensional or in
lack of the coupling effect of varying fluid temperature along the U-tube inside
boreholes. In view of this, a new numerical model which incorporates the
groundwater effect for the entire borefield will be developed. This model is three-
dimensional and considers the interaction of varying fluid temperature along the U-
tube. With the new numerical model, the various aspects of the groundwater effect
on performance of a borefield instead of just a single borehole can be investigated.
The fluid leaving temperature rise will be the main parameter to study.

As the groundwater effect is critical for deciding whether the application of


geothermal heat pump systems in sub-tropical regions will be feasible, the
determination of groundwater velocity becomes very important. Also, the ignorance
of groundwater effect in the thermal response test analysis can lead to erroneous
results in the estimated ground thermal conductivity. Hence, the new numerical
model will be applied to the thermal response test analysis and the parameter
estimation scheme modified to include groundwater velocity as one of the parameters
to be determined. In this way, the correct ground thermal conductivity can be
worked out. Experimental thermal response test will also be conducted to verify the
new parameter estimation scheme.

A hybrid geothermal heat pump system provides an opportunity for


successful application in sub-tropical regions, where part of the loading is shared by
the partner system which usually consists of cooling towers. In this case, the
borehole length will be reduced at the expense of a lower energy efficiency. To
overcome this weakness, a new partner system needs to be explored. As a liquid

6
desiccant dehumidifier coupled with a heat pump is efficient in handling the fresh air,
the combination of a liquid desiccant dehumidifier with the conventional geothermal
heat pump system seems to be a suitable choice where a single compressor will be
used to provide desiccant cooling/heating and air conditioning for the return air.
Computer simulation will be carried out to investigate the potential of this new
hybrid system in terms of the borehole length reduction and overall energy efficiency
change under different groundwater conditions.

7
Chapter 2 : Literature Review

2.1 Geothermal Heat Pump Systems

2.1.1 General

With the awareness of energy crisis in late 1970s and environmental concern,
the quest for sustainable power sources and energy-efficient systems has become a
hot issue. Besides industrial processes, building air conditioning systems, especially
for heating purposes in winter, account for the major portion of the total energy
consumption. The development of better heating systems is one key direction for
improvement. The cogeneration of heat and power from power plants is an obvious
solution, but heating can only be provided to nearby areas around the power stations.
Geothermal energy is another possible choice where high-grade geothermal heat can
be directly used for power generation and medium-grade heat can be extracted for
space heating. ASHRAE (2003) outlined some details in designing geothermal
power system for heating service. Lund (2003) summarised the status of direct-use
of geothermal energy in USA, while Kepinska (2003) discussed the situation in
Poland.

However, the availability of geothermal heat sources for direct heating is still
restricted to limited locations. Alternatively, geothermal heat pump systems can be
developed to extract heat from the ground through conventional vapour compression
systems, where the amount of heat supplied can be much higher than the energy
input due to the high COP value of the systems. Research work increases rapidly in
Europe and USA. Gourlburn and Fearon (1983) investigated a domestic heat pump
system with the low pressure refrigerant directly fed into the U-tube of the ground
heat exchangers for evaporation. Sliwa and Kotyza (2003) discussed the
employment of existing wells for geothermal heat pump systems in Poland. Petit and
Meyer (1998) compared the costs of geothermal and air-cooled systems in South
Africa. Hepbasli and Akdemir (2004) performed energy and exergy case analysis of
particular geothermal heat pump installation in Turkey. Sanner et al. (2003) briefed

8
the current situation of geothermal heat pump system application in Europe. The
systems gain greater acceptance in Sweden, Switzerland and Austria, but the
development is slow in UK, possibility due to the competition from cheap gas supply.

In USA, geothermal heat pump systems are not only used for heating, as cooling
is also required in summer. This helps relieve the ground temperature change, thus
reducing the required borehole lengths and consequently the installation costs and
promoting the wider acceptance of the systems. ASHRAE (1998) presented case
studies of various installations in USA. Various design guidelines/manuals were
issued, including those by Bose et al. (1985), ASHRAE (1995) and Kavanaugh and
Rafferty (1997). Sachs (2002) discussed the geological aspects of geothermal heat
pump system installation. Bose et al. (2002) presented an overview of the recent
situations and developments of geothermal heat pump systems in USA. Spitler
(2005) reviewed the design methodology of geothermal heat pump systems, and
proposed areas for further investigation.

The development of geothermal heat pump systems in Asia concentrates in


North East Asia (Japan, South Korea, China) where space heating demand in winter
is very high. He (2007) highlighted some recent research works in Asia. She also
presented two case studies, one in Shanghai employing energy piles for cold storage,
and the other one the first local installation in the Hong Kong International Wetland
Park (Lau and Suen, 2003) comprising more than 400 hundred boreholes. Lam and
Wong (2005) also investigated the design for the Hong Kong International Wetland
Park, and commented that cyclic operation in cooling and heating modes was
necessary to avoid rapid increase in ground temperature for long-term operation.

To reduce the high installation cost of geothermal heat pump systems due to
drilling of long boreholes, hybrid systems are developed. The most common design
is the parallel operation of ground heat exchangers and cooling towers. Kavanaugh
(1998) outlined a design methodology of such system, and discussed the various cost
consideration and operation criteria for the system. Yavuzturk (1999) also studied
the performance of hybrid geothermal heat pump systems coupled with cooling
towers based on his short time-step model. Gasparella et al. (2005) described the
operation of a combined liquid desiccant and geothermal heat pump system in Italy

9
with different operation modes in various seasons. The reduction in borehole length
can also be achieved by minimising the unbalance of the annual borehole loads. In
heating dominated regions, this can be accomplished by re-charging the heat into the
ground by operating a solar system in summer. Ozgener and Hepbasli (2006)
discussed the costing of a solar-assisted geothermal heat pump system.

In projects where concrete piles are to be used for structural reason, the
insertion of ground heat exchangers into concrete piles (energy piles) can
significantly lower the installation cost of the geothermal heat pump systems, as the
drilling work is significantly reduced. Pahud et al. (1996) developed a modified
version of the Hellstroms DST model implemented in TRNSYS (DSTP) to estimate
the performance of energy piles. Fromentin et al. applied the DSTP model to
simulate the performance of several existing installations. Pahud (1999) further
modified the DSTP model to PILSIM for ease of application. Pahud et al. (1999)
applied PILSIM for design in a project at the Zurich Airport. Morino and Oka (1994)
studied numerically and experimentally the heat transfer between a steel pile
circulated with water and the surrounding soil. Hamada et al. (2007) investigated the
performance of energy piles installed at site. They also analysed the capacity of
energy pile with different configuration of heat exchanger pipes installed, and found
that indirect double pipe provided a higher capacity than single or double U-pipe.
Laloui et al. (2006) studied numerically and experimentally the mechanical
behaviour of energy pile at different mechanical and thermal loading conditions.
Brandl (2006) discusses various types of thermal structures and presented case
studies of various installations.

Research on horizontal ground coil systems is relatively few. Mei (1988)


studied the performance of a horizontal ground coil system having two layers of
tubes. He used a three-dimensional finite difference scheme in cylindrical
coordinates with the centre-line of the lower tube as the cylinder axis, and adopted
the equation proposed by Kusuda and Archenbach (1965) for the variation of far-
field temperature along the ground depth. The numerical model was used to simulate
the performance of an existing installation with good agreement with the site data.
Mei (1991) compared the simulated results using his model with those based on LSM

10
and modified LSM to the experimental data, and found that his model predicted
better than the other two models, especially if the effect of backfill was included.

2.1.2 Modelling of Vertical Ground Heat Exchanger Performance

Various models have been developed and design methodologies implemented


in computer software are available in the market. Thornton et al. (1997a) compared
several sizing tools for the ground heat exchangers with a detailed simulation model
calibrated before with site data (Thornton et al., 1997b). Shonder et al. (1999, 2000)
repeated the comparison with updated version of the software. The TRNSYS DST
model was used as the benchmark. Indeed, each simulation tool usually imposes
some assumptions or restrictions in the design method such as borehole layout and
connection configuration, loading pattern, ground condition, etc. Hence, proper
understanding of the principle behind is important.

Heat is transferred in the ground surrounding the boreholes by conduction


and convection as well if groundwater flow is present. U-tubes are commonly used
inside boreholes to allow the circulating fluid to pass through. The geometry of the
U-tubes and the complex temperature interaction inside and outside the boreholes
render the derivation of an exact analytical solution extremely difficult if not
impossible. Analytical models with assumption in borehole performance and
numerical schemes are thus developed. In the case of analytical models, a single
borehole is usually considered and the effect of entire borefield is obtained by
applying the principle of superposition.

All analytical models are based on a continuous load step. In dealing with a
varying load profile, the loading is piecewise transformed into a number of step loads
as shown in Figure 2.1, and the final solution by superposition. When the simulation
period is long, the quantity of load steps becomes very large, which leads to very
long computation time. Bernier et al. (2004) outlined a multiple load aggregation
algorithm which allowed annual simulation to be conducted efficiently even at
hourly intervals. Past thermal loads were aggregated in parts at different time
intervals corresponding to the past day, week, month and years while recent thermal

11
loads were not aggregated. A similar approach was adopted by Yavuzturk (1999),
but he lumped up the past loads with fixed time intervals.

Load Load

Q1

Q1 Q2 Q3 Q4 Q5 Q2-Q1

Q3-Q2

Time Time
Q4-Q5
Q3-Q4

Figure 2.1 Transformation of load profile to series of step loads

The thermal interaction between a borehole and the surrounding ground


affects the borehole surface temperature. In general,
Tb T0 = qb R g (2.1)

where R g is a time dependent thermal resistance of the ground and q is the heat load

per unit length rejected to the ground. R g characterises the ground response due to

unit loading from the borehole. According to Eskilson (1987),


g
Rg = (2.2)
2k g

where g is also a time-dependent function usually in terms of some dimensionless


parameters. Concerning the heat transfer inside the boreholes, the correlation
between the borehole surface temperature and mean fluid temperature can be
expressed as
T f Tb
qb = (2.3)
Rb

12
where Rb is the thermal resistance of the boreholes. The overall effect of the

circulating fluid on the ground is then defined by combining Eqs. (2.1) and (2.3),
giving
T f T0 = qb ( Rg + Rb ) (2.4)

By applying the analogy between electric current and heat flow, Eqs. (2.3) and (2.4)
can be symbolically represented as shown in Figure 2.2.

Rb Rs
Tf Tb T0
Ground
qb

Borehole

Figure 2.2 Electrical analogy of heat transfer between borehole circulating fluid and
surrounding ground

2.1.2.1 Modelling without Groundwater Effect

One of the earliest adopted analytical models is the line source model (LSM)
from Ingersoll et al. (1954) for calculation of heat conduction around a single
borehole without groundwater effect under a constant heat load step. The line source
model assumes that the borehole can be approximated by an infinite line of radius
zero, and the ground temperature only varies in the radial direction. The temperature
rise is calculated as

e
2
r
I
q qb
Tg T0 = b
2k g
d =
2k g 4a t
(2.5)
r / 4ag t g

e
2
r
where I ( ) = d and = (2.6)

4a g t

Tabulated values for I () are given by Ingersoll et al. (1954). For < 0.2 ,

13
1 4
2
I ( ) ln + 0.2886 (2.7)
2 8

In real situations, the borehole size is finite, and heat is propagated from the
borehole surface instead of the borehole centre. Hence, the heat front from a line
source will lag behind the actual thermal effect, especially soon after the start of the
load step and at distance close to the borehole centre. Ingersoll et al. (1954) stated
that the line source model is not accurate at short time duration and for borehole with
large diameter where a g t / rb2 < 20 . Eq. (2.5) can alternatively be written as

e r2
qb qb
Ei = qb Ei ( )
Tg T0 =
4k g 2

d =
4k g 4a g t 4k g
(2.8)
r / 4ag t

where Y = r 2 / 4a g t (2.9)

and Ei ( ) is the exponential integral function. Tabulated values for Ei ( ) are


given by Poulikalos (1994). According to Abramowitz and Stegun (1964),
Ei( ) ln( ) + 0.99999193 0.24991055 2 + 0.05519968 3

0.00976004 4 + 0.00107857 5 (2.10)


A more general formulation for the exponential integral function is given by Spiegel
and Liu (1999) as

(1) N +1 N
Ei ( ) = ln( ) + (2.11)
N =1 N N!
For small Y (large t), Ei(Y ) ln Y and Eq. (2.10) becomes

q b 4a g t
Tg T0 ln 2 (2.12)
4k g r
Eskilson (1987) suggested that Eq. (2.12) was applicable for 5rb2 / a g < t < H 2 / 90a g .

LSM is adopted by IGSHPA (International Ground-Source Heat Pump


Association) and the approach was detailed in Bose et al. (1985) and quoted by Cane
and Forgas (1991) for the design of ground heat exchangers. They approximated the
integral I(X) in Eq. (2.6) by empirical equations at different ranges of X. For
0 < 1,
I ( ) 0.5( ln 2 0.57721566 + 0.99999193 2 0.24991055 4

14
+ 0.05519968 6 0.00970064 8 + 0.00107857 10 ) (2.13)
while for 1 < ,
1
I ( ) 2
(2.14)
2 2 e
where
= 8 + 8.5733287 6 + 18.059017 4 + 8.637609 2 + 0.2677737 (2.15)
and
= 8 + 9.5733223 6 + 25.632956 4 + 21.0996531 2 + 3.9684969 (2.16)
Eq. (2.10) and (2.13) are basically the same. Bose et al. (1985) defined the ground
formation resistance for single borehole as
I ( rb ) rb
R g ( ) = with rb = (2.17)
2k g 4a g t

By referring to Eq. (2.2), g = I ( rb ) . For multiple borehole systems, the thermal

interference effects from adjacent boreholes are determined by superposition method.


The required borehole length is then calculated with a modified version of Eq. (2.4),
taking into account the run fraction (RF) of the system. The corresponding formulae
for heating and cooling modes are
COPheat 1
CAPheat ( Rb + R g RFheat )
COPheat
H heat = (2.18)
T0 T f ,min

and
COPcool + 1
CAPcool ( Rb + R g RFcool )
COPcool
H cool = (2.19)
T f ,max T0

where CAP is the capacity of the heat pump system, Rb is the borehole thermal
resistance as defined in Eq. (2.3) and Rs is the ground thermal resistance as defined in
Eq. (2.1).

Hart and Couvillion (1986) also applied the LSM and derived method used
by NWWA (National Water Well Association). They assumed that the ground
temperature remained unchanged beyond a far field radius defined as
r = 4 a g t (2.20)

15
They proposed an equation for the calculation of ground temperature rise as

qb r
(1) N +1 4r 2
N

Tg T0 = ln 0.9818 + 2 (2.21)
2k g r 2 N N ! r
N =1

and

1 r
(1) N +1 4r 2
N

Rg = ln 0.9818 + 2 (2.22)
2k g r N =1 2 N N ! r

Similar to the IGSHPA approach, Eq. (2.4) was used to calculate the borehole length.
ASHRAE (1995) outlined design procedures based on their approach. Kyriakis et al.
(2006) adopted LSM and derive maximum borehole loading under different load
duration and undisturbed ground temperatures.

The weakness of requiring long time step for LSM is overcome by the
application of cylindrical source model (CSM) developed by Carslaw and Jaeger
(1959), which assumes the borehole to be a cylinder of infinite length. Constant heat
flux is injected from the cylinder surface. Again, the ground temperature only varies
in radial direction, and groundwater effect is not considered. The solution is
obtained by solving the following one-dimensional heat conduction equation in
cylindrical coordinates by Laplace transform method,
2 1 1
+ =0 for rb < r < (2.23)
r 2 r r a g t

where = Tg T0 , yielding

qb ag t r
= G ( , p ) with = 2
p= (2.24)
kg rb rb

and

e 1
[J 0 ( p )Y1 ( ) J 1 ( )Y0 ( p )] d2
2
1
G ( , p) = 2 2 (2.25)
0 J 1 ( ) + Y1 ( )
2

By comparing Eq. (2.24) with Eqs. (2.1&2),
G ( ,1)
Rg = and g = 2G ( ,1) (2.26)
kg

According to Hellstrom (1991), the evaluation of the integral G ( , p) was


time consuming due to the limit of the integration and the oscillatory behaviour of

16
the Bessel functions. He quoted an alternative form of solution by applying a
numerical inversion technique to the Laplace domain, and the solution becomes
q 10 V j1 K 0 ( j1 r )
=
2rb k g

j1=1 j1 j1 K 1 ( j1 rb )
(2.27)

j1ln(2)
where j1 = (2.28)
ag t

and
Min ( j1, 5 )
(1) j15 n 5 (2n)!
V j1 =
n = Int (( j11) / 2 ) (5 n)!( n 1)! n!( j1 n)!( 2n j1)!
(2.29)

Fujii et al. (2004a) also proposed formulae to approximate the borehole temperature
rise G ( ,1) based on CSM as follows:

G = 0.1443 0.3374 0.0162 for < 1 (2.30)

G = 0.5414 0.0986 0.4166 for 1 < < 100 (2.31)


G = 0.1827 log10 + 0.0668 for 100 < (2.32)
Bernier et al. (2004) adopted an approximate analytical expression derived by
Cooper (1976) for the CSM.

CSM has been widely used by previous workers. Deerman and Kavanaugh
(1991) and Kavanaugh (1992) applied CSM in their analaysis. Bernier (2001) also
adopted CSM in his work. Kavanaugh (1985, 1995) used CSM to calculate the
ground thermal resistance and proposed modified formulae similar to Eqs. (2.18&19)
to calculate the required borehole length in cooling and heating modes. He included
terms to account for thermal interference between adjacent legs of the U-tubes inside
boreholes, and between adjacent boreholes. His approach was implemented in the
GchpCalc design software and presented in Kavanaugh and Rafferty (1997).
Bernier (2000, 2006) proposed modification of Kavanaughs approach.

Dobson et al. (1995) discretised the U-tube of a ground heat exchanger with
surrounding ground by various horizontal layers, and applied CSM for each portion
of tubes. The coupled fluid temperatures along the U-tube were solved numerically.
The variation of ground thermal properties at different depths was also accounted for.
Sutton et al. (2002) adopted a similar approach to estimate the performance of a
vertical ground heat exchanger in a stratified geological regime. They used Eqs.

17
(2.27~29) to calculate the ground temperature change. The U-tube inside the
borehole was represented by a single tube with equivalent diameter.

When applying LSM or CSM, no consideration is made on the ground


surface, as the borehole is assumed to be infinitely long. Indeed, with a continuous
constant load, both LSM and CSM yield solutions which will not converge, i.e. the
ground temperatures will increase continuously with time. To overcome this
problem, a finite line source model (FLSM) was proposed by Eskilson (1987) and
quoted by Zeng et al. (2002) and Diao et al. (2004). The borehole was represented
by a finite line source, and zero ground surface temperature change maintained by
imposing an image line heat sink of equal strength above the ground surface. The
assumption of constant ground surface temperature means that heat is exchanged
between the ground and the ambient. In this sense, a steady state can be reached
even with continuous step load. The solution takes the form
r 2 + ( z s ) 2 2
erfc r + ( z + s )
2
erfc
4a g t 4a g t
qb d + H ds
4k g d
= (2.33)
r 2 + ( z s ) 2 r 2 + ( z + s ) 2
Unlike LSM and CSM, the ground temperature varies in both radial and vertical
directions. Zeng et al. (2002) derived an expression for the steady-state ground
temperature and mean borehole surface temperature. They commented that the
difference between the steady-state mean borehole surface temperature rise and that
at mid-depth of borehole surface was between 3.5 to 5% for 0.0005 < rb / H < 0.005 .
By adopting the borehole mid-depth temperature rise as the representative borehole
temperature rise and combining with Eqs. (2.1&2),
r 2 + ((d + H ) / 2 s ) 2
erfc b
4a g t
1 d +H
g=
2 d rb2 + ((d + H ) / 2 s ) 2

r 2 + ((d + H ) / 2 + s ) 2
erfc b
4a g t
ds (2.34)
rb2 + ((d + H ) / 2 + s ) 2

Lamarche and Beauchamp (2007) defined the g function based on the average
borehole surface temperature, resulting in a double integral. They proposed

18
alternative expression for the double integral which was much faster to compute.
However, in their derivation, they assumed that the borehole heat exchanger had no
insulated depth (i.e. d = 0 ).

Eskilson (1987) also developed a finite difference scheme using a two-


dimensional cylindrical coordinate system to calculate the ground temperature
change around a single borehole based on the same assumption as adopted in FLSM.
He computed the g functions for various configurations of borefield which were
presented graphically. The application of numerical techniques allows more
complex boundary conditions such as those with temperature gradient at far field
ground and even inclined boreholes to be analysed, which can otherwise not be
handled by using FLSM. There are slight differences between the results using
Eskilsons numerical model and FLSM. His approach is applied in the EED
software (Hellstrom and Sanner, 2000) and the GLHEPRO software (Spitler, 2000
and Spitler et al., 2000).

Hellstrom (1989, 1991) developed a duct ground thermal storage model


which assumes that all boreholes be aligned in a closed packed pattern. He divided
the heat transfer process into three parts, the steady-state heat transfer between
circulating fluid and surrounding ground, steady flux condition between adjacent
boreholes inside the thermal store, and global thermal interaction between the store
and the surrounding ground. The final solution was based on the superposition of the
three parts. Unlike LSM, CSM and FLSM, Hellstroms model calculates the
performance of the entire borefield simultaneously. A simplified numerical version
using a two-dimensional finite difference method with cylindrical coordinate was
implemented into the TRNSYS DST module (Hellstrom et al., 1996). Due to the
assumed configuration of the boreholes, Hellstroms model is not accurate for small
scale borefield or the boreholes aligned in a pattern with large aspect ratio.

In applying the models mentioned above, it is assumed that steady state


occurs inside the borehole. This is appropriate when the loading time step is long,
and the heat transfer rate related to the borehole thermal resistance as already
indicated in Eq. (2.3). In most situations where U-tubes are used inside borehole, the
fluid temperature along each leg of the U-tubes affects each other. Gu and ONeal

19
(1998a) developed an expression for an equivalent diameter ( Deq = 2 rpo b ) to

appropriate the two legs of U-tube. The resultant borehole became a composite
cylinder and Gu and ONeal (1995, 1998b) investigated the borehole performance
based on different thermal properties of the backfill.

Claesson and Hellstrom (1987) derived analytical equations to calculate the


thermal interference coefficients between adjacent legs of multiple U-tubes inside
borehole based on line source approximation in a composite circular region while
Bennet at el. (1987) used a multipole method. The results from the former were
adopted by Eskilson and Claesson (1988) to formulate coupled linear equations to
describe the fluid temperature variation along the two legs of a single U-tube inside
borehole. They proposed an expression for the solution which allows the borehole
temperature to vary along the borehole. Zeng et al. (2003) followed a similar
approach but assumed a constant borehole temperature, and derived analytical
solutions for the fluid temperature variation for single and double (with various
connection configurations) U-tubes inside borehole. They also proposed a single
expression for the borehole thermal resistance.

Yazuzturk (1999) investigated the transient response of a single borehole at


short time step using a two-dimensional radial and circumferential cylindrical finite
volume scheme. He finely discretised the interior of the borehole including the U-
tube (approximated by a pie sector), and calculated the transient temperature change
inside the borehole and surrounding ground. The temperature effect in the
longitudinal direction was ignored. He computed a short time step g function curve
to complement the g function curves derived by Eskilson (1987) with long time steps.
Yazuzturks model is also adopted in the GLHEPRO software. Young (2004)
proposed a borehole fluid thermal mass model based on the buried electrical cable
model developed by Carslaw and Jaeger (1959) to analyse the short term borehole
thermal response.

Rottmayer et al. (1997) developed a two-dimensional radial and


circumferential finite difference scheme in cylindrical coordinate to calculate the
transient heat transfer inside and outside a single borehole. The leg of the U-tube

20
was represented by a non-circular section, and a geometry factor introduced to
account for this. Vertical heat conduction in the ground was neglected, and a quasi
three-dimensional analysis is made by applying the numerical scheme layer by layer
so that the temperature variation along the U-tubes could be estimated.

Moujaes (1990) also used a two-dimensional finite difference scheme in


rectangular coordinates to estimate the three-dimensional temperature variation in the
ground and the fluid temperature change along the U-tube under cyclic operation.
Vertical heat conduction in the ground was again neglected. However, he assumed
the two legs of U-tube to be just two pipes in the ground, and no grouting was
considered. He also accounted for the transient response in the fluid. Kavanaugh
(1985) applied a two-dimensional radial and vertical finite difference scheme to
calculate the performance of a single borehole with a concentric tube. Sliwa and
Gonet (2005) also analysed numerically the performance a tube-in-tube borehole
converted from existing oil well. Muraya et al. (1996) analysed the thermal
interference effect of adjacent legs of U-tube using a finite element scheme.
Yavuzturk and Chiasson (2002) compared the performance of borehole with U-tubes,
concentric pipes and standing column well. They commented that a reduction of
22%, 33% and 36% borehole length could be made by using double U-tubes,
concentric pipes and standing column well without bleed rather than single U-tubes.

2.1.2.2 Modelling with Groundwater Effect

One of the simplest analytical model including groundwater effect is the line
source model with groundwater advection (LSMGA) based on the moving line
source concept as stated by Carslaw and Jaeger (1959). They presented results for
steady state condition. Diao et al. (2004) derived an expression for the transient
temperature rise in the ground based on LSMGA as
[ x V ' ( t t ' )] 2 + y 2
q t 1 gw c gwV gw
0 t t ' e
4a g (t t ' )
= b dt ' V' = (2.35)
4k g g cg

or in polar coordinates (r , ) as
V ' r cos 1 V ' 2 r 2

q 2 a g 4ag t / r 2 1 16 a 2
= b e
d
g
e (2.36)
4k g 0

21
They also proposed a formula for the average dimensionless temperature rise around
a circle from the line source which can be used to compute the average borehole
temperature rise as
1 R12

1 R1 4 Fo / R12 1 16


= d =I e
d (2.37)
2
0
0 2 0

2k g V 'r V ' 2t
where = R1 = and Fo = (2.38)
qb ag ag

Claesson and Hellstrom (2000) adopted the solution from Carslaw and Jaeger
(1959) for the ground temperature rise around a finite line source along 0 < z < H in
an infinite surrounding with groundwater advection. They expressed the borehole
temperature rise in terms of a generalised g function based on the same approach as
used in Eqs. (2.1&2) as given by g total = g ngw g gw , and derived a formula for g gw

as

g gw ( , h) =
1
(
0 2s '
1 e h2s' / 4
) (
erfm 1 / s ' ds ' ) (2.39)

where
4a g t H gwcgwVgw
= h' = (2.40)
H2 2k g

and

1 ex
2
1 x
erfm( x) = erf ( s ' )ds ' = erf ( x) (2.41)
x 0 x
For < 1 , they proposed

h '2 1 h ' h'


g gw 0.5Ein ' erf (2.42)

4 h 2
where

1 e s '
'
x
Ein( x) = '
ds ln( x + e x ) + 0.577(1 e 7 x / 4 ) (2.43)
0 s
They further simplified Eq. (2.42) for h ' < 1 as

h'2 4
g gw 1 (2.44)
8
9

22
Following the same approach as in FLSM where no temperature rise at ground
surface is assumed by adding an image sink above the ground and including
insulated depth for borehole, the final g function for finite line source model with
ground water advection (FLSMGA) is given as
g flsmga = g flsm g gw ( , h ' ) + (1 + d ' ) g gw ( / 4(1 + d ' ) 2 ,2(1 + d ' )h ' )

+ d ' g gw ( / 4d '2 ,2d 'h ' ) (1 + 2d ' ) g gw ( /(1 + 2d ' ) 2 , (1 + 2d ' )h ' ) (2.45)

where d ' = d / H . However, the thermal interference effects between boreholes


were not considered by Claesson and Hellstrom (2000). Nevertheless, FLSMGA is
applied in the modified version of EED to include groundwater effect.

Pahud and Hellstrom (1996) modified the DST model to account for
groundwater effect in two manners, long-term and short-term effects. According to
Pahud et al. (1996), the long-term effect, which influences the global temperature
field, determines the heat transferred by groundwater. Two methods were proposed.
The first one is based on the average temperature difference between the inner and
outer surface of the store boundary as
Econv = Vgw S store gwcgw (Tstore ,in Tstore , out )t (2.46)

where t is the time step for calculating the global temperature field in the ground.
The second method is based on the difference between the average store temperature
and the undisturbed ground temperature as
Econv = Vgw S store wcw (Tsore, avg T0 )t (2.47)

and an upper limited is defined as


Econv , max = VOLstore g cg (Tstore , avg T0 ) (2.48)

The calculated Qconv is then equally distributed as a temperature correction on each


cell of the ground layer within the store.

The short-term effect influences the local solution governing the heat transfer
between borehole and surrounding ground. Based on the equation for calculating the
average steady state heat transfer at a cylinder surface inside a porous medium (Nield
and Bejan, 1992) applied to the boreholes,
0.5
q conv 2V r
Nu = Nu = 1.015 gw b (2.49)
(Tb T0 )k g a
g

23
where q conv is the average convective heat transfer on the borehole surface. As DST
assumes the borehole to be closely packed, the heat flux at mid-distance between
boreholes becomes zero, and Hellstrom (1991) gave

q r12 r1 1 rb2
Tb T0 = 2 ln + 2 (2.50)
2k g r1 rb
2
rb 2 2r1

where r1 is the mid centre-to-centre distance between two boreholes. Combining Eq.
(2.49) and (2.50),

q conv Nu r12 r1 1 rb2


= ln + (2.51)
q 2 r12 rb2 rb 2 2r12

According to Pahud and Hellstrom (1996), Eq. (2.51) represents the difference
between actual heat transfer as compared with that without groundwater effect, and
the ratio is used as a correction factor in calculating the heat transfer rate of the
boreholes only if the ratio is greater than 1.
Chiasson et al. (2000) applied a two-dimensional finite element scheme to
discretise the boreholes and the surrounding ground under groundwater advection.
They analysed the variation of average borehole fluid temperature at different soil
materials under a preset hydraulic gradient. They found that with shale where the
groundwater velocity is very low, the borehole performance was not much different
from that without groundwater flow. With fine sand, the difference was only minor.
However, for coarse sand with high groundwater velocity, the difference was very
significant. They also investigated the effect of groundwater velocity on the thermal
response test result by simulating test data using the numerical scheme and analysed
the data based on the method adopted by Austin et al. (2000). The predicted thermal
conductivity increases with groundwater velocity, and the variation is more
pronounced with longer test period. Finally, the numerical scheme was used to
simulate the performance of a 4x4 borefield under various groundwater conditions.

Niibori et al. (2005) used a two-dimensional finite difference scheme to study


the ground temperature variation around a single borehole at different groundwater
conditions. The numerical model was also verified with site data with fairly good
accuracy in one loading cycle of several hours. Gehlin and Hellstrom (2003b) also
developed a two-dimensional finite difference model to evaluate the effect of

24
groundwater in thermal response test. Li et al. (2005) applied a two-dimensional
finite difference scheme which assumed that heat and mass transfer in the vertical
direction was negligible, and studied the quasi three-dimensional heat and moisture
transfer in a borefield. The boreholes were taken as inner heat sources in the
discretisation control volume. The performance of single borehole at different soil
properties and loading patterns was analysed. The model was further used (Li et al.,
2006) to simulate the performance of an existing installation comprising both
conventional vertical borehole heat exchangers and energy piles.

Fujii et al. (2004b) utilised the FEFLOW software (Diersch, 2002) to estimate
the groundwater flow pattern and heat transfer in an existing installation based on
site measured boundary parameters. Each borehole was approximated as a thermal
well. They compared the simulated data with CSM at no groundwater flow, giving
good agreement. The simulated performance for a single borehole with groundwater
flow was validated with thermal response test data, showing very little difference in
the heat exchange rate. The long-term behaviour of the entire borefield was then
predicted under four loading patterns. They commented that in designing a ground
thermal store with groundwater advection, the heat storage period and frequency
should be carefully selected depending on the groundwater velocity. He (2007) also
used FEFLOW to simulate the groundwater flow pattern in the Hong Kong
International Wetland Park. The estimated groundwater velocity was very small
with the same order of magnitude as the one calculated based on geological data
from construction site in the same district area.

2.1.3 Thermal Response Test Analysis

The thermal properties of the ground strongly affect the performance of a


borefield. Hence, an accurate prediction of the soil properties is very important,
especially for the thermal conductivity. A thermal response test (TRT) was
developed to estimate various parameters which are important in the design of the
geothermal heat pump system. According to Georgiev et al. (2006), the first TRT
was presented by Mogensen (1983) for stationary systems. TRT with mobile device
was later developed by Eklof and Gehlin (1996) and Austin (1998) respectively.
Basically, a constant load was injected into the borehole through circulating fluid,

25
and the fluid temperatures at inlet and outlet of borehole recorded. Medium-scale
laboratory TRT was made by Smith et al. (1999b) and recently, micro-scale
laboratory TRT was conducted by Katsura et al. (2006).

The underlying principle of TRT falls into two main approaches. The first
one is the application of LSM to determine the ground thermal conductivity
graphically. By combining Eq. (2.3) and (2.12),

qb 4a g t
T f T0 ln 2 + qRb (2.52)
4k g rb

Hence, a plot of T f T0 vs ln(t ) should give a straight line of slope q / 4k g . By

measuring the gradient of the fluid temperature rise profile, the ground thermal
conductivity can be calculated. Besides, the borehole thermal resistance can also be
estimated.

Pahud and Mathey (2001) applied this method to determine the thermal
resistance of a borehole with double U-tubes at different alignments and backfills,
and the results were compared with those calculated based on the equations from
Hellstrom (1991). Pahud (2001) also used this approach to analyse the TRT results
of two boreholes near Luzern, and the results were compared with those obtained by
laboratory test of soil samples. The advantage of this approach is ease of application
without the need of computation work. However, LSM assumes no groundwater
flow. In case this is not fulfilled, the calculated ground thermal conductivity will
increase with time, as noted by Katsura et al. (2006). The calculated ground thermal
conductivity increases with groundwater velocity and may eventually go to infinity
when the fluid temperature rise becomes constant after a certain test period.

The second approach is based on parameter estimation technique where


various selected parameters such as ground thermal conductivity, borehole thermal
resistance, ground volumetric heat capacity, etc. are to be determined simultaneously.
With a model for the borehole selected from LSM, CSM or any others, simulated
data is compared with the test data. An error term is defined based on the difference
between the simulated and the test data. The parameter estimation technique is then
employed for adjusting the selected parameters using optimisation method in order to

26
minimise the error term. There are various kinds of optimisation methods available.
Jain (1999) reviewed the characteristics of different methods and compared their
performance when applied in the analysis of TRT results. He commented that the
Nelder-Mead simplex method (Nelder and Mead, 1965) was one of best in terms of
rate of convergence and precision of method. This is a direct search method using
three basic operations (reflection, expansion and contraction) to optimise a function
with multi-dimensional domains.

Austin et al. (2000) adopted the short time step model developed by
Yavuzturk (1999) to analyse the TRT result using the Nelder-Mead simplex method.
They investigated the data from site and the results from Smith et al. (1999). They
also studied the sensitivity of parameters, and the effects of test period, far field
temperature, shank spacing, ground volumetric heat capacity, etc on the estimated
ground thermal conductivity. They commented that the TRT period should be at
least 50 hours. Gehlin and Hellstrom (2003a) used the Nelder-Mead simplex method
to compare the results of TRT analysis based on LSM (using Eqs. (2.8~10)),
simplified LSM (using Eq. (2.12)), CSM and a one-dimensional numerical model.
They also concluded that the test period should be at least 50 hours. They found that
the results obtained by using LSM and simplified LSM basically have no difference.
LSM predicted better than CSM and numerical model at the early stage of test. By
excluding results for the first 10-15 hours, CSM performed the best. If results for the
first 5-10 hours were excluded, the numerical model gave the nearest estimate.

Wagner and Clauser (2005) developed a numerical model for the borehole
and applied parameter estimation graphically using the ground thermal conductivity
and thermal diffusivity as parameters to minimise a misfit function. Shonder and
Beck (1999) derived a one-dimensional numerical model with the U-tube replaced by
a single pipe of equivalent diameter and applied parameter estimation using the
Gauss minimisation technique to investigate the TRT results from a laboratory setup.
Georgiev et al. (2006) analysed the TRT data of a shallow borehole using both
approaches. The model of Shonder and Beck (1999) was used in the second
approach and results compared. The estimated parameters were then applied to
simulate the performance of the borehole with periodic charging from a solar panel
and discharging using the TRNSYS DST model.

27
Researches on TRT analysis including groundwater effect were also made.
Gehlin and Hellstrom (2003b) developed three models to study the groundwater
effect on the borehole performance installed in hard rock under different assumptions
of groundwater flow conditions. At low groundwater velocity, the three models
differ mildly, while at high groundwater, the difference is prominent. They
presented graph showing the ratio of effective ground thermal conductivity to actual
thermal conductivity at different groundwater velocity based on the three models.
With the effective thermal conductivity determined from TRT and actual thermal
conductivity from laboratory test or other means, the groundwater velocity can be
estimated from the graph.

Katsura et al. (2006) applied the first approach to analyse the TRT results
performed using a laboratory setup at no groundwater flow in order to estimate the
effective soil thermal conductivity. They then used the estimated soil thermal
conductivity and adopt the LSMGA stated by Diao et al. (2004) to compare the
simulated data with the experimental TRT results under controlled groundwater
velocities. The differences were found to be small. They further proposed a method
to estimate the groundwater velocity from the TRT results based on known ground
thermal properties.

2.2 Liquid Desiccant Dehumidification Systems

2.2.1 General

The application of liquid desiccant dehumidification systems in commercial


installations arises from the need for a better indoor environment. Arundel et al.
(1992) investigated various adverse effects caused by bad control indoor relative
humidity. High relative humidity enhances the growth of bacteria and virus, which
endangers the health of occupants. The problem of molding is also embarrassing
(Wageningen et al., 1992). The risk is especially significant in locations such as
markets, kitchens, toilets, hospitals, etc. where the internal latent load and the
necessity for good hygiene are both high.

28
As noted by Griffiths (1992), the sole employment of conventional vapour
compression system for achieving a low sensible/latent load ratio usually required
sub-cooling to a very low temperature and reheating of the supply air. This implies
that the capacity of the air conditioning system should be higher than what is actually
necessary and sub-cooling to low temperature reduces the energy efficiency of the
system, resulting in a high operating cost. Liquid desiccant dehumidification
systems offer the advantages that the duty of the dehumidifier matches exactly the
requirement and that low-grade or even waste heat can be utilised, thus reducing the
operating cost.

The implementation of liquid desiccant dehumidifiers in air conditioning


systems is achieved in various ways. Griffiths (1992) described several
configurations of the liquid desiccant air conditioning systems incorporating other
auxiliary equipment such as cooling towers, heat pumps, chillers, diesel engine
generators, etc. Studak and Peterson (1992) compared the performance of a liquid
desiccant dehumidifier coupled with heat pump using different liquid desiccant
solutions. The condenser was used to reheat the weak desiccant solution before
entering the regenerator tower and the evaporator used to cool the strong desiccant
solution before entering the dehumidifier tower. They found that calcium chloride
gave the best overall performance based on various parameters.

Lazzarin and Castellotti (2007) analysed the performance of a modified heat-


pump-coupled liquid desiccant dehumidifier which was similar to that studied by
Studak and Peterson (1992) but with additional features to enhance the energy
efficiency of the unit for use in conjunction with conventional air handling unit
installed in a supermarket in Italy. The liquid desiccant dehumidifier designed for
treating the fresh air was mainly used for handling the latent load while the air
handling unit provided sensible cooling. They found that annual energy saving of
the new system was only mild, probably due to the fact that dehumidification was
only necessary for not more than two months in a year. They commented that the
benefits of the new system should be more prominent when applied to localities with
a humid climate condition.

29
Ghaddar et al. (2003) studied a solar-assisted liquid desiccant
dehumidification system installed in Beirut where the regenerative heat came partly
from the solar system and partly from a gas heater. A cooling tower is used to pre-
cool the strong desiccant solution before entering the dehumidifier tower. They
found that if no solar panel was used, the payback was immediate. With more solar
panels used, the payback period would be longer although the energy cost would be
lower. The regeneration temperature was about 45-50oC, which offered extra saving
potential if waste heat was available.

The application of liquid desiccant to air conditioning system can also take in
other form. Yuan and Alberce (2004) developed a cross-flow air-to-air total heat
recovery unit incorporating membranes soaked with desiccant solution placed
between the two air streams. Heat and moisture is transferred through the
membranes across the two air streams. They claimed that the total heat recovery
efficiency could be more than 80%, and was maintenance free as no moving part was
included. This was considered to be more beneficial as compared with a system
using a desiccant heat wheel.

2.2.2 Modelling of Liquid Desiccant Dehumidification Systems

The key item in a liquid desiccant dehumidifier is the packed tower. The
performance of the packed tower depends on the simultaneous heat and mass transfer
along the tower. There are mainly two modelling approaches. The first one is based
on coupled fundamental heat and mass transfer equations which are solved
numerically by using the finite difference method to determine the variation of states
in the air and desiccant stream along the tower. As counter-flow packed tower is
usually studied, a fully implicit scheme is used. Treybal (1969) applied this
approach and developed a design method for the packed tower which accounted for
the mass and heat transfer resistances of the liquid phase. Factor and Grossman
(1980) followed similar ideas to develop a model for the packed tower. They
compared the simulated data with the test results from an experimental setup using
lithium bromide solution and good agreement was found.

30
Lazaarin et al. (1999) also adopted the same method to derive a numerical
model for a packed tower and compared the computed data with the test results using
lithium bromide and calcium chloride solution, showing a maximum error of about
20%. They compared the performance of the packed tower in terms of humidity
reduction and tower efficiency at different solution-to-air mass flowrate ratios. Both
parameters increased with the mass flowrate ratio. They commented that maximum
humidity reduction was attained when the leaving air water vapour pressure was the
same as the saturated water vapour pressure of the entering desiccant solution. Fumo
and Goswami (2002) studied experimentally the dehumidification and regeneration
performance of a liquid desiccant packed tower using lithium chloride solution.
They investigated the dehumidification and regeneration rate at different air and
desiccant conditions, showing a linear correlation in most cases. They also
compared the test results with the numerical tower model.

Mago et al. (2006) applied the numerical tower model to investigate the
performance of a hybrid liquid desiccant system with an auxiliary process air cooler
after the dehumidifier tower. The tower performance agreed with experimental
results from previous research work. They also found that at constant desiccant inlet
temperature and incoming air condition, a higher dehumidification rate in the packed
tower would result in a lower capacity required for the auxiliary air cooler to meet
the same cooling demand. Khan and Ball (1992) derived the governing heat and
mass equations in a simplified way. They then proposed a generalised formula for
the leaving air humidity ratio which related to the entering air humidity ratio,
solution temperature and concentration, and determined the corresponding
coefficient from the simplified numerical model.

In the numerical model, the heat and mass transfer coefficient are important
parameters. Most workers adopted equation for the mass transfer coefficient with
different formulations for different packing types in various air and liquid mass flow
range and derived solution for the heat transfer coefficient based on the analogy
between heat and mass transfer. Gandhidasan et al. (1986) proposed simplified
correlations suitable for use within a wide range of air and liquid flow conditions.
Chung et al. (1996) proposed empirical equations for the heat and mass transfer
coefficients based on experimental data for random and structured packed tower

31
using lithium chloride solution. Their formulations did not follow the analogy
between heat and mass transfer. Elasrrag et al. (2004) conducted a similar study for
a structured packed column using triethylene glycol. They proposed alternative
equations for the mass transfer coefficient in terms of the Sherwood number based on
different solution-to-air mass flow ratio ranges. Analogy between heat and mass
transfer was then applied to evaluate the heat transfer coefficient.

The numerical model approach requires the solution to numerous coupled


finite difference equations which becomes very time consuming, especially if a
dynamic system simulation for long time period is performed. Hence, the
effectiveness approach is suggested. The overall heat and mass transfer in the
packed tower is described by a corresponding effectiveness. Chung (1994) derived
an empirical expression for the humidity ratio effectiveness based on experimental
results from a desiccant tower with random packing using triethylene glycol and
lithium chloride solution. Martin and Goswami (2000) proposed formulae for the
humidity ratio and enthalpy effectiveness from experimental data of a randomly
packed desiccant tower using triethylene glycol with an average error of 15%. Liu et
al. (2006) investigated the performance of a cross-flow tower, and deduced
correlations for the humidity ratio and enthalpy effectiveness from test data. Ren et
al. (2006) simplified the governing heat and mass transfer equations and developed
analytical solutions for the humidity ratio and enthalpy effectiveness of a desiccant
packed tower.

Elsarrag (2006) compared the simulation results from the effectiveness


formulae proposed by previous researchers with the test data for a structured packed
tower using triethylene glycol. He investigated the influence of various parameters
on the humidity ratio and enthalpy effectiveness. The numerical model was also
used for comparison. He remarked that the numerical model was more accurate.
Gandhidasan (2004) proposed a simplified model for the dehumidifier tower relating
the dehumidification rate with various parameters, but the formulation seemed to be
just a re-arrangement of basic equations, and the proof of the model using results
from previous workers (Fumo and Goswami, 2002) appeared to be misleading.

32
Elsayed (1994) studied the performance of a liquid desiccant
dehumidification system with desiccant cooler before the dehumidifier tower and
regenerative air heat exchanger and heater but without desiccant-to-desiccant heat
exchanger. He assumed fixed values for the humidity ratio and enthalpy
effectiveness for the dehumidifier and regenerator towers. He found that for normal
operation of the system, there was a minimum value for the desiccant to air mass
flow ratio which depended on the operating conditions. He also derived expression
for the COP of the system. He investigated the effects of various parameters on the
COP and concludes that for high COP, regenerative heater was not required if the
system operates at the minimum desiccant to air mass flow ratio, and the design mass
ratio should be close to the minimum value.

2.3 Vapour Compression Systems

The vapour compression system employing the reverse Carnot cycle is the
most common type of equipment used for air conditioning and refrigeration.
Gaseous refrigerant is compressed to high pressure and temperature through the
compressor before entering a condenser where the refrigerant is condensed, in most
cases, to become a sub-cooled liquid. The high pressure liquid refrigerant is then
expanded through an expansion device before entering the evaporator where it is
usually evaporated to a super-heated gas. ASHRAE (2004, 2006) described the
various aspects of vapour compression systems used for air conditioning and
refrigeration, and ASHRAE (2005) provided information of properties in the form of
tables and charts for a number of refrigerants commonly used.

According to Jin and Spitler (2002), modelling of vapour compression cycle


is performed in two approaches. The first one applies equation fitting to the entire
system from catalog data, which is easy to use but can be dangerous if extrapolation
beyond the catalog data is required. The second one involves a detailed description
of the individual components such as compressor, condenser, expansion device and
evaporator, etc. based on fundamental thermodynamic equations and the assumption
of conditions at different stages of the cycle such as fixed degree of sub-cooling at
condenser outlet, fixed degree of superheat at evaporator outlet, etc. Solution is

33
obtained by balancing the energy change and refrigerant mass flow through the cycle
iteratively until convergence is reached. This most popular approach allows
investigation of the effects of various parameters on performance of the vapour
compression system. In some cases, performance of individual component is
estimated by interpolation or equation fitting of the performance data from the
component suppliers.

Domanski and Didion (1984) studied the performance of an air-to-air heat


pump using a capillary tube as an expansion device. A polytropic process was
assumed for the compressor. The heat transfer in the condenser and evaporator was
calculated on a tube-by-tube basis employing detailed heat transfer equations. For
the capillary tube, a governing differential equation similar to the one proposed by
Stoecker and Jones (1982) was adopted which was solved numerically. No enthalpy
drop was assumed in the capillary tube. Besides energy and refrigerant flow, the
seemingly more reasonable refrigerant charge balance was also used in the cycle
iteration instead of a specified degree of sub-cooling. In calculating the refrigerant
charge in the condenser and evaporator within the saturated region of the refrigerant,
mean density was calculated on a tube-by-tube basis. No assumption was made on
the degree of superheat and sub-cooling leaving the evaporator and condenser.

Parise (1986) developed a model for water-to-water vapour compression heat


pumps using a thermostatic expansion valve which maintained a constant degree of
superheat at the compressor inlet. He also assumed a polytropic process in the
compressor. For the condenser and evaporator, he adopted a generalised heat
transfer coefficient to calculate the energy change based on the difference between
the arithmetic mean evaporating fluid and refrigerant temperatures across the coils.
The appropriateness of this approach was considered questionable. Specified degree
of sub-cooling was assumed without stating any justification for the cycle iteration.
The model was verified with experimental results.

Cecchini and Marchal (1991) followed a similar algorithm to estimate the


performance of a piece of vapour compression equipment suitable for any type of
condensing and evaporating fluid except that the calculation of heat transfer in the
condenser and evaporator was conducted in a more precise way. Formulae for the

34
heat transfer coefficients were adopted instead of just assigning a fixed value, and the
log mean temperature was used instead of the arithmetic mean temperature. For the
evaporator, different formulations were used depending on whether the evaporating
fluid was water or air. Condensation and frosting was accounted for if the
evaporating fluid was air. The model was also verified with test data.

Stefanuk et al. (1992) followed the same approach as Domanski and Didion
(1984) to simulate the performance of a water-to-water heat pump with thermostatic
expansion valve. Refrigerant charge balance was used instead of specified sub-
cooling at condenser outlet for cycle iteration. The calculation of heat transfer and
refrigerant charge in the coils within the saturated region of the refrigerant was
carried out in a numerical way by subdividing those parts of the coils into various
segments. The parameters used to model the compressor were estimated from the
suppliers performance data/curves by trial-and-error. Again, the model was
validated with experimental results.

Jin and Spitler (2002) also developed a model for a water-to-water heat pump
using thermostatic expansion valve. Isentropic compression was assumed, and
pressure drops at compressor inlet/outlet were accounted for. For the condenser and
evaporator, the refrigerant was assumed to be mostly in the saturated region. Hence,
the simple thermal effectiveness equation with fixed temperature on one side was
adopted. The refrigerant leaving the condenser was a saturated liquid with no sub-
cooling. Refrigerant charge balance was not applied. All design parameters
including coil effectiveness, compressor efficiency, etc. were then determined using
a parameter estimation technique from the system catalog data. No experimental
work or data for individual components were required, and the error level was still
satisfactory.

Jung et al. (1999) modified the model developed by Stoecker and Jones
(1982), including the effect of pressure drop due to sub-cooling and area contraction,
to determine the characteristic of capillary tube. They derived empirical equations to
relate the refrigerant mass flowrate with tube diameter, tube length, condensing
temperature and degree of sub-cooling suitable for different refrigerants with
different sets of coefficients.

35
Chapter 3 : Mathematical Formulations for
Geothermal and Liquid Desiccant Systems

3.1 Ground Heat Exchanger Borefield

In developing the numerical model for the ground heat exchanger borefield,
the following assumptions are adopted in order that comparison with the analytical
models can be made:

- The ground is homogeneous;


- The thermal properties of all materials remain constant within the
temperature range investigated;
- The boreholes and the ground have no contact resistance;
- The ground temperature remains unchanged at the top surface and at
distances far below the boreholes and away in the transverse directions;
- Thermal capacitance effects of the tubes and grouting inside the boreholes are
ignored so that the heat dissipated by the circulating fluid is the same as that
transferred from the boreholes to the surrounding ground;
- Fluid flow rate in each tube of boreholes is the same;
- Fluid flow rate in each borehole is the same;
- Groundwater flows in parallel to the ground surface with the same velocity
along the ground depth.

All subsequent analyses in this research will thus be based on the above
assumptions.

3.1.1 Heat Transfer Around Borehole

Figure 3.1 shows the grid scheme used for the ground around a borehole.
The implicit difference equation with regular grids can be obtained by direct

36
discretisation of the following governing differential equation for conductive heat
transfer in the ground with ground water advection:
Tg Tg Tg
g cg = k g 2Tg gw c gwV gw cos + sin + Qs (3.1)
t x y
Eq. (3.1) can be written in energy balance form with non-uniform grids as
Tgn,i+,1j ,m Tgn,i , j ,m Q x + + Q x + Q y + + Q y + Q z + + Q z + Qcx + Qcy + Qs q fact
= (3.2)
t g c g x j y m z i

where
Tgn,+i ,1j +1,m Tgn,i+,1j ,m
Q x + = k g y m z i
dx j

Tgn,+i ,1j 1,m Tgn,i+,1j ,m


Q x = k g y m z i
dx j 1

Tgn,+i ,1j ,m +1 Tgn,i+,1j ,m


Q y + = k g x j z i
dy m

Tgn,+i ,1j ,m 1 Tgn,+i ,1j ,m


Q y = k g x j z i
dy m 1

Tgn,+i +11, j ,m Tgn,+i ,1j ,m


Q z + = k g y m x j
dz i

Tgn,+i 11, j ,m Tgn,+i ,1j ,m


Q z = k g y m x j
dz i 1

Tgn,i+,1j 1,m Tgn,i+,1j +1,m


Qcx = gw c gwV gw cos x j y m z i
dx j 1 + dx j

Tgn,+i ,1j ,m 1 Tgn,+i ,1j ,m +1


Qcy = gw c gwV gw sin x j y m z i
dy m 1 + dy m
qfact is a load factor which will be introduced and determined in Section 6.1.1. The
source term Qs equals one quarter of the borehole load at each corner point of the
borehole and zero elsewhere.

37
Ground surface
y dx
d
x
Tg,i-1,j,m
Tg,i,j,m+1
Borehole

H zi
Tg,i,j,m dy
dzi
Tg.i,j-1,m Tg,i,j,m Tg,i,j+1,m y
Tg,i+1,j,m

Tg,i,j,m-1

Vgw

x
z

Figure 3.1 Discretisation scheme for the ground surrounding a borehole

3.1.2 Heat Transfer Inside Borehole

Tube u Tube v

Borehole
buv
z
Tg,i,j,m+1 Tg,i,j+1,m+1

Tg,i,j,m Tf,i,u Tf,i,

Ground Tb,i Tfl,i,u Tfl,i, zi


qb,i
Tg,i,j,m Tg,i,j+1,m
Tg,i+1,j,m Tf,i+1,u Tf,i+1,v
di

Figure 3.2 Discretisation scheme inside borehole

Figure 3.2 shows the grid scheme used for the boreholes. With fluid flowing
in the tubes, the fluid temperatures along the tubes are strongly coupled. Hence, all
grid fluid temperatures have to be solved simultaneously. However, assuming
conduction along the boreholes to be small compared with that in the transverse
direction when there is no fluid flowing in the tubes, the fluid temperatures along the

38
tubes not within the same control volume are only weakly coupled. Hence, the
numerical formulation is different. The borehole temperature (Tb,i) and loading (qb,i)
are defined at mid level of the control volume while the ground temperatures are
defined at the grid points. The borehole temperature is computed by averaging the
ground temperatures around the borehole surface. Hence, eight ground temperatures
are required to calculate the borehole temperature in each control volume, and that
the source term (Qs) at particular ground grid point on the borehole surface should be
determined from the borehole loadings above and below the ground grid point.

3.1.2.1 Flow Process

For a borehole containing nt tubes, the energy transfer at any tube u inside the
borehole can be determined from
dT fl ,u nt
T fl ,u T fl ,v T fl ,u Tb
(1) C f
m1
=
+
Cf = mf cf (3.3)
dz v =1
Ruv Ruu
v u

where m1=1 for downward flowing tubes


m1=2 for upward flowing tubes
By replacing the fluid temperatures with the grid point temperatures, eq. (3.3) can be
re-written as
T f ,i +1,u T f ,i ,u nt
(T f ,i ,u + T f ,i +1,u ) (T f ,i ,v + T f ,i +1,v )
(1) m1 C f =
dz i v =1
2 Ruv
vu

(T f ,i ,u + T f ,i +1,u ) 2Tb ,i
+
(3.4)
2 Ruu
Eq. (3.4) represents (nt x (nzbore -1)) coupled equations in (nt x nzbore) variables (Tf,i,u)
for one borehole by writing Eq. (3.4) consecutively for i=1 to (nzbore -1) for all
control volumes and u=1 to nt for all tubes, which are solved iteratively based on the
prescribed fluid inlet temperature, pipe connection configuration and borehole
temperatures. If all boreholes are connected in parallel, the borehole fluid inlet
temperature will be the borefield fluid entering temperature, and the borefield fluid
leaving temperature will be the average of all the borehole fluid outlet temperatures
which may be different depending on the relative positions of the boreholes in the
borefield. If some of the boreholes are connected in series, the fluid inlet and outlet

39
temperatures for the inter-connected boreholes will be coupled. The borehole
loading of each borehole is then calculated as
nt
T f ,i +1,u T f ,i ,u nt
(T f ,i ,u + T f ,i +1,u ) 2Tb ,i
qb ,i = (1) m1 C f =
(3.5)
u =1 dzi u =1 2 Ruu

The derivation of Ruv is given in the Appendix A. If all tubes are identical with

equal distance from the borehole centre, Ruu will be the same for all u, and Eq. (3.5)
can be re-written as
1 nt
nt Tb,i
qb,i = (T
u =1
f ,i ,u + T f ,i +1,u )
(3.6)
2 R11 R11
It should be noted that due to the difference in the borehole temperature profile, the
loading of each borehole in a borefield can still be different even if the borehole fluid
inlet temperature is the same.

Recalling Eq. (2.3) and defining the fluid temperature, borehole temperature
and borehole load as a weighted average along the borehole, then
~ ~
~ T f Tb
qb = ~ (3.7)
Rb
where
nzbore 1
qb ,i dz i ~ nt nzbore 1
(T f ,i ,u + f ,i +1,u )dz i ~
nzbore 1
Tb ,i dz i
q~b = Tf = Tb = (3.8)
i =1 H u =1 i =1 2nt H i =1 H

By combining Eqs. (3.6~8),



~ R
Rb = 11 (3.9)
nt
The weighted average accounts for the non-uniform grid spacings along the borehole,
which will simply become the arithmetic mean if the borehole is discretised into
equal segments. The borehole thermal resistance defined in this way will not be
affected by the temperature variation along the tubes and borehole surface, and the
pipe connection configuration inside borehole. This is contrary to the work by Zeng
et al. (2003) who defined the average borehole fluid temperature as the mean of the
inlet and outlet fluid temperature only. This approach is also commonly used by

40
other researchers as the outlet fluid temperature is the most important parameter to be
determined rather than the actual fluid temperature variation inside the borehole.

3.1.2.2 No Flow Process

When no fluid is flowing in the tubes, there is no directional dependence, and


the governing differential equation (assuming that the heat conduction in the vertical
direction along the borehole is small compared with that in the transverse direction)
becomes
dT fl ,u nt
T fl ,u T fl , v T fl ,u Tb
rpi2 f c f =
+
(3.10)
dt v =1 Rnf ,uv Rnf ,uu
vu

which can be discretised as


T fln,+i ,1u T fln,i ,u nt
T fln,+i ,1u T fln,+i ,1v T fln,+i ,1u Tbn,i
rpi2 f c f =
+
t v =1 Rnf ,uv Rnf ,uu
vu

nt
T fln,+i ,1u T fln,+i ,1v T fln,+i ,1u Tbn,i
Cf (T fln,+i ,1u T fln,i ,u ) =
+
(3.11)
v =1 Rnf ,uv Rnf ,uu
vu

rpi2 f c f
where Cf = (3.12)
t

The derivation of Rnf ,uv is also given in the Appendix A. Eq. (3.11) represents nt

equations in nt unknowns (T fln,+i1,u ) by writing Eq. (3.11) consecutively for u=1 to nt for

all tubes at the ith control volume. To solve T fln,+i1,u , Eq. (3.11) is re-written as

nt n +1
Tbn,i nt
T fln,+i ,1v 1 T fl ,i ,u
CfT n
fl , i , u +
=
+ + Cf (3.13)
Rnf ,uu v =1 Rnf ,uv v =1 R R
vu nf , uv nf ,uu
Eq. (3.13) can be expressed in matrix form as
n
CfT n + Tb ,i = R n +1
[ ]
*

nf ,uv T fl ,i ,v
fl ,i ,u
Rnf ,uu

where
* 1
Rnf ,uv = for u v and
R nf ,uv

41

* N 1
nf ,uu + Cf
R =
v =1 R
nf ,uv
Hence,

1
* n
Tbn,i
[
n +1
Rnf ,uv CfT fl ,i ,v + = T fl ,i ,u ] (3.14)
Rnf ,vv

All the borehole fluid temperatures will be solved by applying Eq. (3.14)
consecutively for each control volume of the boreholes in the borefield. Following a
similar transformation methodology applied to flow processes with identical tubes
and at the same distance from the borehole centre,

~ R
Rnf ,b = nf ,11 (3.15)
nt
Again, the borehole thermal resistance will be independent of the temperature
variation along the tubes and borehole surface, and the pipe connection configuration
inside the borehole. The weighted average of the fluid temperatures inside the
borehole will be used as the borehole fluid outlet temperature for the next fluid-
flowing time step.

3.1.3 Solution Methodology

The ground temperature around boreholes depends on the borehole loadings,


but the determination of the borehole loadings requires the borehole surface
temperature to be known. The two problems are coupled and have to be solved
iteratively at each time step to get the complete implicit solution. Gauss-Seidel
iteration with under-relaxation is adopted in the calculation of ground temperature
around boreholes and fluid temperature inside boreholes, i.e.
Tnew iteration step, final = Told iteration step + URF(Tnew iteration step Told iteration step) (3.16)
where URF is the under-relaxation factor. A small value of URF implies that the
temperature modifies very mildly in each iteration step, which enhances the stability
of the iteration process. This is necessary for small borehole sizes and/or high
groundwater velocities but will inevitably increase the computation time. The
iteration at each time step stops when all the ground and fluid temperatures converge
to a specified tolerance, i.e.

42
|Tnew iteration step Told iteration step| < Tolerance (3.17)

3.1.4 Borefield Performance Formulation

The inlet fluid temperature can be calculated as


Q AC
Tbfn +, 1f ,in = Tbfn , f ,out + (3.18)
m f ,bf c f

where QAC is the applied air conditioning load to the borefield. In practice, the
applied borefield loading varies with time or the system does not operate
continuously within a day. Hence, there is a transient response in the borefield
performance, usually resulting in a time delay. The load transferred to the borefield
is then evaluated as
Qbf = mbf , f c f (Tbfn +, 1f ,in Tbfn +. 1f ,out ) (3.19)

As already stated in Section 3.1.2.1, the representative fluid temperature used


in Eq. (2.3) is usually defined as the arithmetic mean of the borehole entering and
leaving fluid temperatures. By adopting this definition, the thermal resistance for a
borehole and the entire borefield can be defined as
~ ~
H (Tb , f ,in + Tb , f ,out 2Tb ) nbore H (Tbf , f ,in + Tbf , f , out 2Tbf )
Rb = Rbf = (3.20)
2Qb 2Qbf

The borefield thermal resistance indicates the overall performance of the entire
borefield while the borehole thermal resistance shows the situation for individual
borehole. Unlike Eqs. (3.9&15), the borefield and borehole thermal resistances will
depend on the temperature variation along the borehole surface and within the
borehole fluid. The borehole thermal resistance can also be different depending on
the relative position of the borehole in the borefield.

A borefield usually contains more than one borehole, and the effect of
groundwater on each borehole is different. The global effect on the entire borefield
depends on the position and connection configuration of the boreholes. In this way,
a change in the groundwater direction may result in a substantial difference in the
performance of the borefield under a specified groundwater flowrate, system
configuration and loading profile. To analyse the problem, a percentage variation

43
(PV) is defined based on the borefield leaving fluid temperature rise applicable to
sub-tropical regions under different groundwater directions as

PV = (Maximum leaving fluid temperature rise


- Minimum leaving fluid temperature rise) x100
/ Minimum leaving fluid temperature rise (3.21)
PV represents the maximum percentage difference in the borefield leaving fluid
temperature rise that can be obtained if the groundwater direction is wrongly
estimated.

3.1.5 Thermal Response Test Analysis

To estimate the various parameters namely ground thermal conductivity,


borehole thermal resistance, groundwater velocity, etc. from TRT results, an
optimisation tool is used to adjust the parameters in order to minimise a pre-defined
misfit function. In this research, the function to be minimised is the mean square
difference (MSD) between the measured data (usually the mean fluid temperature)
and the simulated values from the selected model and is defined as
ndata

(T f , simulated T f ,measured ) 2
MSD = 1
(3.22)
ndata

The Nelder-Mead simplex method, developed by Nelder and Mead (1965),


will be used as the optimisation tool for reason given in Section 6.2.2. With n
parameters to be determined, (n +1) test points in a n-dimensional solution space will
be used to determine the final results. Two criteria for completion of optimisation
procedure will be adopted. The first one is that when the MSDs of all the (n +1) test
points are below a preset value. The second one is that when the variation of the
MSDs is below a prescribed limit, i.e.
2( MSDmax MSDmin )
< Limit (3.23)
MSDmax + MSDmin
which is the same as that used by Jain (1999).

44
3.2 Liquid Desiccant Systems

According to Figure 1.4, a conventional liquid desiccant system comprises a


dehumidifier tower, regenerator tower, desiccant heater, desiccant-to-desiccant heat
exchanger, desiccant pumps and fans. In some cases, a desiccant cooler may also be
employed to cool the desiccant before it enters the dehumidifier tower. Assuming
very small duties for the pumps and fans, the change of states across the pumps and
fans can thus be neglected and will not be considered in the system model to be
developed.

3.2.1 Desiccant Packed Tower

TG,out TL,in
out in
G'out L'in

PACKED TG + dTG TL
TOWER + d
G' + dG' L'

dz
TG TL + dTL
+ d
z G' L' + dL'

TG,in TL,out
in out
G'in L'out

Figure 3.3 Discretisation scheme for a packed tower

Figure 3.3 indicates an elemental section of a vertical counter-flow liquid


desiccant packed tower utilising lithium chloride solution as the liquid desiccant.
According to Factor and Grossman (1980), the moisture transfer rate between air and
liquid desiccant depends on the corresponding partial pressures as

d M w FwG a 1 PwL / Pt
= ln (3.24)
dz G 1 PwG / Pt
dG = G d (3.25)
dL = dG (3.26)

45
dL
d = (3.27)
L + dL
The sensible heat transfer between air and liquid desiccant is given as
dTG hc ,G a (TG TL )
'

= (3.28)
dz G cG

dH G = G cG dTG + dG [c w (TG TL ) + ] (3.29)


By balancing the energy change between the air and liquid desiccant,
dH G + hd dL
dTL = (3.30)
L c L
The water vapour pressure in, and specific enthalpy of dilution of the liquid desiccant
solution can be calculated based on the empirical formulae proposed by Conde
(2004). Should lithium bromide solution be used as the liquid desiccant, the
formulae given by Patek and Klomfar (2006) can be employed.

Chung et al. (1996) proposed empirical correlations for determination of the


heat and mass transfer coefficients for lithium chloride liquid desiccant packed tower
with random packing as
M w d p2
= 1.326 x10 4 (1 ) 0.94 L Sc w0.333 Re1w.16
0.27

FwG a (3.31)
Dw w G

d p2
= 5.20 x10 5 (1 )1.56 L Prw0.333 Re1w.6
0.50

h a
'
(3.32)
c ,G
kw G

and for the structured packing case as
M w d eq2
= 2.25 x10 4 (1 ) 0.75 L Sc w0.333 Re1w.0
0.10

FwG a (3.33)
Dw w G

d eq2
= 2.78 x10 6 (1 )1.8 L Prw0.333 Re1w.6
0.40

h a
'
(3.34)
c ,G
kw G

Here,
w
Sc w = (3.35)
Dw w

d c wVG
Re w = (3.36)
w

46
cw w
Prw = (3.37)
kw
The change of states along the entire tower will be solved simultaneously by using an
iterative method for all segments along the tower. This approach is similar to the
calculation of the variation of fluid temperature inside a borehole in a flow process.

3.2.2 Desiccant-to-desiccant Heat Exchanger

Assuming no change in the thermal properties for the desiccant across the
heat exchanger and employing an average value for the overall heat transfer
coefficient along the heat exchanger, the temperature effectiveness of a counter-flow
heater exchanger can be determined using the approach by Bejan (1993) as
1 e [ NTU (1Cmin / Cmax )] UA
= NTU = (3.38)
1 (C min / C max )e [ NTU (1Cmin / Cmax )] C min
Eq. (3.38) basically applies to any two fluids with no change in physical states and
nearly constant heat capacities within the temperature range considered. It is
difficult to say if C will be higher for desiccant solution at dehumidifier tower outlet
or regenerator tower outlet. Indeed, they may differ only slightly. If C min C max ,
Eq. (3.38) becomes
NTU
= (3.39)
1 + NTU

3.2.3 Desiccant Heater/Cooler

The performance of a desiccant heater depends on the type of heat source


used such as by heat exchanger coil, solar, electric, etc. Even for the case of a heat
exchanger coil, the characteristic of the circulating fluid will also influence the
behaviour of the desiccant heater or cooler. Since refrigerant is to be used in this
study for the ground-coupled liquid desiccant air conditioner, detailed mathematical
modelling will be carried out in subsequent sections to incorporate the performance
of the condenser and evaporator of a heat pump. In general,
QL ,heat = C L (TL ,ri TL ,dhi ) (3.40)

QL ,cool = C L (TL ,dci TL ,di ) (3.41)

47
3.3 Vapour Compression Air Conditioning Systems

A vapour compression air conditioning system consists of various


components. For the refrigerant circuit, it mainly includes a compressor, a condenser
coil, an expansion valve and an evaporator coil. For the air side, it may comprise a
cooling coil, a heating coil (when operated in reverse mode), and an air-to-air heat
exchanger, etc. The modelling of each component may vary depending on the
medium handled by each component. In this analysis, a thermostatic expansion
valve is to be used for maintaining a constant degree of superheat at the compressor
inlet. No assumption on the degree of sub-cooling at the condenser outlet will be
made. The approach based on energy, refrigerant flow and refrigerant charge
balance as adopted by Domanski and Didion (1984) will be used for cycle iteration.
In calculating the refrigerant charge, only the condenser and evaporator (holding the
majority of the liquid refrigerant) will be considered. Pressure drops across the
condenser, evaporator and all connecting pipelines are neglected. Figure 3.4 denotes
the various states in a vapour compression cycle which will be used for analysis in
the subsequent sections.

T
sr,dis

suc dis compressor


Pr,cond dis co condenser
sr,co co ei expansion valve
sr,suc ei suc evaporator
Pr,evap
sr,ei

Figure 3.4 Notations used for vapour compression cycle modelling

48
3.3.1 Compressor

The precise process occurring inside a compressor depends on the type of


compressor used such as screw, scroll, reciprocating, centrifugal, etc. Modelling of
compressor performance can be very complex and many parameters are involved.
To avoid too complicated formulations, only the global behaviour of compressor will
be investigated. The simplest way is probably to employ the interpolation of
performance data from the compressor supplier (www.emersonclimate.com).
However, the main disadvantage of this method is that extrapolation of performance
data may lead to erroneous results. Another way is to approximate the compressor
performance by a particular thermodynamic process, and the equation of states
across the compressor derived from basic thermodynamic equations as adopted by
most previous workers. However, this method will not be used in this study, as the
simulated compressor performance based on this approach does not cohere with
manufacturers catalogue data. Detailed explanation will be given in Section 6.3.2.
Instead, empirical correlations for the refrigerant mass flowrate and compressor
power provided by the compressor manufacturer (www.emersonclimate.com) will be
used which are based on a specified degree of superheat equal to 5.56oC at the
compressor suction. The refrigerant discharge condition is determined from
Wcomp = mr (hr ,dis hr , suc ) (3.42)

3.3.2 Condenser or Desiccant Heater

In the condenser, the refrigerant can change from a superheated gas to a sub-
cooled liquid if sufficient cooling is provided and the refrigerant leaves the
condenser at least as a saturated mixture. However, there will be no change in the
composition of the condensing fluid if it is a liquid desiccant, and thus a constant
specific heat capacity can be assumed. As the refrigerant undergoes phase changes
during the cycle, the heat transfer coefficient will vary substantially within the coil
depending on the state of the refrigerant. Since an actual coil configuration is much
more complicated than just a single straight pipe, it is not easy to follow the
refrigerant path to determine the heat transfer between the refrigerant and the liquid
desiccant. Hence, instead of applying relevant formulae to evaluate the heat transfer

49
coefficient at different states of the refrigerant, an average value will again be
adopted throughout the coil. To calculate the global performance, the condenser coil
is divided into three parts, namely superheated, saturated and sub-cooled stage, as
shown in Figure 3.5.
T
Tr,dis

Trs,cond

Tr,co TL,ri
TL,dh2
TL,dhi
TL,dh1

0 x'cond,sc x'cond,sc+x'cond,sat 1

Figure 3.5 Change of state of refrigerant and condensing fluid across condenser

Here, x (0 < x < 1) is the proportion of coil for each stage, and l the characteristic
coil length where 0 refers to desiccant inlet and 1 the refrigerant inlet. Clearly,
, sc + xcond
xcond , sat + xcond
, sh = 1 (3.43)

To further simplify the analysis, average specific heat capacities are used for the
refrigerant along the superheated and sub-cooled regions.

In the superheated stage, Eq. (3.38) is applicable. Since the mass flowrate of
refrigerant is generally much lower than that for the liquid desiccant, the temperature
effectiveness for the superheated stage can be calculated as
[ NTU cond , sh (1C r , cond , sh / C L , dhi )]
1 e
cond , sh = [ NTU cond , sh (1 Cr , cond , sh / C L , dhi )]
(3.44)
1 (C r ,cond , sh / C L ,dhi )e

, shUAcond
xcond
where NTU cond , sh = (3.45)
C r ,cond , sh

The temperature change for the refrigerant and liquid desiccant can then be
determined as

50
Trs , cond = Tr , dis cond , sh (Tr , dis TL , dh 2 ) (3.46)

Cr , cond , sh (Tr , dis Trs , cond )


TL , ri = TL , dh 2 + (3.47)
CL , dhi

In the saturated stage, the refrigerant temperature remains constant except for
the zeotropic refrigerant mixtures like R407C. The temperature effectiveness
becomes
cond , sat = 1 e NTU cond , sat
(3.48)

, satUAcond
xcond
where NTU cond , sh = (3.49)
C L ,dhi

The temperature change for the liquid desiccant is calculated as


TL ,dh 2 = TL ,dh1 + cond , sat (Trs ,cond TL ,dh1 ) (3.50)

Should no sub-cooling occurs, the enthalpy change for the refrigerant is given by
C L ,dhi (TL ,dh 2 TL ,dh1 )
hr ,co = hr ,cond (3.51)
mr

with TL ,dh1 = TL ,dhi . The same condition will apply for Eq. (3.50).

In the sub-cooled stage, the coil performance is similar to that at the


superheated stage. Hence,
[ NTU cond , sc (1C r , cond , sc / C L , dhi )]
1 e
cond ,sc = [ NTU cond , sc (1 Cr , cond , sc / C L , dhi )]
(3.52)
1 (C r ,cond , sc / C L ,dhi )e

xcond , scUAcond
where NTU cond , sc = (3.53)
Cr , cond , sc

The temperature change for the refrigerant and liquid desiccant are evaluated as
Tr , co = Trs , cond cond , sc (Trs , cond TL , dhi ) (3.54)

Cr , cond , sc (Trs , cond Tr , co )


TL , dh1 = TL , dhi + (3.55)
CL , dhi

The complete coil performance is determined by solving Eqs. (3.43~55)


, sh ,
simultaneously. This requires an iterative method with initial guesses for xcond

, sat and xcond


xcond , sc respectively. The algorithm for the method is shown in

51
Appendix B. With all parameters estimated, the refrigerant charge in the condenser
can be calculated by
+ r ,co
, sat r ,cond , sat + x cond
M r ,cond = VOLcond xcond ,sc r ,cond (3.56)
2
The derivation of r ,cond , sat is given in Appendix C.

3.3.3 Expansion Device

As mentioned before, a thermostatic expansion valve is to be used in the


analysis. Assuming no enthalpy drop in the expansion valve,
hr ,co = hr ,ei (3.57)

and
DSH = Tr , suc Trs ,evap = Const. (3.58)

3.3.4 Evaporator or Desiccant Cooler

TL,dc1 TL,dci

TL,di Tr,suc

Trs,evap

0 x'evap,sat 1
Figure 3.6 Change of state of refrigerant and evaporating fluid across evaporator

The modelling of an evaporator is similar to that for a condenser except that


no sub-cooling of refrigerant is considered and the refrigerant is assumed to enter the
evaporator as a saturated mixture. The refrigerant may leave the evaporator as a

52
superheated gas if it is sufficiently heated. All other assumptions remain the same
unless the evaporating fluid is air, which will be dealt with in Section 3.3.8. The
calculation can be much simpler if there is no superheating of refrigerant, as direct
evaluation can be performed without any need for an iterative method. Figure 3.6
indicates the various states for the refrigerant and evaporating fluid (liquid desiccant)
in the evaporator. The derivations of the governing equations are similar to those for
condensers. It follows that
, sat + xevap
xevap , sh = 1 (3.59)

In the saturated stage, the temperature effectiveness is given by


cond , sat = 1 e
NTU evap , sat
(3.60)

, satUAevap
xevap
where NTU evap , sh = (3.61)
C L ,dci

The temperature change for the liquid desiccant is calculated as


TL ,di = TL ,dc1 evap , sat (TL ,dc1 Trs ,evap ) (3.62)

If no superheating occurs, the enthalpy change for the refrigerant is given by


C L ,dci (TL ,di TL ,dc1 )
hr , suc = hr ,ei (3.63)
mr

with TL ,dc1 = TL ,dci . The same condition will apply for Eq. (3.62).

In the superheated stage, the temperature effectiveness can be calculated as


[ NTU (1 C
1 e evap , sh r ,evap , sh /C
L , dci )]

evap, sh = [ NTU evap , sh (1 C r ,evap , sh / C L ,dci )]


(3.64)
1 (Cr , evap , sh / CL , dci )e

, shUAevap
xevap
where NTU evap , sh = (3.65)
C r ,evap , sh

The temperature change for the refrigerant and liquid desiccant can then be
determined by
Tr , suc = Trs ,evap + evap , sh (TL ,dci Trs ,evap ) (3.66)

C r ,evap , sh (Tr , suc Trs ,evap )


TL ,dc1 = TL ,dci (3.67)
C L ,dci

53
The evaporator performance can be determined by solving Eqs. (3.59~67).
As mentioned before, an iterative method may not be necessary. The algorithm for
the solution methodology is given in Appendix D. The refrigerant charge in the
evaporator is then given by
, sat r ,evap , sat
M r ,evap = VOLevap xevap (3.68)

3.3.5 Air-to-air Heat Exchanger

Tia,p+1,q

EA Tea,p,q Tea,p,q+1

q
Tia,p,q

IA
Figure 3.7 Discretisation scheme for air-to-air heat exchanger

An air-to-air heat exchanger usually appears as a cross-flow system in square


shape used to recover part of the heat or cold energy from the exhaust air to the
intake air. To simplify the analysis, the occurrence of condensation is not considered,
as in most applications, the temperature of the cold air stream is not much lower than
the dewpoint of the hot air stream. Detailed modelling has to be carried out
numerically by subdividing the heat exchanger into a number of tiny cells, as shown
in Figure 3.7. Again, an average overall heat transfer coefficient will be used.

Assume that both the intake air and exhaust air streams are discretised into
the same number of equal segments (naahxr ) . The heat transfer across each tiny cell
can be determined from

54
UAaahxr (Tea , p , q Tia , p , q )
Cia , aahxr (Tia , p +1, q Tia , p , q ) = Cea , aahxr (Tea , p , q Tea , p , q +1 ) = 2
(3.69)
naahxr
Eq. (3.69) is an explicit formulation of the finite difference equation, which is
appropriate for a cross-flow system. For a counter-flow system, an implicit scheme
must be used. The temperature variation inside the heat exchanger is obtained by
solving Eq. (3.69) for p=1 to naahxr and q=1 to naahxr . The newly estimated
temperatures are used to calculate the temperatures in the downstream cells. The
outlet temperatures for the intake and exhaust air streams will be given by
naahxr naahxr

Tia, p,q +1
p =1
T
q =1
ea , p +1, q

Tiao = Teao = (3.70)


naahxr naahxr

3.3.6 Air Heater

The model for an air heater is basically the same as a condenser by using air
as the condensing fluid. The much higher mass flowrate of condensing air as
compared with that for the refrigerant allows the same formulations to be applied
except by replacing liquid desiccant properties with air properties.

3.3.7 Ground-coupled Coil

For a geothermal heat pump, energy is transferred to the borefield circulating


fluid through the ground coupled coil, which can be a condenser or an evaporator
depending on the mode of the vapour compression cycle, i.e. heating or cooling. All
governing equations are the same as those for condensers and evaporators presented
previously except by replacing the liquid desiccant properties with appropriate fluid
properties, usually those of water.

3.3.8 Air Cooler

The modelling of an air cooler is more complicated than the desiccant cooler
due to the possibility of condensation of moisture from the air. Two approaches can
be used, the first one being an approximate iterative method which is similar to that

55
for a desiccant cooler but with additional iteration work required, and the second one
the finite difference method.

3.3.8.1 Approximate Iterative Method

To simplify the analysis, an average specific heat capacity is adopted for the
air along the saturation line. The enthalpy of the condensing water is neglected. In
this way, the air cooler can be considered to be composed of two partial coils (dry
and wet if sufficiently cooled) with similar formulations but having different air heat
capacities. The assumptions and governing equations for the partial coils are
basically the same as those for the desiccant cooler (Eqs.(3.59~68)) except that the
refrigerant may enter the coil as a superheated gas.

Figure 3.8 denotes the parameters used for the air cooler. x ac , wet is the

portion of wet air cooler which can be zero. The refrigerant entering the dry coil
may be saturated (Tr ,ac1 = Trs ,evap ) or superheated (Tr ,ac1 > Trs ,evap ) but the air enters

the wet coil at the dewpoint of air entering the air cooler (Tadp ,aci ) . The algorithm

for the air cooler is presented in Appendix E.

Ta,aci
Tadp,aci
Tr,suc
Tsa

Trs,evap Tr,ac1

0 x'ac,wet 1
Figure 3.8 Change of state of refrigerant and air across air cooler

56
3.3.8.2 Finite Difference Method

With the air cooler sub-divided into numerous segments (nac ) , the heat
transfer rate across each segment is calculated from
H ac = UAac (Ta Tr ) / nac (3.71)
The enthalpy changes in the air and refrigerant streams are given by
ha = H ac / ma and hr = H ac / mr (3.72)
The complete solution is obtained by writing Eqs. (3.71&72) consecutively for each
coil segments and solving them iteratively. Different formulations will be used for
the air enthalpy change in the dry and wet regions. Basically, the finite difference
approach can also be used to obtain solutions for the case of an air condenser.

3.3.9 Single Zone

Figure 3.9 shows a single-zone under air conditioning control. The zone
material is assumed to be a lumped mass, and no time delay is considered in the
response of the zone temperature and relative humidity to the corresponding heat
load and supply air condition. By balancing the energy and moisture across the zone
at one time interval,
Cap zone
Qsen = C sa (Trta Tsa ) + C zone (Trta Tzone ) C zone =
t
Qsen + C sa Tsa + C zoneTzone
Trta = (3.73)
C sa + C zone

Qlat M dzone ( rta zone )


= mdsa ( rta sa ) +
t
Qlat M dzone zone
+ mdsa sa +
rta = t (3.74)
M
mdsa + dzone
t
The zone condition will become the return air condition in the next time step.

57
SA RTA

Tzone, zone

Single Zone
Figure 3.9 Single zone with air conditioning

58
Chapter 4 : Laboratory Setup for Experimental
Thermal Response Tests

4.1 Thermal Response Test Rig

Figure 4.1 shows a schematic diagram of the thermal response test rig
designed for use in this study. A test tank of size 0.8m(W)x0.8m(D)x1.2m(H) and
made of galvanized steel is used to hold the soil for the test. A model borehole,
made of a PVC tube of diameter 34mm and length 200mm, is inserted into the test
tank at 0.4m above the tank bottom level. An electric heater and a PT100
temperature sensor of 150mm in length is inserted inside the model borehole and
backfilled with a thermal enhanced grouting. A voltmeter and current meter is used
to measure the power consumption of the electric heater. Due to the high design
rating (650W at 220VAC) of the electric heater, the supply voltage to the electric
heater is stepped down by using a variable transformer to about 24VAC and the
rating dropped to about 6W. The temperature sensor is connected to a transmitter,
powered by a 30VDC supply, to convert the temperature signal from 0~104oC to
1~5VDC signal through a 250-ohm high precision resistor for logging by a computer.
Two remote temperature logging buttons are used to record the circulating water
temperature and the test room temperature.

The test tank is filled with sand up to a height of about 0.8m from the tank
bottom. A water pool is maintained on top of the sand by circulating water from and
to the circulating tank of size 0.4m(W)x0.5m(L)x0.4m(H) by a submersible pump
inside the circulating tank. The height of the water level in the test tank can be
controlled by regulating the supply and return water valves. This provides a
hydraulic head to push the water downward into the soil, thus simulating a
groundwater flow across the model borehole, to a drain connection at the bottom
level of the test tank. A valve is installed at the drain outlet for controlling the
groundwater flowrate which is measured by a direct readout flowmeter. The model

59
groundwater is directed to the floor drain of the test room. Fresh tap water is fed into
the circulating tank to make up for the lost groundwater.

Supply pipe

Testroom temperature Circulating water


log button temperature log button

Overflow pipe

Return pipe
Soil
Circulating tank
From tap water

Groundwater
Constant Circulating flowmeter
temperature pump
water bath
To floor drain Test tank

Borehole
temperature
sensor

Borehole

Figure 4.1 Schematic diagram of thermal response test rig

Figure 4.2 Layout of thermal response test setup

60
As actual groundwater temperature at site tends to be constant, especially
below some depth from the ground surface, it is essential to maintain the circulating
water temperature within a narrow range. This is achieved by first passing the
circulating water through a constant-temperature water bath set at 23~24oC provided
by using a temperature sensor calibrating equipment before it enters the test tank. A
1500W electric heater, controlled by a built-in thermostat, is used to maintain the
required room temperature for tests conducted in winter. To further minimise the
effect of room temperature variation on the soil temperature, the test tank and the
circulating tank are wrapped with thermal insulation. A top cover is provided to the
test tank. A photo of the entire test rig inside the test room is shown in Figure 4.2.

4.2 Soil Thermal Conductivity Measurement Setup

Copper plates
Thermal insulation

Soil
Plate type heater

Cooling fan

Acrylic frame

Figure 4.3 Details of experimental setup for measuring soil thermal conductivity

Fig. 4.3 shows the details of the experimental setup used for measuring the
soil thermal conductivity. An acrylic frame of internal size 140mmx140mm and
thickness 6mm is used to house a lump of wet soil sandwiched by two copper plates.
A plate-type electric heater of capacity 125W at 120VAC is stuck on the exterior
surface of one copper plate, and the exterior surface of the other copper plate open to
the surrounding with a cooling fan placed to enhance the heat transfer in order to
reduce the temperature rise in the soil. The temperature difference across the soil is

61
measured by two thermocouples placed on each exterior surface of copper plates. To
ensure that most of the heat generated by the heater be transferred to the soil, the
acrylic frame and the heater side is covered with a layer of thermal insulation. A
variable transformer is used to control the heater output, and the power consumption
measured by a high precision power meter. A photo of the complete experimental
setup is shown in Figure 4.4.

Figure 4.4 Layout of experimental setup for soil conductivity measurement

4.3 Calibrating Equipment

All measuring instrument need to be calibrated before use. For the


temperature measuring devices including borehole temperature logger, room and
circulating water temperature logging buttons, thermocouples for the soil
conductivity measurement, a temperature calibrator as shown in Figure 4.5 is used,
which can maintain the temperature of a testing water pool within 0.1oC accuracy.
The borehole temperature logger and the logging buttons will be calibrated from 20

62
to 40oC with 1oC interval while the thermocouples will be calibrated from 30 to 50oC
with 1oC interval.

Figure 4.5 Temperature calibrator

The voltmeter and ammeter used to measure power consumption of the


electric heater are calibrated against a high-precision power meter, a schematic
diagram of which is shown in Figure 4.6. A variable transformer is used to adjust the
voltage input applied to a dummy load in the form of a variable resistor. Different
voltages and currents across the dummy load, generated by modulating the
transformer and resistor separately, will be measured both by the high-precision
power meter and the measuring devices to be used in the test for comparison
purposes. Figure 4.7 shows the layout of the calibration setup.

The groundwater flowmeter will not be calibrated. Instead, the groundwater


flowrate will be calculated by measuring the groundwater flow volume within a time
period (1.5~2 minutes) after each test using a measuring cylinder. All calibration
curves are shown in Appendix I.

63
High
precision
power meter load

Figure 4.6 Schematic diagram of calibration setup of voltmeter and ammeter

(a) (b)
Figure 4.7 Layout of calibration setup for voltmeter and ammeter: (a) front view, (b)
rear view

The precision of all the measuring devices results in some uncertainties in the
measured or logged data. For the borehole temperature logger, a possible noise of
0.104oC is found, which will affect the logged borehole temperature profile and the
determination of the initial or farfield ground temperature (T0). The voltmeter and
the ammeter used for measuring the borehole loading (Qb) have a precision of
0.01V and 0.01A respectively, which accounts for an uncertainty in the borehole
loading of 0.0001W. The effect of uncertainties in T0 and Qb on the thermal
response test analysis results will be given in Section 6.2.3.4. The thermocouples
used in the soil thermal conductivity measurement test have a precision of 0.1oC.

64
The resulting uncertainty in the estimated soil thermal conductivity (kg) depends on
the actual kg and the heat flux through the soil. Based on a heat load of 20W passing
through the soil of thermal conductivity 3.0W/mK, the uncertainty in kg will range
from -0.268 to +0.326W/mK.

65
Chapter 5 : Methodologies for Simulation and
Analysis

5.1 Study of Ground Heat Exchanger Borefield Performance

To apply the numerical model developed in Section 3.1, the setting of the
grids and the load factor are to be determined first by comparing the simulated
ground temperatures with those using analytical models. This is essential in order
that the accuracy of the numerical scheme can be maintained without becoming
computationally expensive. Once the grid system is finalised, borefield performance
simulation can then be conducted both with and without groundwater advection.
Comparisons are also made with relevant analytical models. The effect of loading
profiles and groundwater conditions on the borefield performance will also be
investigated. Four loading profiles will be analysed:
1. Continuous constant load;
2. Constant load with a daily operating schedule of twelve hours;
3. Annual periodic load (approximated by a sine curve) with a zero minimum load
and daily operating schedule;
4. Annual periodic load with heating period (peak cooling load/peak heating load =
4) and daily operating schedule.
Unless otherwise specified, the peak cooling load for each loading profile will be
assumed to be the same. The annual heat injection to the ground corresponding to
Loading Profiles 2, 3 and 4 will then be 0.5, 0.25 and 0.1875 times that under
Loading Profile 1 respectively.

The analytical models usually require the evaluation of integrals. To reduce


computation time, suitable simplifications need to be made. When applying CSM,
Eqs. (2.24&25) will be used. The calculation of the improper integral will be
simplified by replacing the lower integration limit by 10-13, and the upper limit by
100 in view of the variation of the integrand at different . A fixed step of 0.00001
will be applied for the numerical integration. With FLSM, Eq. (2.33) will be adopted.

66
The determination of the definite integrals will be accomplished as indicated in
Appendix F. When determining the g-function, a weighted average is used instead of
a double integral using the same discretisation method as that used for the numerical
model. The thermal interference between adjacent boreholes will also be estimated
in a similar way. For LSMGA, Eqs. (2.37&38) will be used to calculate the average
borehole temperature with simplification made as shown in Appendix G. When
evaluating the thermal interference effect, Eq. (2.35) will be applied. The estimation
of the definite integral follows the approach presented in Appendix F. When dealing
with FLSMGA, the average borehole temperature is determined based on Eqs.
(2.42~45). The thermal interference effect is computed as shown in Appendix H.
Again, a weighted average for the ground temperature within the effective borehole
depth will be adopted.

Table 5.1 Default values of parameters used for simulation of borefield performance

Parameter Value Parameter Value


Insulated depth of borehole 5 Effective ground thermal 3.5
below ground surface, d (m) conductivity, kg (W/mK)
Effective length of borehole, 110 Effective ground thermal 1.62x10-6
H (m) diffusivity, ag (m2/s)
Borehole radius, rb (m) 0.055 Fluid heat capacity, cf 4,190
(J/kgK)
U-tube outer radius, rpo (m) 0.016 Fluid density, f (kg/m3) 1,000
U-tube inner radius, rpi (m) 0.013 Fluid thermal conductivity, kf 0.614
(W/mK)
Distance for tube centre from 0.03 Fluid dynamic viscosity, f 0.00086
borehole centre, di (m) (kg/ms)
Number of discretisation 20 Pipe thermal conductivity, kp 0.4
segments along borehole (W/mK)
Number of tubes inside 2 Grout thermal conductivity, 1.3
borehole, nt kb (W/mK)
Fluid mass flowrate inside 0.2 Peak applied load, QAC (W) 4,000
tube, mf (kg/s)

67
Unless otherwise specified, the values given in Table 5.1 will be used in the
analysis. The effective ground thermal properties refer to those for wet soil, but the
groundwater velocity may be zero. This is to avoid a sudden discontinuity in the
ground thermal properties between dry and wet soil when the groundwater velocity
approaches zero for ease of comparison between the borefield performance with and
without groundwater advection. All boreholes are assumed to be connected in
parallel. The borefield load refers to the total loading handled by the borefield, but
the loadings for individual boreholes will be different in most cases.

5.2 Thermal Response Test Analysis

Before performing the TRT analysis, besides the groundwater velocity, the
number of independent parameters to be estimated simultaneously has to be finalised
first. A sensitivity analysis will be made to select suitable parameters for analysis.
Once the parameter list is set, analysis can be conducted. In applying the parameter
estimation, initial guess points have to be assigned. For ease of generating the initial
guess points, each parameter is given a range with presumed minimum and
maximum values. This will not actually limit the search range in the simplex method
and the final result can be outside this range. The first guess point is selected
randomly within 30~50% of the corresponding parameter range for each parameter.
The second guess point will be the same as the first guess point except that the first
parameter will be increased by 20% of the first parameter range. The third guess
point differs from the first guess point by increasing the second parameter by 20% of
the second parameter range. The remaining guess points can be obtained by
following the same method.

In each analysis, three trials will be made. This allows the elimination of pre-
mature completion according to the second completion criterion as stated in Eq.
(3.23) where the MSDs of all the test points may be very close to each other but
none of them are the optimum points. Moreover, the confidence of the parameter
estimation method in determining the groundwater velocity can be investigated. This
is achieved by first analysing the test data generated from the same model but
excluding groundwater velocity in the parameter list. The result is then compared

68
with that analysed with groundwater velocity in the parameter list. Three models
will be applied, namely LSMGA, FLSMGA and the present numerical model. The
effect of the test period on the sensitivity of estimated groundwater velocity will also
be studied.

Analysis will then be made on the TRT results from an experimental setup
with controlled groundwater flowrate in order to verify the method. Regarding the
diameter and length of the experimental borehole with a length-to-diameter ratio of
less than 7, only the present numerical model will be used. Different groundwater
velocities will be tested and the estimated groundwater velocities compared with the
test values. The estimated soil thermal conductivity will also be compared with that
determined by laboratory method. Finally, the new parameter estimation method
including groundwater velocity will be used to analyse site TRT results provided by
others to investigate the possible outcome.

5.3 Modelling of Ground-coupled Liquid Desiccant Air Conditioners

Before the performance of ground-coupled liquid desiccant air conditioners


can be investigated, the configuration of the systems has to be finalised first. For the
dehumidifying cycle, the concern will not be just the dehumidification rate, but also
the enthalpy change of the process air across the dehumidifier tower. The effect of
each component in the dehumidifying cycle in terms of these two parameters will be
analysed to decide the ultimate arrangement of the dehumidifying cycle. The
parameter values shown in Table 5.2 will be used to determine the final
configuration of the liquid desiccant dehumidification circuit. In the peak-load
period, the design indoor temperature is lower and drier than the ambient. This
requires a larger temperature difference between the desiccant solution entering the
dehumidifier and the regenerator tower if the process air is room air and the
regenerative air is ambient air. As a heat pump will be used to provide desiccant
cooling and heating in the present study, the COP of the heat pump will be reduced.
In view of this, the liquid desiccant cycle will only be used to treat the fresh air, as
indicated by the same process and regenerative air entering conditions in Table 5.2.
The outdoor condition selected in Table 5.2 (31oC/83%RH) is the one with the

69
largest humidity ratio (and hence water vapour pressure in the air) from 1986 to 1995
based on the weather data from the Hong Kong Observatory. This condition will
also be used in the design of the ground-coupled liquid desiccant air conditioning
system. Random packing will be assumed throughout the analysis.

Table 5.2 Default parameter values used for determination of final configuration of
liquid desiccant dehumidifier

Height of dehumidifier 0.6 Height of regenerator tower, 0.6


tower, Zdeh (m) Zreg (m)
No. of discretisation segment 50 No. of discretisation 50
for dehumidifier tower, NZ,deh segment for regenerator
tower, NZ,reg
Cross-sectional area of 0.16 Cross-sectional area of 0.16
dehumidifier tower, Sdeh (m2) regenerator tower, Sreg (m2)
Packing size for dehumidifier 0.02 Packing size for regenerator 0.02
tower, dp,deh (m) tower, dp,reg (m)
Process air entering 31 / 84 Regenerative air entering 31 / 84
conditions, Tpa,di / RHpa,di conditions, Tra,di / RHra,di
(oC/%) (oC/%)
Process air volume flowrate, 0.2 Regenerative air volume 0.2
V&pa (m3/s) flowrate, V&ra (m3/s)

Overall heat transfer 1.0 Total system pressure of 101.325


coefficient of desiccant heat dehumidifier, Pt (Pa)
exchanger, UAL,he (kW/K)
No. of discretisation segment 50 Diffusion coefficient of 2.56x10-5
for regenerative air heat water pressure in air, Dw
exchanger, Nra,he (m2/s)
Liquid desiccant (LiCl) 0.00015 Desiccant heating load, 5,500
solution volume flowrate, QL,heat (W)
V&L (m3/s)

With the configuration for the liquid desiccant dehumidifier fixed, a heat
pump will be incorporated into the system, and the performance analysed in terms of

70
the COP for which only the compressor power is considered. Comparison will be
made with the conventional air-cooled fresh air unit (ACFAU) based on the same
coil properties, air flow and refrigerant charge under different outdoor conditions for
1986. Extra design parameters to be used are given in Table 5.3. Scroll compressor
(Model ZH15K4E) and R134a refrigerant will be used in the analysis.

Table 5.3 Additional parameter values used for analysis of heat-pump-coupled liquid
desiccant dehumidifier

Overall heat transfer 0.8 Overall heat transfer 0.8


coefficient of desiccant coefficient of desiccant
cooler/evaporator, UAL,dc heater/condenser, UAL,dh
(kW/K) (kW/K)
Volume of refrigerant in 0.003 Volume of refrigerant in 0.003
desiccant cooler/evaporator, desiccant heater/condenser,
VOLdc (m3) VOLdh (m3)
Volume of refrigerant in 0.0001 Refrigerant charge, Mr (kg) 0.6
3
liquid line, VOLll (m )
Liquid desiccant (LiCl) 0.00012 Degree of superheat at 5.56
solution volume flowrate, compressor suction, DSH
V&L (m3/s) (K)

Finally, a ground-coupled liquid desiccant air conditioner will be devised.


The dynamic performance of the new system operating under Hong Kong weather
conditions (from 1986 to 1995) will be simulated and compared with that using the
conventional geothermal heat pump system. Similar coil and refrigerant parameters
will be used in both systems and other parameters selected so that the capacity of
both systems will be almost the same. The corresponding borehole length based on
same maximum borefield fluid leaving temperature within the simulated time period
and the compressor power input will be investigated to justify the benefit of the new
hybrid system for application in sub-tropical regions. A simple single-zone building
will be used for analysis, with details shown in Table 5.4. Two cases will be
considered, the first one with ten occupants and second case with twenty occupants,
which allows the effect of fresh air ratio to be investigated. No infiltration will be

71
considered in the building. The TRNSYS software package will be used to generate
a room loading profile for simulation. The building thermal capacitance indicated in
Table 5.4 includes the effect of building material and furniture as suggested by
Bradley (2007).

Table 5.4 Details of single zone building for dynamic simulation

Building size (W x D x H) (m) 10 x 10 x 4


Building thermal capacitance (kJ/K) 4,800
Lighting load (W/m2) 5
Internal gain, radiative (kW) 1
Internal gain, convective (kW) 1
o
Cooling set point (db/RH) ( C / %) 24 / 54
Heating set point (oC) 20
Occupant activity Seating, light work

72
Chapter 6 : Results and Discussions

6.1 Study of Ground Heat Exchanger Borefield Performance

6.1.1 Setting of Ground Grids and Load Factor

Before any simulation could be carried out, the grid sizes had to be carefully
set so that accuracy could be maintained without use of excessive computation time.
An analytical model was to be chosen for calibration. CSM was selected since it was
an exact solution under the presumed boundary conditions. As CSM excluded
groundwater effect, the calibration was performed for zero groundwater velocity.

Before setting grid sizes, the boundary of the borefield in vertical and
transverse directions had to be determined first. According to Eskilson (1987), the
distance travelled by a heat front from the heat source after time t could be estimated
as r = 3 a g t . By defining a maximum time (tmax) for analysis which was assumed

to be one hundred years, a boundary distance (rmax) could be calculated as


rmax = 3 a g t max . This boundary distance was the margin from the outmost

boreholes in the transverse directions and from the bottom of boreholes in vertical
direction where the ground was discretised.

To set the grid sizes in the transverse directions, Eq. (3.2) was simplified to a
two-dimensional case for a single borehole of infinite length, with the results
compared with those from CSM based on a constant continuous load along the
borehole for ten years. The grids and the load factor were adjusted so that the
calculated ground temperatures along borehole and those at particular distances from
the boreholes would be very close to those obtained by using CSM. Upon repeated
trial and errors, the first grid on the borehole surface would be rb / 2 from the
borehole centre. The second, third and fourth grid would be 3rb, 5rb and 7rb from the
borehole centre respectively. Beyond that, the grid spacing would be 1.65 times the
preceding grid spacing. This grid scheme applied to both transverse directions. The

73
load factor was set to 1.047. A comparison of ground temperatures with CSM based
on the above grid system and load factor was shown in Table. 6.1 based on a
constant borehole loading of 30W/m.

Table 6.1 Comparison of present model with CSM

Distance Ground temperature Borehole temperature rise


Time
from rise after 10 years (K) (K)
(year)
borehole (m) Present CSM Present CSM
0.055 8.75 8.75 1 7.18 7.18
0.165 7.38 7.26 2 7.66 7.66
0.385 6.17 6.11 3 7.94 7.93
0.825 5.11 5.07 4 8.13 8.13
1.705 4.10 4.08 5 8.29 8.28
3.465 3.12 3.11 6 8.41 8.41
6.985 2.16 2.17 7 8.51 8.51
14.03 1.25 1.27 8 8.61 8.60
28.11 0.48 0.49 9 8.69 8.68
56.26 0.06 0.06 10 8.75 8.75

With the transverse grids fixed, the vertical grids were to be set. The method
used by Eskilson (1987) was adopted for depths within and below the boreholes.
The number of discretisation segments along the borehole was chosen so that the
minimum segment length along the borehole would be around one metre. For depths
above the boreholes, the grid spacing from top of boreholes would be the minimum
segment length along the borehole as determined by Eskilson (1987). With the entire
grid scheme fixed, a comparison of ground temperatures was made with FLSM based
on a continuous constant load of 30W/m along borehole, as shown in Table 6.2. The
simulated results between the two models were quite close. Comparison was also
made between the simulated fluid leaving temperature from a single borehole using
Eq. (3.4) by adopting the finalised borehole discretisation scheme with the analytical
solution stated by Zeng (2002) based on a single U-tube inside the borehole and a
constant borehole temperature. The results agreed well with each other.

74
Table 6.2 Comparison of present model with FLSM

Distance Temperature rise at mid level Average borehole


from of borehole after 10 years (K) Time temperature rise (K)
borehole (m) Present FLSM (year) Present FLSM
0.055 8.73 8.73 1 7.07 7.08
0.165 7.29 7.24 2 7.50 7.50
0.385 6.14 6.08 3 7.73 7.74
0.825 5.08 5.04 4 7.90 7.90
1.705 4.08 4.05 5 8.02 8.03
3.465 3.09 3.09 6 8.12 8.12
6.985 2.14 2.14 7 8.20 8.20
14.03 1.23 1.24 8 8.26 8.27
28.11 0.48 0.48 9 8.32 8.34
56.26 0.06 0.06 10 8.37 8.38

With the inclusion of groundwater effect, the numerical model for the two-
dimensional case was then compared with LSMGA based on a loading 30W/m after
10 years, as shown in Table 6.3. It could be noted that the differences were small
and no more than 2.8%. The complete three-dimensional model with groundwater
effect was finally compared with FLSMGA as shown in Table 6.4 using the
parameters indicated in Table 5.1.

Table 6.3 Comparison of present model with LSMGA

Average borehole Average borehole


Darcy velocity Darcy velocity
temperature rise (K) temperature rise (K)
(10-7m/s) (10-7m/s)
LSMGA Present LSMGA Present
0 8.76 8.75 10 4.82 4.80
1 7.91 7.92 20 3.88 3.86
2 7.01 7.01 50 2.66 2.67
5 5.76 5.75 100 1.81 1.85

75
Table 6.4 Comparison of present model with FLSMGA

Average borehole Average borehole


Darcy velocity Darcy velocity
temperature rise (K) temperature rise (K)
(10-7m/s) (10-7m/s)
FLSMGA Present FLSMGA Present
0 10.15 10.15 10 5.81 5.78
1 9.31 9.32 20 4.68 4.65
2 8.35 8.35 50 3.22 3.20
5 6.92 6.91 100 2.18 2.19

6.1.2 Performance of Borefield without Groundwater

The previous comparisons made with analytical models were based on a


constant load along the boreholes. In real situations, the fluid temperature varied
along the U-tubes in the boreholes, and the coupling effect with the ground could be
complex. The present numerical model provided a solution to such situations.

6.1.2.1 Temperature and Loading Variation along Single Borehole

Figures 6.1&2 showed the temperature rise and loading along single borehole
at different times against a dimensionless parameter (z-d)/H which assumed values of
one and zero at the borehole bottom and top respectively. It was found that neither
the temperature nor the loading was constant along the borehole. The borehole
temperature reached a maximum near the top part of borehole rather than near the
middle level of borehole if a finite line source model was used. The borehole
loading decreased with depth up to the bottom end of borehole. This could be
explained by considering the fact that the mean fluid temperature inside the borehole
decreased with depth as shown in Figure 6.3. Hence, near the top of borehole where
the borehole temperature increased with depth, the borehole loading would decrease
due to reduced temperature difference between the fluid and the borehole. The
situation changed only beyond the depth where the borehole temperature decreased
with depth at a higher rate than the mean fluid temperature near the lower part of
borehole. Basically, the borehole temperature and loading profiles changed very

76
mildly with time. Indeed, very similar profiles were obtained when using a
cylindrical coordinate system for a single borehole.

Borehole temperature rise variation Borehole loading variation

Borehole temperature rise (K) Borehole loading (W/m)


6 7 8 9 10 11 30 35 40 45 50 55 60
0 0

0.2 0.2
1 year 1 year

(z-d)/H
(z-d)/H

0.4 2 years 0.4 2 years


0.6 5 years 0.6 5 years
10 years 10 years
0.8 0.8

1 1

Figure 6.1 Temperature rise along single Figure 6.2 Loading along single borehole
borehole without groundwater flow at different without groundwater flow at different times
times

Comparison of fluid and borehole temperature rise after


1 year

16
Temperature rise (K)

14
Down flow fluid
12
Up flow fluid
10
Mean fluid
8
Borehole
6
4
0 0.2 0.4 0.6 0.8 1
(z-d)/H

Figure 6.3 Comparison of fluid and borehole temperature rise in single borehole
without groundwater flow after 1 year

6.1.2.2 Ground Temperature Profiles around a Single Borehole

Figures 6.4&6.5 showed the ground temperature profiles at various distances


from a borehole after five and ten years respectively against the dimensionless length
( z d ) / H . The reason for just indicating the ground temperatures within this depth
range was that the thermal interferences from adjacent boreholes were accounted for
only within this region. As the dimensionless length was linear in z, the shapes of
the ground temperature profiles were maintained. The temperature profiles at the
borehole were different from those away from borehole. At distances far from the
borehole, the depth with maximum temperature shifted to mid-level of the borehole.
This meant that the thermal interference effect between two boreholes in a borefield
would depend both on the borehole spacing and the depth which consequently
influenced the fluid temperature variation inside the boreholes. The resulting

77
borehole temperature profile and thermal resistance of each borehole could thus be
different. The present numerical method accounted for the varying coupling effect
between the ground and the borehole fluid temperatures along the boreholes. Hence,
individual finite difference scheme would be needed for each borehole for a precise
simulation when the method of superposition was used to determine the performance
of a borefield. A simpler way would be to discretise the entire borefield, and
perform the simulation for all boreholes simultaneously. This was easily achieved
when using a rectangular coordinate system by expanding the original grid system in
transverse directions to cover the entire borefield.

Ground temperature profile after 5 years Ground temperature profile after 10 years

Ground temperature rise (K) Ground temperature rise (K)


0 2 4 6 8 10 12 0 2 4 6 8 10 12
0 0

0.2 Borehole 0.2 Borehole


1m 1m
(z-d)/H
(z-d)/H

0.4 0.4
2m 2m
0.6 0.6
5m 5m
0.8 10m 0.8 10m

1 1

Figure 6.4 Ground temperature profiles around Figure 6.5 Ground temperature profiles around
single borehole without groundwater flow after 5 single borehole without groundwater flow after 10
years years

6.1.2.3 Comparison with the Finite Line Source Model

Table 6.5 Comparison of average ground temperature rise between present model
and FLSM after 10 years

Distance from Average ground temperature rise (K)


borehole Numerical FLSM
Borehole surface 10.10 10.15
1m 5.29 5.38
2m 4.26 4.25
5m 2.77 2.79
10m 1.55 1.74

Table 6.5 compared the mean ground temperature rise within the borehole
depth at various distances from the borehole calculated using the present simulation

78
tool and the finite line source model. It could be observed that at a distance of 5m or
less, the differences between numerical and analytical results were small and less
than 2%. However, at 10m from borehole, the situation changed with a resultant
difference of up to 12.5%, contrary to a value of less 1% for ground temperature at
mid borehole level as shown in Table 6.2 which assumed a constant borehole load
profile. This could lead to substantial deviation in estimating the performance of a
large borefield when considering the thermal interference effects from distant
boreholes. To verify this, the performance of various borefields with a borehole
spacing of 5m was simulated using the present finite difference scheme and the finite
line source model, as shown in Table 6.6. For a small borefield of up to 3x3
boreholes, the percentage difference was no more than 2.1. However, for a 5x5
borefield, the percentage difference rose up to more than 5. It could be anticipated
that for a larger borefield of more than 10x10 boreholes, the percentage difference
could be more than 10%. The present model was considered to be more accurate as
it accounted for the variation of temperature and loading along the boreholes which
occurred in real-life situations.

Table 6.6 Comparison of average borehole temperature rise between present model
and FLSM after 10 years for various borefields

Average borehole temperature rise (K)


Borefield Numerical FLSM
single 10.10 10.15
2x2 17.90 18.05
3x3 26.87 27.40
5x5 44.43 46.98

6.1.2.4 Borefield Thermal Resistance Variation

Table 6.7 showed the variation of borefield thermal resistance (Rbf) according
to Eq. (3.20) with time for various borefield configurations. For a single borehole,
the thermal resistance changed very little with time. This could be explained by
considering the fact that the borehole temperature profile, which affected the thermal
resistance, did not change much with time as already shown in Figure 6.1. For a

79
borefield with more than one borehole, the thermal interference from adjacent
boreholes depended on distance, depth and time. This caused the borefield thermal
resistance to vary more with time, especially for larger borefields. When using
FLSM based on an average borehole temperature along borehole, the thermal
resistance was fixed at around 0.1253. This deviation of thermal resistance between
the present numerical method and FLSM would lead to an extra difference when
estimating the fluid temperature rise.

Table 6.7 Variation of borefield thermal resistances with time for various borefields
without groundwater flow

Time Borefield thermal resistance (mK/W)


(year) Single 2x2 3x3 5x5
1 0.1255 0.1255 0.1254 0.1254
2 0.1255 0.1254 0.1254 0.1253
5 0.1254 0.1253 0.1252 0.1250
10 0.1254 0.1252 0.125 0.1247

6.1.2.5 Borehole Performance with Time Schedule

The effect of applying a time schedule on borehole performance was also


important, as the majority of systems did not operate on a schedule of 24 hours per
day. With a time schedule, the amount of energy transferred would be reduced and
the off-period helped relieve the temperature change along the borehole surface. At
present, attention was paid to the load transferred to borehole during each schedule.
Figure 6.6 indicated the variation of the borefield load for a single borehole with time
based on a constant applied load with a daily schedule of 12 hours per day. The load
increased at the start of the on-period to nearly 4000W at the end of on-period.
During the off-period, the residual temperature difference between the fluid and
borehole surface contributed an off-period loading to the ground although this was
small. However, the total energy transferred to the ground during one daily schedule
was still lower than the total energy applied to the borehole. This deficiency could
lead to certain errors in the simulated data if the load was assumed to be constant
when using an analytical model. The effect could be significant when the system

80
was switched on and off frequently during the low-load period. This also illustrated
the benefit of applying numerical modeling for the borefield in the dynamic
simulation of an actual system as compared with the load aggregation method using
an analytical model.

Variation for borefield load with time

4000
3500

Borefield load (W)


3000
2500
2000
1500
1000
500
0
0 10 20 30 40 50 60 70 80
Time (hour)

Figure 6.6 Variation of borefield load during a daily operation schedule of 12 hours
for single borehole without groundwater flow

6.1.3 Performance of a Single Borehole with Groundwater Advection

6.1.3.1 Leaving Fluid Temperature Rise of a Single Borehole at Different Darcy


Velocities of Groundwater Flow

Leaving fluid temperature rise of single borehole

12.5
12

11.5
Tbf,f,out - To (K)

DV=0
11
DV=1E-7
10.5
DV=2E-7
10
DV=5E-7
9.5
9
8.5
0 20000 40000 60000 80000
Simulation time (Hours)

Figure 6.7 Variation of leaving fluid temperature rise of single borehole at different
Darcy velocities of groundwater flow

Figure 6.7 showed the variation of fluid leaving temperature rise of a single
borehole at different Darcy velocities of groundwater flow through a 10-year period
under a continuous constant load. At a very low Darcy velocity, the leaving fluid
temperature increased throughout the 10-year period. With a sufficiently high
groundwater velocity, the leaving fluid temperature reached a steady state within 1
year. The temperature rise was reduced with an increase in the groundwater velocity,

81
indicating the importance of an accurate prediction of the groundwater velocity in the
optimisation of a borefield design.

6.1.3.2 Temperature and Loading Variation along a Single Borehole at Different


Darcy Velocities of Groundwater Flow

Figures 6.8&9 showed the temperature rises and Figures 6.10&11 the
loadings along a single borehole at different Darcy velocity of groundwater. The
curves were basically similar to those of Figures 6.1&2. The temperature profiles
and consequently the borehole thermal resistance did not change much with
groundwater velocity. This was important in thermal response test analysis which
would be discussed in Section 6.2, as borehole thermal resistance could be
considered as independent of groundwater velocity.

Borehole temperature rise variation after 5 years Borehole temperature rise variation after 10 years

Borehole temperature rise (K) Borehole temperature rise (K)


5 6 7 8 9 10 5 6 7 8 9 10 11
0 0

0.2 0.2
DV=0 DV=0
(z-d)/H
(z-d)/H

0.4 DV=1E-7 0.4 DV=1E-7


DV=2E-7 DV=2E-7
0.6 0.6
DV=5E-7 DV=5E-7
0.8 0.8

1 1

Figure 6.8 Temperature rise along single Figure 6.9 Temperature rise along single
borehole after 5 years at various groundwater borehole after 10 years at various groundwater
velocities velocities

Borehole loading variation after 5 years Borehole loading variation after 10 years

Borehole loading (W/m) Borehole loading (W/m)


30 35 40 45 50 55 60 30 35 40 45 50 55 60
0 0

0.2 0.2
DV=0 DV=0
(z-d)/H

(z-d)/H

0.4 DV=1E-7 0.4 DV=1E-7


DV=2E-7 DV=2E-7
0.6 0.6
DV=5E-7 DV=5E-7
0.8 0.8

1 1

Figure 6.10 Loading along single borehole after Figure 6.11 Loading along single borehole
5 years at various groundwater velocities after10 years at various groundwater velocities

6.1.3.3 Performance of a Single Borehole at Different Loading Profiles

82
Figures 6.12~15 indicated the performance of a single borehole under
different loading profiles but with the same peak cooling load. It could be found that
the introduction of a daily operating schedule of 12 hours reduced the fluid leaving
temperature rise by less than 35% at no groundwater flow although the total heat
rejection to the ground was decreased by 50%. The application of a periodic load
only changed the borehole performance a little, and the inclusion of heating period
had insignificant effects. In contrast to a 81.25% reduction in the total heat injection
to the ground, the fluid leaving temperature rise for Profile 4 was only reduced by
42%. Indeed, the effect of loading profiles decreased with increase in groundwater
velocity, and the borehole performance became independent of the loading profiles at
a very high groundwater velocity, say 10-5m/s or higher. This indicated that the peak
cooling load was a very important parameter in the design of the borefield. The
reduction of the peak cooling load by other means, such as adopting a hybrid system
to share the cooling load at the peak-load season, could significantly reduce the
required borehole lengths especially when the groundwater velocity was high.

Leaving fluid temperature rise with no groundwater flow Leaving fluid temperature rise with groundwater velocity
at different load profiles 1E-7m/s at different load profiles

14 12
12
Tbf,f,out - To (K)

10
Tbf,f,out - To (K)

10 Profile 1 Profile 1
8
8 Profile 2 Profile 2
6
6 Profile 3 Profile 3
4 4
Profile 4 Profile 4
2 2
0 0
0 20000 40000 60000 80000 100000 0 20000 40000 60000 80000 100000
Time (Hour) Time (Hour)

Figure 6.12 Performance of single borehole at Figure 6.13 Performance of single borehole at
no groundwater flow for different load profiles groundwater velocity 10-7m/s for different load
profiles

Leaving fluid temperature rise with groundwater velocity Leaving fluid temperature rise after 1 year with high
2E-7m/s at different load profiles groundwater velocity at different load profiles

12 12
10
Tbf,f,out - To (K)

10
Tbf,f,out - To (K)

Profile 1 Profile 1
8 8
Profile 2 Profile 2
6 6
Profile 3 Profile 3
4 4
Profile 4 Profile 4
2 2
0 0
0 20000 40000 60000 80000 100000 0 20 40 60 80 100

Time (Hour) Groundwater velocity (1E-7m/s)

Figure 6.14 Performance of single borehole at Figure 6.15 Performance of single borehole after
groundwater velocity 2x10-7m/s for different load 1 year at high groundwater velocity for different
profiles load profiles

83
6.1.4 Effect of Groundwater Velocity on Borefield Design Length at Different
Load Profiles

As already shown in Figure 6.7, the presence of groundwater reduced the


leaving fluid temperature rise. Hence, the borehole length could be shortened while
the leaving fluid temperature could still be kept at the maximum allowable value.
Comparison was made between the borehole length of a single borehole and a 2x2
borefield for an assumed groundwater direction parallel to the x-axis under different
load profiles for a period of ten years. The effects were studied with and without
groundwater flow based on an approximately equal leaving fluid temperature rise
which was set at 12.269oC corresponding to the case of a single borehole with load
profile 1 without groundwater flow. The peak borefield loads were adjusted so that
for no groundwater flow, the borehole length remained the same for different
borefields and load profiles, as summarised in Table 6.8. The variations of the
borehole lengths and capacities per unit length with groundwater flow were
expressed as ratios to those values at no groundwater flow, as shown in Figures
6.16~19.

Table 6.8 Variation of peak borefield load with fixed leaving fluid temperature rise
at different load profiles for single and 2x2 borefield without groundwater flow

Peak borefield load at different loading profiles (W)


Borefield Profile 1 Profile 2 Profile 3 Profile 4
Single 4000 5952 6598 6786
2x2 9793 16467 21206 22856

Variation of borehole length ratio with groundwater Variation of borehole capacity ratio with groundwater
velocity at different load profiles for single borehole velocity at different load profiles for single borehole

1.2 2.5
Borehole capacity ratio
Borehole length ratio

1
Profile 1 2 Profile 1
0.8
Profile 2 Profile 2
0.6 1.5
Profile 3 Profile 3
0.4
Profile 4 1 Profile 4
0.2
0 0.5
0 20 40 60 80 100 0 20 40 60 80 100

Groundwater velocity (1E-7m/s) Groundwater velocity (1E-7m/s)

Figure 6.16 Variation of borehole length ratio Figure 6.17 Variation of borehole capacity ratio
with groundwater velocity at different load with groundwater velocity at different load
profiles with fixed leaving fluid temperature rise profiles with fixed leaving fluid temperature rise
for single borehole for single borehole

84
Variation of borehole length ratio with groundwater Variation of borehole capacity ratio with groundwater
velocity at different load profiles for 2x2 borefield velocity at different load profiles for 2x2 borefield

1.2 3.5

Borehole capacity ratio


Borehole length ratio
1 3
Profile 1 Profile 1
0.8 2.5
Profile 2 Profile 2
0.6 2
Profile 3 Profile 3
0.4 1.5
Profile 4 Profile 4
0.2 1
0 0.5
0 20 40 60 80 100 0 20 40 60 80 100
Groundwater velocity (1E-7m/s) Groundwater velocity (1E-7m/s)

Figure 6.18 Variation of borehole length ratio Figure 6.19 Variation of borehole capacity ratio
with groundwater velocity at different load with groundwater velocity at different load profiles
profiles with fixed leaving fluid temperature rise with fixed leaving fluid temperature rise for 2x2
for2x2 borefield borefield

As expected, the effect of groundwater on borehole length and capacity was


more pronounced at low groundwater velocity, and reduced with an increase in
groundwater velocity. The influence was greatest for a continuous constant load, and
became lower with the introduction of a daily operating schedule and periodic load.
Moreover, the effect of groundwater increased with the scale of the borefield. The
borehole length ratio was nearly 20% lower for a 2x2 borefield as compared with
that for a single borehole even for Load Profile 4. This illustrated the importance of
consideration of the borefield size when discussing groundwater effect on borefield
performance rather than just investigating a single borehole as carried out by some
previous researchers, and the significance of including groundwater effect in the
design and modelling of a large borefield.

6.1.5 Effect of Borehole Length on Performance of Borefield at Different


Groundwater Velocities and Load Profiles

The borehole length determined the initial cost of geothermal heat pump
systems. Hence, the understanding of the effect of borehole length on borefield
performance helped select the best length to be used. Analysis was made for single
and 2x2 borefields by comparing the borefield performance having longer ( H = 120)
and shorter ( H = 100) effective borehole lengths with that using the default length
( H = 110) in terms of the leaving fluid temperature rise ratios under different
groundwater velocities assumed to be parallel to the x-axis and load profiles, as
shown in Figures 6.20~23.

85
Variation of leaving fluid temperature rise ratio of single Variation of leaving fluid temperature rise ratio of single
borehole with short borehole length borehole with long borehole length

1.14 0.915

Temperature rise ratio

Temperature rise ratio


1.135 0.91
1.13 Profile 1 Profile 1
0.905
1.125 Profile 2 Profile 2
0.9
1.12 Profile 3 Profile 3
1.115 0.895
Profile 4 Profile 4
1.11 0.89
1.105 0.885
0 20 40 60 80 100 0 20 40 60 80 100

Groundwater velocity (1E-7m/s) Groundwater velocity (1E-7m/s)

Figure 6.20 Variation of leaving fluid temperature Figure 6.21 Variation of leaving fluid temperature
rise ratio of single borehole with short borehole rise ratio of single borehole with long borehole
length at different groundwater velocities and load length at different groundwater velocities and load
profiles profiles

Variation of leaving fluid temperature rise ratio of 2x2 Variation of leaving fluid temperature rise ratio of 2x2
borefield with short borehole length borefield with long borehole length

1.135 0.92
Temperature rise ratio

Temperature rise ratio


1.13 0.915
1.125
Profile 1 0.91 Profile 1
1.12
1.115 Profile 2 0.905 Profile 2
1.11 Profile 3 0.9 Profile 3
1.105
Profile 4 0.895 Profile 4
1.1
1.095 0.89
1.09 0.885
0 20 40 60 80 100 0 20 40 60 80 100

Groundwater velocity (1E-7m/s) Groundwater velocity (1E-7m/s)

Figure 6.22 Variation of leaving fluid temperature Figure 6.23 Variation of leaving fluid temperature
rise ratio of 2x2 borefield with short borehole rise ratio of 2x2 borefield with long borehole
length at different groundwater velocities and load length at different groundwater velocities and load
profiles profiles

It could be found that the temperature rise ratio departed more from unity
with an increasing groundwater velocity, indicating that the borefield performance
was more sensitive to borehole length at a higher groundwater flow. The curves
basically differed very little under different loading profiles, unlike the borefield
performance which changed substantially with the application of daily operating
schedule. Hence, the study of borefield size on the temperature rise ratio needed
only be based on one loading profile such as a periodic load with no heating with a
daily operating schedule.

Figures 6.24&25 depicted the variation of the leaving fluid temperature rise
ratio for various borefields. The curves shifted towards the unity line where
temperature ratio equals 1 when the borefield size increased, indicating that the
percentage change in borefield performance became smaller under the same change
in borehole length. In other words, the benefits in the form of a much higher energy
efficiency for the geothermal heat pumps due to a lower leaving fluid temperature

86
rise from the borefield could only be achieved at the expense of a much higher cost
because of the larger increase in the borehole length for a large borefield. This
illustrated again the importance of considering every possible means for improving
the borefield performance when designing a large borefield, including groundwater
direction effect to be discussed later.

Variation of temperature rise ratio with short borehole Variation of temperature rise ratio with long borehole
length for different borefields length for different borefields

1.14 0.93
Temperature rise ratio

Temperature rise ratio


0.925
1.13
0.92
Single Single
1.12 0.915
2x2 0.91 2x2
1.11
3x3 0.905 3x3
1.1 0.9
5x5 5x5
1.09 0.895
0.89
1.08 0.885
0 20 40 60 80 100 0 20 40 60 80 100
Groundwater velocity (1E-7m/s) Groundwater velocity (1E-7m/s)

Figure 6.24 Variation of leaving fluid temperature Figure 6.25 Variation of leaving fluid temperature
rise ratio with short borehole length at different rise ratio with long borehole length at different
groundwater velocities and borefields under load groundwater velocities and borefields under load
profile 3 profile 3

In the above analysis, the peak applied borefield load per borehole was fixed
at the default value. This meant that the leaving fluid temperature rise would be
lower at higher groundwater velocities. It might not be clear if the temperature rise
ratios shown in Figures 6.20~25 could not reflect the actual change in the leaving
fluid temperature rise which determined the heat pump performance. Hence, another
analysis was made for a 2x2 borefield where the borefield load was adjusted so that
the same leaving fluid temperature rise of 12.269oC would apply as used in Section
6.1.4 for the default borehole length, as shown in Table 6.9.

Table 6.9 Variation of peak borefield load at different load profiles and groundwater
flow for 2x2 borefield based on fixed leaving fluid temperature rise at default
borehole length

Load Peak borefield load (W) at different groundwater velocities (m/s)


profile 1x10-7 2x10-7 5x10-7 1x10-6 2x10-6 5x10-6 1x10-5
1 11325 13709 18276 22348 26905 34726 42960
2 18494 21360 26112 31367 35244 40721 45728
3 22826 24826 27544 31652 35280 40725 45730
4 24249 25881 27929 31724 35289 40726 45731

87
Figures 6.26&67 showed the variation of the leaving fluid temperature rise
ratio at different groundwater velocities and load profiles. The trends were very
similar except that the introduction of daily operating schedule produced a greater
difference in the temperature rise ratio, especially at higher groundwater velocities.
Hence, it could be expected that the same patterns could be obtained for Figures
6.24&25 even with a fixed leaving fluid temperature rise at the default borehole
length.

Variation of leaving fluid temperature rise ratio of 2x2 Variation of leaving fluid temperature rise ratio of 2x2
borefield with short borehole length borefield with long borehole length

1.14 0.92

Temperature rise ratio


Temperature rise ratio

0.915
1.13
Profile 1 0.91 Profile 1
1.12 Profile 2 0.905 Profile 2

1.11 Profile 3 0.9 Profile 3


Profile 4 0.895 Profile 4
1.1
0.89
1.09 0.885
0 20 40 60 80 100 0 20 40 60 80 100

Groundwater velocity (1E-7m/s) Groundwater velocity (1E-7m/s)

Figure 6.26 Variation of leaving fluid temperature Figure 6.27 Variation of leaving fluid temperature
rise ratio of 2x2 borefield with short borehole rise ratio of 2x2 borefield with long borehole
length at different groundwater velocities and load length at different groundwater velocities and load
profiles based on fixed leaving fluid temperature profiles based on fixed leaving fluid temperature
rise at default borehole length rise at default borehole length

6.1.6 Performance of Borefield at Different Groundwater Directions

So far in the previous analysis, groundwater was assumed to flow parallel to


the x-axis. Indeed, with more than one borehole, the borefield performance might be
different at different groundwater directions, and a study of the effect of groundwater
direction allowed optimised design to be achieved, especially for large borefield as
mentioned in previous sections. As mentioned in Section 3.1.4, the effect of
groundwater direction on borefield performance was represented by the percentage
variation (PV) which indicated the maximum percentage difference in the borefield
fluid temperature rise at different groundwater directions. Detailed information
about borefield performance at different groundwater directions could be found in
Appendix J.

6.1.6.1 Variation of PV for Square Borefields

88
Figures 6.28&29 showed the variation of PV for 2x2 and 3x3 borefield at
different groundwater velocities. It could be found that PV generally increased with
increase in groundwater velocity. PV was highest for the case of a continuous
constant load. With the introduction of a daily operating schedule, the maximum PV
was reduced by more than 40%. Replacing the constant load with a periodic load
approximated by a sine curve had insignificant effects even when a heating period
was applied to the load cycle (peak cooling load/peak heating load = 4). The patterns
were very similar to the performance for a single borehole as shown previously in
Figures 6.12~15. Basically, the effect of groundwater direction was small for square
borefields, and was negligible for groundwater velocities below 10-6m/s.

Change of PV with groundwater velocity after 10 years at Change of PV with groundwater velocity after 10 years at
different load profiles different load profiles

4 6
5
3 Profile 1 Profile 1
4
PV (%)

PV (%)

Profile 2 Profile 2
2 3
Profile 3 Profile 3
Profile 4
2 Profile 4
1
1
0 0
0 20 40 60 80 100 0 20 40 60 80 100
Darcy velocity of groundwater (10-7m/s) Darcy velocity of groundwater (10-7m/s)

Figure 6.28 Variation of PV with groundwater Figure 6.29 Variation of PV with groundwater
velocity for 2x2 borefield velocity for 3x3 borefield

6.1.6.2 Variation of PV for Non- square Borefields

Change of PV with groundwater velocity after 10 years at Change of PV with groundwater velocity after 10 years at
different load profiles different load profiles

8 14
12
6 Profile 1 10 Profile 1
PV (%)

PV (%)

Profile 2 8 Profile 2
4
Profile 3 6 Profile 3

2 Profile 4 4 Profile 4
2
0 0
0 20 40 60 80 100 0 20 40 60 80 100
Darcy velocity of groundwater (10-7m/s) Darcy velocity of groundwater (10-7m/s)

Figure 6.30 Variation of PV with groundwater Figure 6.31 Variation of PV with groundwater
velocity for 1x2 borefield velocity for 1x3 borefield

Figures 6.30~31 depicted the variation of PV at different groundwater


velocities for 1x2 and 1x3 borefields. The effects of loading profiles were similar to
those for square borefields except that the application of periodic load contributed a
mild effect on PV. Moreover, PV increased with groundwater velocity until it

89
reached a maximum, and then decreased with further increase in groundwater
velocity. The maximum PV was nearly twice that at a Darcy velocity of 10-5m/s
under a constant load and with a daily operating schedule.

6.1.6.3 Variation of PV for Different Borefield Scales

Figure 6.32 showed the PV for various borefields of different shapes at a


groundwater velocity of 10-5m/s under a constant load and with a daily operating
schedule of 12 hours. A constant load was used since the application of a periodic
load only changed the results mildly at low groundwater velocities. It could be
observed that PV generally increased with the scale of borefields. From Figures
6.28&29, the maximum PV with a daily operating schedule was only slightly higher
(less than 10%) than that at a Darcy velocity of 10-5m/s. Hence, it could be
anticipated that the maximum PV for a 10x10 borefield would be around 5%.
However, from Figures 6.30&31, the maximum PV with a daily operating schedule
could be 40% higher than that at a Darcy velocity of 10-5m/s. Hence, the maximum
PV could reach 12% for a 5x10 borefield and 10% for a 6x9 borefield. The range of
groundwater flow direction between the maximum and minimum fluid leaving
temperatures did not always equal the full groundwater direction range (45o for a
square borefield and 90o for a non-square borefield as shown in Appendix J). Indeed
for a 2x2 borefield, the maximum and minimum fluid leaving temperatures occurred
within a range of 15o of groundwater flow direction, while for 1x2 borefield with a
very high groundwater flow velocity, the range was 60o. This meant that in the
installation of the borefield, precise site alignment for the boreholes should be made
in order that the best borefield orientation was used.

PV after 10 years at Darcy velocity 1E-5m/s and a


constant load with a daily operating schedule of 12 hours

10
8
Square
PV (%)

6
1:2
4
2:3
2
0
0 2 4 6 8 10 12
Number of boreholes in x-direction

Figure 6.32 PV of various borefields with groundwater velocity of 10-5m/s at


constant load with a daily schedule

90
The reduction in the leaving fluid temperature rise affected the design length
of the boreholes, and the effect could be investigated by reviewing Figures 6.24&25.
With the temperature rise ratio closer to unity, a larger reduction in borehole length
could be achieved for the same improvement in borefield performance. In Figure
6.24 where the effective borehole length was reduced from 110 to 100m (about
9.091% reduction in length), a temperature rise ratio of lower than 1.1 would mean
that the percentage reduction in borehole length could be higher than the percentage
decrease in borefield performance. This applied to large borefields at low
groundwater flow velocities. The same finding could be made if the temperature rise
ratio is higher than 0.9167 as shown in Figure 6.25. Hence, a 10% decrease in the
leaving fluid temperature rise could possibly result in up to 15% decrease in the
borehole length for very large borefields, leading to a significant saving in the
installation cost. This was based on the assumption that the same allowable
maximum leaving fluid temperature and the recognition that a 10% decrease in the
temperature rise required a 11.1% increase in order to bring the temperature rise back
to the maximum limit. This highlighted the importance of understanding the
groundwater conditions in regard to both the magnitude and the direction in order to
achieve an optimal design. A correct setting of the orientation of the borefields could
cause a substantial reduction in the installation cost and shortening of the payback
period.

6.2 Thermal Response Test Analysis

6.2.1 Finalisation of Parameters to be Determined

There were many parameters that affected the performance of a single


borehole, namely, ground thermal conductivity, ground volumetric heat capacity,
groundwater velocity, groundwater volumetric heat capacity, borehole backfill
thermal conductivity, U-tube spacing, U-tube position inside the borehole, etc. It
would be inefficient to attempt estimating all of them in TRT. Rather, the
parameters inside the borehole could be consolidated into a single parameter,
borehole thermal resistance. As already mentioned in Section 6.1.3.2, the thermal
resistance of a single borehole was considered independent of the groundwater

91
velocity. By performing simulation using Eq. 3.3 based on the correlations for the
thermal interference coefficients as proposed by Hellstrom (1991), it was found that
the borehole thermal resistance was very weakly affected by the ground thermal
conductivity. Hence, there were basically five independent parameters left.

To select if particular parameters were to be determined in TRT, the


sensitivity of each parameter besides groundwater velocity on the borehole
performance in TRT was investigated, as shown in Figures 6.33~36. The variation
of the borehole performance in respect of the fluid temperature rise due to change of
a particular parameter while the others were kept constant was compared with that
using the nominal value of that parameter through a borehole temperature rise ratio.
The nominal values used were: 3.5W/mK for the ground thermal conductivity,
2.16x106J/m3K for the ground volumetric heat capacity, 0.1253mK/W for the
borehole thermal resistance and 4.19x106J/m3K for the groundwater volumetric heat
capacity.

Sensitivity of ground volumetric heat capacity on Sensitivity of ground thermal conductivity on borehole
borehole performance at different groundwater velocity performance at different groundwater velocity
Borehole fluid temperature rise
Borehole fluid temperature rise

1.025 1.15
1.02 DV=0 DV=0
1.015 1.1
DV=1E-7 DV=1E-7
1.01
DV=2E-7 1.05 DV=2E-7
1.005
ratio
ratio

1 DV=5E-7 DV=5E-7
1
0.995
DV=1E-6 DV=1E-6
0.99 0.95
0.985 DV=2E-6 DV=2E-6
0.98 DV=5E-6 0.9 DV=5E-6
0.8 0.9 1 1.1 1.2 DV=1E-5 0.8 0.9 1 1.1 1.2 DV=1E-5
Ground volumetric heat capacity ratio Ground thermal conductivity ratio

Figure 6.33 Sensitivity of ground volumetric heat Figure 6.34 Sensitivity of ground thermal
capacity on borehole performance at different conductivity on borehole performance at different
groundwater velocities in TRT groundwater velocities in TRT

Sensitivity of borehole thermal resistance on borehole Sensitivity of groundwater volumetric heat capacity on
performance at different groundwater velocity borehole performance at different groundwater velocity
Borehole fluid temperature rise
Borehole fluid temperature rise

1.2 1.07
1.15 DV=0 1.05 DV=0
1.1 DV=1E-7 DV=1E-7
1.03
1.05 DV=2E-7
DV=2E-7
ratio
ratio

1 1.01
DV=5E-7 DV=5E-7
0.95 0.99
DV=1E-6 DV=1E-6
0.9
0.97 DV=2E-6
0.85 DV=2E-6
0.8 DV=5E-6 0.95 DV=5E-6
0.8 0.9 1 1.1 1.2 0.8 0.9 1 1.1 1.2 DV=1E-5
DV=1E-5
Borehole thermal resistance ratio Groundwater volumetric heat capacity ratio

Figure 6.35 Sensitivity of borehole thermal Figure 6.36 Sensitivity of groundwater volumetric
resistance on borehole performance at different heat capacity on borehole performance at
groundwater velocities in TRT different groundwater velocities in TRT

92
For a parameter variation of 20% in the nominal parameter value, it could be
found that the ground volumetric heat capacity only changed the borehole
performance by less than 2%. The sensitivity was reduced as the groundwater
velocity increased, and became zero at high groundwater velocities. For the
groundwater volumetric heat capacity, the sensitivity was also low (less than 2%) for
groundwater velocity below 2x10-6m/s. A low sensitivity for a particular parameter
would result in a lower precision for the respective parameter value obtained by
applying appropriate parameter estimation techniques. Accordingly, the ground and
groundwater volumetric heat capacities would be specified instead of being
determined in the TRT analysis.

The role of groundwater was complicated, as its effect could be small in TRT
at low groundwater velocities as depicted in Figure 6.37 with Darcy velocities less
than 10-6m/s, but still had substantial influence on the long-term borehole
performance as already discussed in Section 6.1.3.1. Hence, a separate study needed
to be conducted first to investigate the behaviour of groundwater velocity in TRT
analysis before the final parameter estimation scheme could be set.

Fluid leaving temperature rise at different groundwater


velocity

9
8 DV=0
7
DV=1E-7
Tout - To (K)

6
5 DV=2E-7
4 DV=5E-7
3
DV=1E-6
2
1 DV=2E-6
0 DV=5E-6
0 20 40 60 80 100 120 DV=1E-5
T ime (Hours)

Figure 6.37 Performance of single borehole in TRT at different groundwater


velocities

6.2.2 Analysis Based on the Generated TRT Data with Specified Groundwater
Velocity

To understand the behaviour of groundwater velocity in TRT analysis, the


groundwater velocity was initially specified and only the ground thermal
conductivity and the borehole thermal resistance were determined from the
parameter estimation scheme. Sample thermal response test data for five days

93
generated from the models were used. The advantage was that a very small MSD
(theoretically zero) could result if the estimated and specified parameters matched
with the ones used to generate the data. Three models were tried, namely the
LSMGA, FLSMGA and present numerical model. A borehole thermal resistance of
0.1253mK/W was used to generate the data. For LSMGA, a borehole loading of
30W/m was adopted. Initially, several optimisation methods had been applied,
namely the gradient search method, the genetic algorithm and the Nelder-Mead
simplex method. It was found that the simplex method gave the fastest convergence
and highest precision, and was thus adopted in all the subsequent analysis.

6.2.2.1 Analysis Based on a Normal Test Period

Table 6.10 Parameter estimation results with specified groundwater flow using
LSMGA on generated TRT data for 5 days without groundwater

Specified groundwater Average analysis results based on data generated


velocity used in parameter without groundwater
estimation (10-8m/s) kg (W/mK) Rb (mK/W) MSD (K2)
0 3.500 0.1253 0.000000
1 3.499 0.1253 0.000000
2 3.501 0.1253 0.000000
5 3.499 0.1253 0.000000
10 3.498 0.1253 0.000000
20 3.490 0.1251 0.000001
50 3.438 0.1243 0.000030
100 3.258 0.1212 0.000516

Table 6.10 showed the results using LSMGA based on generated TRT data
without groundwater flow for 5 days. At a specified groundwater velocity of 10-7m/s
or less, the estimated average MSD, k g and Rb were almost the same and very close

to the theoretical values of 3.5 for k g and 0.1253 for Rb . This agreed with the

borehole performance previously shown in Figure 6.35 that at low groundwater


velocities, the curves were very close to each other. When the specified groundwater

94
velocity went beyond 2x10-7m/s, a more significant deviation of the analysis results
from the theoretical values was observed, and the MSD became much higher.

Table 6.11 Parameter estimation results with specified groundwater flow using
LSMGA on generated TRT data for 5 days with groundwater velocity 2x10-7m/s

Specified groundwater Average analysis results based on data generated


velocity used in parameter with groundwater velocity 2x10-7m/s
estimation (10-8m/s) kg (W/mK) Rb (mK/W) MSD (K2)
0 3.510 0.1255 0.000001
1 3.510 0.1255 0.000001
2 3.510 0.1255 0.000001
5 3.509 0.1255 0.000001
10 3.508 0.1254 0.000001
20 3.499 0.1253 0.000000
50 3.448 0.1244 0.000021
100 3.268 0.1214 0.000472

Table 6.11 indicated the analysis results based on the generated data
corresponding to a groundwater velocity of 2x10-7m/s. Again, the results were
almost the same as those for a specified groundwater velocity of 10-7m/s or less,
which were slightly differentiable from those at a specified groundwater velocity of
2x10-7m/s. It could be expected that if the actual groundwater velocity in the thermal
response test was very low (say 10-7m/s or less), a possible range of groundwater
velocity might be obtained from parameter estimation ranging from 0 to 10-7m/s.
However, at high groundwater velocities, the results seemed promising as shown in
Table 6.12. In other words, precise determination of low groundwater velocity from
the thermal response test was not possible, or there existed a confident minimum
for the groundwater velocity above which it was able to estimate its value from TRT
analysis with confidence. The significance of this confident minimum was
demonstrated in Section 6.2.3 where the groundwater velocity was also included in
the parameter estimation scheme. Analysis results for other specified groundwater
velocities were presented in Appendix K. Corresponding studies based on the

95
generated data using FLSMGA and the present numerical model followed the same
trends as those using LSMGA.

Table 6.12 Parameter estimation results with specified groundwater flow using
LSMGA on generated TRT data for 5 days with groundwater velocity 10-6m/s

Specified groundwater Average analysis results based on data generated


velocity used in parameter with groundwater velocity 10-6m/s
estimation (10-8m/s) kg (W/mK) Rb (mK/W) MSD (K2)
0 3.752 0.1292 0.000391
1 3.751 0.1292 0.000391
2 3.751 0.1292 0.000390
5 3.750 0.1292 0.000389
10 3.748 0.1291 0.000383
20 3.741 0.1290 0.000361
50 3.688 0.1282 0.000225
100 3.500 0.1253 0.000000

6.2.2.2 Analysis Based on a Long Test Period

As mentioned in previous section, the analysis results were very close at low
groundwater velocities of less than 10-7m/s. To differentiate the thermal response
test performance at this low velocity range, a longer test period of 30 days was tried
with data generated using LSMGA as shown in Table 6.13&14 under groundwater
velocities of 2x10-8m/s and 5x10-8m/s. It could be observed that the MSDs were
almost the same for groundwater velocities 5x10-8m/s or less. This meant that it
should be able to predict the groundwater velocity with confidence when it was
greater than 5x10-8m/s with the test period extended to 30 days. The confident
minimum for the estimated groundwater velocity decreased with increase in the test
period. The significance of groundwater effect at this velocity range depended on
the scale of the borefield, which could still be substantial for a large borefield. In
other words, in designing a large borefield, it would be beneficial to perform the
TRT with a longer testing period. The corresponding results at other groundwater

96
velocities were shown in Appendix L. Again, the analysis using other models led to
similar findings.

Table 6.13 Parameter estimation results with specified groundwater flow using
LSMGA on generated TRT data for 30 days with groundwater velocity 2x10-8m/s

Specified groundwater Average analysis results based on data generated


velocity used in parameter with groundwater velocity 2x10-8m/s
estimation (10-8m/s) kg (W/mK) Rb (mK/W) MSD (K2)
0 3.500 0.1253 0.000000
1 3.500 0.1253 0.000000
2 3.500 0.1253 0.000000
5 3.497 0.1252 0.000000
10 3.487 0.1249 0.000002
20 3.446 0.1238 0.000032
50 3.177 0.1162 0.001422
100 2.560 0.0999 0.027382

Table 6.14 Parameter estimation results with specified groundwater flow using
LSMGA on generated TRT data for 30 days with groundwater velocity 5x10-8m/s

Specified groundwater Average analysis results based on data generated


velocity used in parameter with groundwater velocity 5x10-8m/s
estimation (10-8m/s) kg (W/mK) Rb (mK/W) MSD (K2)
0 3.503 0.1254 0.000000
1 3.504 0.1254 0.000000
2 3.504 0.1254 0.000000
5 3.500 0.1253 0.000000
10 3.490 0.1250 0.000001
20 3.449 0.1239 0.000029
50 3.179 0.1163 0.001395
100 2.556 0.0995 0.027203

97
6.2.3 Analysis Based on the Generated TRT Test Data Including Groundwater
Velocity in the Parameter Estimation Scheme

In the previous section, analysis based on specified groundwater velocities


revealed that it was unable to differentiate the TRT performance when the
groundwater velocity was below 10-7m/s for a normal test period. To illustrate this,
groundwater velocity was also included in the parameter estimation scheme, and
again analysis made using the generated data.

6.2.3.1 Analysis Based on a Normal Test Period

Tables 6.15~17 indicated the analysis results based on generated data using
LSMGA with groundwater velocities of 5x10-8, 10-7 and 2x10-7m/s respectively for a
test period of five days. From Tables 6.15&16, the estimated groundwater velocity
in the three trials scattered and deviated much from the theoretical values while the
estimated ground thermal conductivity and borehole thermal resistance converged to
the theoretical values. It should be noted that the sign was unimportant for single
borehole performance. This agreed with the findings in Section 6.2.2.1. In Table
6.17, the situation was quite different. The estimated groundwater velocity was
much closer to the theoretical value, and the confident minimum for the estimated
groundwater velocity under a normal test period of five days was 2x10-7m/s. Above
the confident minimum, the convergence of the estimated groundwater velocity was
even much better as shown in Appendix M. Analysis with other models reached
very similar results.

Table 6.15 Parameter (including groundwater velocity) estimation results using


LSMGA on generated TRT data for 5 days with groundwater velocity 5x10-8m/s

Analysis results based on data generated with


groundwater velocity 5x10-8m/s
Trial kg (W/mK) Rb (mK/W) Vgw (m/s) MSD (K2)
1 3.498 0.1253 1.19x10-7 0.000000
2 3.497 0.1253 1.45x10-7 0.000000
3 3.498 0.1253 -1.36x10-7 0.000000

98
Table 6.16 Parameter (including groundwater velocity) estimation results using
LSMGA on generated data for 5 days with groundwater velocity 10-7m/s

Analysis results based on data generated with


groundwater velocity 10-7m/s
Trial kg (W/mK) Rb (mK/W) Vgw (m/s) MSD (K2)
1 3.503 0.1254 4.77x10-8 0.000000
2 3.495 0.1252 1.53x10-7 0.000000
3 3.493 0.1252 1.79x10-7 0.000000

Table 6.17 Parameter (including groundwater velocity) estimation results using


LSMGA on generated data for 5 days with groundwater velocity 2x10-7m/s

Analysis results based on data generated with


groundwater velocity 2x10-7m/s
Trial kg (W/mK) Rb (mK/W) Vgw (m/s) MSD (K2)
1 3.494 0.1252 -2.48x10-7 0.000000
2 3.499 0.1253 -2.07x10-7 0.000000
3 3.498 0.1253 2.16x10-7 0.000000

6.2.3.2 Analysis Based on a Long Test Period

Tables 6.18&19 showed the analysis results based on the generated data
using LSMGA with groundwater velocities of 5x10-8 and 10-7m/s for a test period of
30 days. It was evident that the estimated groundwater velocity converged much
more readily to the theoretical value at a groundwater velocity of 10-7m/s than that at
a groundwater velocity of 5x10-8m/s. This totally agreed with the analysis in Section
6.2.2.2 that for a longer test period, confident results could still be obtained even at
lower groundwater flow velocities. The confident minimum for the estimated
groundwater velocity was reduced to 10-7m/s at a long test period of 30 days.
Corresponding analysis results with other groundwater velocities were shown in
Appendix N, and the results using other models were basically the same at that using
LSMGA.

99
Table 6.18 Parameter (including groundwater velocity) estimation results using
LSMGA on generated data for 30 days with groundwater velocity 5x10-8m/s

Analysis results based on data generated with


groundwater velocity 5x10-8m/s
Trial kg (W/mK) Rb (mK/W) Vgw (m/s) MSD (K2)
1 3.504 0.1254 -2.65x10-8 0.000000
2 3.499 0.1253 5.86x10-8 0.000000
3 3.495 0.1252 -7.87x10-8 0.000000

Table 6.19 Parameter (including groundwater velocity) estimation results using


LSMGA on generated data for 30 days with groundwater velocity 10-7m/s

Analysis results based on data generated with


groundwater veloctiy 10-7m/s
Trial kg (W/mK) Rb (mK/W) Vgw (m/s) MSD (K2)
1 3.499 0.1253 1.00x10-7 0.000000
2 3.498 0.1252 1.03x10-7 0.000000
3 3.498 0.1252 1.03x10-7 0.000000

6.2.3.3 Effects of Other Parameters on the Confident Minimum of Estimated


Groundwater Velocity

As the ground thermal conductivity and borehole thermal resistance affected


the borefield performance, one might suspect if the confident minimum of estimated
groundwater velocity would also be strongly related. To check this, analysis was
made with different k g (2 and 4) and Rb (0.1 and 0.2) on generated data using

LSMGA under groundwater velocities of 10-7, 2x10-7 and 5x10-7m/s as shown in


Table 6.20. Compared with Table 6.17 and Tables M.4&5, the change of Rb from
0.1 to 2 did not seem to alter the confident minimum of the estimated groundwater
velocity. This could be explained by reviewing Eq. (2.4) showing that for the same
loading, a change in Rb would only shift the fluid temperature profiles up or down

with the same extent even at different groundwater velocities. For k g , a lower value

tended to yield less scattering in the estimated groundwater velocities. This was due

100
to the fact that a lower k g resulted in a higher ground temperature rise and

consequently a higher fluid temperature rise and the convective effect of the
groundwater became more substantial. Hence, the separation of the borehole
temperature rise profiles at different groundwater velocities would increase, meaning
that it would become easier to differentiate the groundwater effect, thus leading to a
reduction of the confident minimum of the estimated groundwater velocity.

Table 6.20 Estimated groundwater velocities using LSMGA based on generated data
with various ground thermal conductivities and borehole thermal resistances

Estimated V based on generated data with


kg Rb Trial different groundwater velocities (m/s)
(W/mK) (mK/W) 10-7 m/s 2x10-7 m/s 5x10-7 m/s
1 9.63x10-8 2.00x10-7 4.97x10-7
2 0.1253 2 8.72x10-8 2.01x10-7 5.07x10-7
3 7.73x10-8 2.11x10-7 4.97x10-7
1 1.73x10-7 1.01x10-7 5.14x10-7
4 0.1253 2 1.81x10-7 2.09x10-7 5.13x10-7
3 1.15x10-7 2.66x10-7 5.03x10-7
1 9.10x10-8 1.58x10-7 5.08x10-7
3.5 0.1 2 9.66x10-8 2.21x10-7 4.88x10-7
3 1.17x10-7 1.97x10-7 4.94x10-7
1 1.33x10-7 1.27x10-7 4.96x10-7
3.5 0.2 2 1.33x10-7 1.99x10-7 5.02x10-7
3 1.33x10-7 2.38x10-7 4.96x10-7

6.2.3.4 Effects of Far Field Temperature and Borehole Load on the Estimated
Parameters

When conducting the TRT, the borefield load (Qb ) was provided by an
electric water heater and calculated from the logged temperature difference between
the borehole inlet and outlet. There might be a small fluctuation in Qb due to noise
in the supply power, temperature logger and change in heat transfer rate in the

101
electric heater due to the water temperature rise. Usually, an average value was used
in the analysis and error might be induced. The far field temperature (T0 ) was
normally estimated from the borehole temperature or the ground temperature
somewhere in the borefield within a period before the start of the test. The
appropriateness of the estimation method might also lead to a certain degree of error.
To investigate their effects on the estimated parameters, analysis was made using
LSMGA with Qb and T0 in the parameter estimation scheme assuming values

different from the default values of 30W/m for Qb and 20oC for T0 as adopted in the
generated data. The details were shown in Table 6.21&22 for groundwater velocities
of 5x10-7 and 10-6m/s respectively.

Table 6.21 Effects of Qb and T0 on parameter estimation results using LSMGA on


generated data with groundwater velocity 5x10-7m/s

Values used in parameter Average analysis results based on data generated


estimation scheme with groundwater velocity 5x10-7m/s
Qb (W/m) T0 (oC) kg (W/mK) Rb (mK/W) Vgw (m/s)
25 20 2.902 0.1549 4.90x10-7
35 20 4.098 0.1046 5.07x10-7
30 18 3.499 0.1919 5.05x10-7
30 22 3.499 0.0586 5.01x10-7

Table 6.22 Effects of Qb and T0 on parameter estimation results using LSMGA on


generated data with groundwater velocity 10-6m/s

Values used in parameter Average analysis results based on data generated


estimation scheme with groundwater velocity 10-6m/s
Qb (W/m) T0 (oC) kg (W/mK) Rb (mK/W) Vgw (m/s)
25 20 2.902 0.1549 9.29x10-7
35 20 4.098 0.1046 1.07x10-6
30 18 3.500 0.1920 1.00x10-6
30 22 3.502 0.0587 9.97x10-7

102
It could be observed that T0 had negligible effect on the estimated k g and Vgw.

The reason was that with different T0 , the borehole temperature rise profile would

only shift up and down without changing the shape of the curve. Hence, only Rb
would be affected which was similar to what was discussed in Section 6.2.3.3. Of
course, T0 was assumed to be constant throughout the test, which was appropriate in
actual TRT where a long borehole was used and the boundary ground temperature
was stable below certain depths. However, the situation could be different when
laboratory TRT was conducted where the boundary temperature might vary more
substantially. For Qb , the influence was more significant for k g and Rb . A 16%

change in Qb led to a 14% change in k g and at least 17% change in Rb . However,

the estimated Vgw varied only by not more than 7.1% which became smaller for a
lower groundwater velocity. As a 16% error in Qb was unlikely, the actual effect on
V would be very small.

6.2.4 Analysis of Experimental Thermal Response Test Results

As mentioned in Section 4.1, the borehole temperature was recorded by a


temperature logger. It was found that the logger had an inherent noise source
corresponding to a fluctuation of within 0.1oC over the measuring range. In some
situation as was indicated in the logged temperature profiles shown in Appendix O,
there was substantial perturbation in certain points of the recorded data. To eliminate
these effects, the borehole temperature was logged with a time interval of 3 minutes.
The parameter estimation would be based on the adjusted hourly recorded data, using
the median between the hourly logged data and those within 3 minutes before and
after the hour being considered.

Six formal tests were conducted with the applied conditions given in Table
6.23. As stated in Section 4.3, the groundwater flow was measured after each test,
and the groundwater velocity calculated accordingly. The loading was determined
by averaging the recorded values over the entire test period which was around 4 days
for each test. The far field temperature was estimated by averaging the logged
borehole temperature over one day before the start of each test.

103
Table 6.23 Applied conditions for different experimental TRTs

Test A Test B Test C Test D Test E Test F


Groundwater flow (l/min) 260 260 185 150 100 70
Groundwater velocity 6.77 6.77 4.82 3.91 2.60 1.82
(10-6m/s)
Average loading (W) 9.708 8.245 8.249 8.284 8.180 5.878
o
Far field temperature ( C) 24.11 24.39 23.37 22.78 23.48 23.19

As already shown in Figure 6.37, the influence of groundwater in TRT was


weak at the early stage of test. Also, it usually needed some time in order to achieve
steady state inside the borehole and that a constant Rb could be assumed for use in
Eq. (2.3). Hence, analysis with different starting recorded data would be tried so that
some of the recorded data at the early stage of the test period would not be used in
the determination of the MSD.

6.2.4.1 Parameter Estimation Results of Experimental TRT

Table 6.24 showed the analysis results. Basically, the results varied with the
starting point of the analysed data, but they seemed to approach certain limits if more
of the starting data record was discarded except Test B where the estimated
groundwater velocity with analysis starting at Test Record 8 appeared to be off-track.
Nevertheless, the results of the analysis starting at Test Record 10, i.e. skipping the
earlier 9 hours of data, were taken as the representative ones. The estimated k g

ranged from 2.399 to 2.602 while Rb ranged from 0.2461 to 0.2826. The percentage
differences were less than 15% when compared with the lowest values, showing that
the method yielded similar results for k g and Rb at different groundwater flow

velocities. Table 6.25 compared the estimated groundwater velocities with the actual
values. The percentage errors were large except for Test C, and the error seemed
smaller with a higher applied load as indicated in the results of Tests A & B.
However, a higher applied load was not used at other groundwater flow since the
logged borehole temperature as shown in Appendix O in Test A slightly exceeded
the calibration range of the logger which was 20~40oC.

104
Table 6.24 Analysis results of experimental TRT data

Test Estimated Starting data record to analyse


ref. parameter 1 2 4 6 8 10
kg (W/mK) 3.176 2.833 1.959 2.662 2.360 2.598
A Rb (mK/W) 0.2598 0.2505 0.2135 0.2479 0.2371 0.2461
Vgw (10-6m/s) 2.19 1.49 1.99 4.84 4.56 4.88
kg (W/mK) 4.297 3.884 1.876 2.012 2.249 2.602
B Rb (mK/W) 0.2823 0.2775 0.224 0.2272 0.2284 0.2478
Vgw (10-6m/s) 16.6 15.9 9.44 8.52 4.45 8.99
kg (W/mK) 2.630 2.53 3.526 1.673 2.453 2.399
C Rb (mK/W) 0.2630 0.2601 0.2829 0.2193 0.2583 0.2496
Vgw (10-6m/s) 8.32 8.34 9.02 6.60 5.60 4.31
kg (W/mK) 3.083 2.873 2.341 2.914 2.279 2.534
D Rb (mK/W) 0.2735 0.2695 0.2520 0.2677 0.2468 0.2551
Vgw (10-6m/s) 10.8 11.0 8.85 8.97 7.27 7.13
kg (W/mK) 1.623 1.801 2.534 2.489 2.575 2.510
E Rb (mK/W) 0.2271 0.2438 0.2843 0.2822 0.2851 0.2826
Vgw (10-6m/s) 0.117 2.17 5.00 4.25 3.87 3.40
kg (W/mK) 1.855 1.834 2.419 2.745 2.493 2.519
F Rb (mK/W) 0.2316 0.2306 0.2652 0.2769 0.2697 0.2703
Vgw (10-6m/s) 0.153 0.921 4.25 4.60 5.14 4.74

Table 6.25 Percentage error of estimated groundwater flow based on experimental


TRT data

Percentage error of estimated groundwater flow


Test A Test B Test C Test D Test E Test F
-27.93 32.77 -10.54 82.54 30.57 160.01

6.2.4.2 Laboratory Soil Thermal Conductivity Measurement Test Results

Table 6.26 indicated the measured temperature difference across the


sandwiched soil at different heater loads. The corresponding thermal characteristic

105
curve was shown in Figure 6.38. Under steady state conditions, the temperature drop
across the sandwiched soil should be proportional to the heat flux passing through
the soil. By applying linear regression method, the calculated k g was 3.73W/mK.

Table 6.26 Temperature differences across soil at different applied loads in


laboratory soil thermal conductivity measurement test

Copper plate Copper plate Calibrated


Applied temperature with temperature temperature difference
load (W) heater (oC) without heater (oC) across soil (oC)
10 34.5 33.4 1.2
15 37.5 36.0 1.5
20 39.2 37.2 2.0
25 43.5 41.1 2.3
30 49.0 46.0 2.8

Thermal conductivity test of soil

35
30
Applied load (W)

25
20
15
10
5
0
1 1.5 2 2.5 3

Temperature difference across soil (K)

Figure 6.38 Variation of applied load with temperature difference across soil in
laboratory soil thermal conductivity measurement test

6.2.4.3 Source of Experimental Errors

In the laboratory TRT, large errors were found in the estimated groundwater
flow in most cases. The major reason was that the boundary temperature of the test
tank, as reflected by the circulating water temperature indicated in Appendix O,

106
could not be maintained constant over the test period. This was different from the
actual site TRT condition. Although efforts had been made, including applying
insulation to the test tank, providing an electric heater inside test room, and passing
the circulating water through the constant-temperature water bath inside the
temperature calibrator before entering the test tank, the result was still unsatisfactory.
This inevitably affected the estimated groundwater velocity. The solving of the
problem required the re-design and erection of a new experimental setup, which was
not feasible within the limited time schedule in the study.

Another cause of error came from the size of the test tank. The distance
travelled by the heat front as calculated from the numerical model was far longer
than the width of the test tank within the test period. Thus, the actual heat transfer
pattern was disturbed and different from the one simulated by the numerical model.
The difference was large at small groundwater flow as indicated by the high
percentage error of estimated groundwater flow for Test F, and reduced with an
increasing groundwater velocity. It might be contemplated if better results could be
obtained by testing with a much higher groundwater flow. However, the hydraulic
conductivity and depth of the soil limited the groundwater flow that could be
modelled, and it was found that 260 l/min was basically the maximum that could be
achieved.

The laboratory-determined ground thermal conductivity was higher than the


average estimated value from the experimental TRT results by 47.66%. The main
reason was that all the heat generated from the electric heater was assumed to have
transferred through the soil. This was definitely not the real situation. Although the
back side of the electric heater and the acrylic frame was insulated, part of the heat
energy would still leak to the surrounding. Unfortunately, there was no way to
determine the magnitude of this leakage energy. Hence, the determined value could
only be considered as an upper limit. The resolution of the thermocouple reader was
0.1oC. This was relatively large as compared with the measured temperature
difference across the soil (only 1.1oC at a 10W heating load). The uncertainty could
account for several percents of error in the calculated k g from the measured data.

Finally, it might actually need a much longer time to achieve steady state in each

107
measurement. Thus the assumption of attainment of steady state led to errors in the
recorded temperatures.

6.2.5 Analysis of Site Thermal Response Test Results Provided by Others

To verify the new parameter estimation scheme in actual TRT analysis, the
site TRT data provided by Pahud (2006) and Spitler (2006) were used. One set of
data from Pahud (2006) was provided from a test near Luzern, Switzerland at the
north borehole (Pahud 2001), while four sets of data from Spitler (2006) were
obtained from tests at two schools (two tests for each school), Campbell and Maxey,
in Lincoln, Nebraska, USA. Analysis would be based on the hourly recorded data.

Table 6.27 Parameter estimation results of TRT data at north borehole of Luzern
using different models

Average Starting record data to analyse


estimated
Model parameters 1 2 4 6 8 10
kg (W/mK) 2.907 2.921 2.920 2.894 2.894 2.885
LSMGA Rb (mK/W) 0.0830 0.0835 0.0834 0.0826 0.0825 0.0822
Vgw (10-9m/s) 1.58 12.3 2.22 111 19.4 12.0
kg (W/mK) 2.889 2.906 2.903 2.884 2.874 2.864
FLSMGA Rb (mK/W) 0.0830 0.0989 0.1159 0.1259 0.1332 0.1390
Vgw (10-9m/s) 6.85 8.10 7.33 32.4 5.85 4.77
kg (W/mK) 2.903 2.920 2.914 2.898 2.884 2.871
Present Rb (mK/W) 0.0858 0.0864 0.0862 0.0857 0.0852 0.0847
Vgw (10-9m/s) 8.93 8.34 27.3 33.3 2.77 8.14

Table 6.27 summarised the analysis results of TRT data provided by Pahud
(2006) using various models of ground heat exchangers under different starting data
records. The estimated V varied from 1.58x10-9 to 1.11x10-7m/s. The testing period
was 5 days. According to Section 6.2.3.1, V should be more than 2x10-7m/s for
confident results. Hence, it could be deduced that the actual groundwater flow was
below this level. Indeed, Pahud (2001) stated that no groundwater flow was present

108
at that site, which agreed with the analysis results. The estimated k g varied mildly,

with an average value of 2.896W/mK, which was close to the value of 3W/mK
determined based on the LSM approach by Pahud (2001). According to Pahud
(2001), laboratory tests were conducted on soil samples collected at different depths
of the site, yielding an average value of 3.5W/mK. The estimated Rb changed very
little at different starting record data for LSMGA and the present numerical model,
but the trend with FLSMGA was completely different where Rb increased
substantially when more record data at beginning of test were skipped. If excluding
those using FLSMGA, the average estimated Rb was 0.0843mK/W which was not
far from the value of 0.088mK/W determined by Pahud (2001). Generally speaking,
the estimated results from the new parameter estimation scheme agreed with the
findings from Pahud (2001).

Tables 6.28&29 indicated the analysis results based on TRT data from Spitler
(2006) for Maxey School. The average estimated values of k g and V gw differed in

the two particular tests. For the test conducted on 18th September 1998, they were
1.915W/mK & 1.5x10-6m/s respectively, and the corresponding values for the test
conducted on 30th June 1999 were 2.29W/mK & 1.14x10-6m/s respectively. The
inhomogeneity in the ground might account for the difference in the estimated k g , as

the tests were conducted using separate boreholes at some distance apart. The
difference in the estimated V gw might reflect the change in the groundwater condition

at site, as the dates for the two tests differed by nearly 9 months. Again, the trend for
the estimated Rb was different with FLSMGA as compared with those from other
numerical models.

109
Table 6.28 Parameter estimation results of TRT data conducted on 18th September
1998 for Maxey School at Lincoln using different models

Average Starting record data to analyse


estimated
Model parameters 1 2 4 6 8 10
kg (W/mK) 1.979 1.934 1.827 1.806 1.954 2.025
LSMGA Rb (mK/W) 0.1087 0.1070 0.1025 0.1016 0.1083 0.1114
Vgw (10-9m/s) 1.39 1.47 1.62 1.65 1.48 1.39
kg (W/mK) 1.972 1.928 1.820 1.799 1.948 2.012
FLSMGA Rb (mK/W) 0.1088 0.1307 0.1540 0.1695 0.1828 0.1918
Vgw (10-9m/s) 1.37 1.46 1.62 1.64 1.46 1.39
kg (W/mK) 1.957 1.933 1.830 1.803 1.935 2.001
Present Rb (mK/W) 0.1105 0.1096 0.1053 0.1040 0.1102 0.1132
Vgw (10-9m/s) 1.40 1.45 1.61 1.65 1.49 1.40

Table 6.29 Parameter estimation results of TRT data conducted on 30th June 1999
for Maxey School at Lincoln using different models

Average Starting record data to analyse


estimated
Model parameters 1 2 4 6 8 10
kg (W/mK) 2.293 2.300 2.345 2.270 2.285 2.321
LSMGA Rb (mK/W) 0.0904 0.0905 0.0920 0.0895 0.0901 0.0913
Vgw (10-9m/s) 1.17 1.15 1.05 1.20 1.18 1.12
kg (W/mK) 2.285 2.292 2.337 2.267 2.274 2.323
FLSMGA Rb (mK/W) 0.0903 0.111 0.1338 0.1458 0.1553 0.1631
Vgw (10-9m/s) 1.14 1.13 1.03 1.17 1.17 1.07
kg (W/mK) 2.241 2.268 2.315 2.248 2.259 2.303
Present Rb (mK/W) 0.0909 0.0917 0.0933 0.091 0.0914 0.093
Vgw (10-9m/s) 1.23 1.16 1.05 1.19 1.18 1.11

Table 6.30 Parameter estimation results of TRT data conducted on 15th September
1998 for Campbell School at Lincoln using different models

110
Average Starting record data to analyse
estimated
Model parameters 1 2 4 6 8 10
kg (W/mK) 2.004 2.045 2.140 2.164 2.178 2.149
LSMGA Rb (mK/W) 0.1338 0.1355 0.1388 0.1401 0.1408 0.1391
Vgw (10-9m/s) 89.9 76.9 23.5 8.80 3.43 4.61
kg (W/mK) 1.996 2.047 2.125 2.149 2.162 2.129
FLSMGA Rb (mK/W) 0.134 0.1584 0.1826 0.1996 0.2102 0.2173
Vgw (10-9m/s) 87.6 68.9 24.9 4.43 1.97 5.59
kg (W/mK) 1.973 2.033 2.101 2.135 2.146 2.115
Present Rb (mK/W) 0.1357 0.1379 0.1405 0.1419 0.1426 0.1408
Vgw (10-9m/s) 92.8 69.9 26.0 3.03 3.24 2.51

Table 6.31 Parameter estimation results of TRT data conducted on 8th October 1998
for Campbell School at Lincoln using different models

Average Starting record data to analyse


estimated
Model parameters 1 2 4 6 8 10
kg (W/mK) 1.894 1.938 2.021 2.068 2.073 2.064
LSMGA Rb (mK/W) 0.1041 0.1059 0.1092 0.1113 0.1116 0.1111
Vgw (10-9m/s) 937 791 438 13.6 42.8 9.86
kg (W/mK) 1.886 1.930 2.013 2.054 2.058 2.047
FLSMGA Rb (mK/W) 0.1041 0.1297 0.1573 0.1733 0.1841 0.1922
Vgw (10-9m/s) 919 770 394 6.23 45.9 70.4
kg (W/mK) 1.872 1.924 1.996 2.039 2.044 2.035
Present Rb (mK/W) 0.1064 0.1084 0.1113 0.1133 0.1137 0.1131
-9
Vgw (10 m/s) 955 772 422 30.8 17.4 54.8

Tables 6.30&31 showed the corresponding results for Campbell School. The
estimated V gw in the two tests varied considerably from 9.86x10-9 to 9.55x10-7m/s.

In fact, the estimated V gw scattered very much in each trial. As the duration of each

test was only 50 hours, the actual groundwater flow, if any, was probably outside the

111
range for confident estimated result which was similar to the situation when using
Pahuds data. The average estimated k g for the two tests, i.e. 2.1 and 1.998W/mK,

differed only mildly. Neglecting results using FLSMGA, the average estimated
values of Rb at 0.139 and 0.11mK/W were substantially different for the two tests.
This might be due to the structural difference induced during the construction of the
boreholes including orientation of U-tube, workmanship of backfilling, etc. In
practice, the thermal resistance for each borehole in a borefield would not be exactly
the same.

6.3 Modelling of Ground-coupled Liquid Desiccant Air Conditioners

6.3.1 Determination of Final Configuration of Liquid Desiccant Dehumidifying


Cycle

Desiccant Desiccant Regenerative


cooler heater air fan Trai
TL,di TL,dci TL,dhi TL,ri
Tpa,do Tra,ro

Desiccant
heat Regenerative
Process Dehumidifier exchanger Regenerator air heat
air fan tower tower exchanger

Weak Strong
Tpa,di Tra,ri Trahi
desiccant desiccant
pump pump
TL,do TL,ro
Regenerative Trao
air heater

Figure 6.39 Schematic of advanced liquid desiccant dehumidifier

For a ground-coupled liquid desiccant air conditioner, the design concern was
the total enthalpy drop rather than just the dehumidification rate. Hence, the
selection of components for the liquid desiccant dehumidifying cycle might not be
exactly the same as that for a pure liquid desiccant dehumidifier. Figure 6.39
showed a schematic diagram of a more advanced liquid desiccant dehumidifier
which incorporated additional components including a desiccant cooler, regenerative

112
air heat exchanger and regenerative air heater as compared with the basic system as
shown in Figure 1.4. The effect of incorporating each additional component into the
basic system on the dehumidification rate and enthalpy drop of process air would be
investigated to decide if a particular component would be used. The parameters
values set in Table 5.2 were used in the analysis.

6.3.1.1 Effect of Desiccant Cooler on Performance of Liquid Desiccant Dehumidifier

Table 6.32 summarised the change of dehumidification rate and process air
cooling capacity at different desiccant cooling loads. Clearly, the introduction of
desiccant cooling rendered both the dehumidification rate and the cooling capacity to
improve. Hence, desiccant cooler would be used.

Table 6.32 Effect of desiccant cooler on performance of liquid desiccant


dehumidifier

Desiccant Process air Process air Dehumidification Process air


cooling discharge discharge rate (g/s) cooling
capacity (kW) temperature relative capacity (kW)
(oC) humidity (%)
0 37.71 51.27 0.567 -0.120
1 35.22 56.91 0.712 0.876
2 32.26 65.50 0.837 1.892
3 29.30 75.59 0.953 2.903
4 26.34 87.57 1.066 3.902

6.3.1.2 Effect of Regenerative Air Heater on Performance of Liquid Desiccant


Dehumidifier

Table 6.33 showed the variation of liquid desiccant dehumidifier performance


at different regenerative air heating loads under a constant desiccant cooling load of
4kW. The dehumidification rate increased with the regenerative air heating load, but
the additional sensible heat transferred to the process air exceeded that gained from
the increased latent cooling. Hence, the total cooling capacity of process air dropped,

113
which was not beneficial for an air conditioner. In view of this, regenerative air
heater was discarded.

Table 6.33 Effect of regenerative air heater on performance of liquid desiccant


dehumidifier

Regenerative Process air Process air Dehumidification Process air


air heater discharge discharge rate (g/s) cooling
capacity (kW) temperature relative capacity (kW)
(oC) humidity (%)
0.5 27.29 82.36 1.089 3.737
1.0 28.22 77.57 1.115 3.583
1.5 29.07 73.79 1.117 3.385
2.0 30.13 68.72 1.161 3.249

6.3.1.3 Effect of Regenerative Air Heat Exchanger on Performance of Liquid


Desiccant Dehumidifier

Table 6.34 Effect of regenerative air heat exchanger on performance of liquid


desiccant dehumidifier

Regenerative air Process air Process air Dehumidification Process air


heat exchanger discharge discharge rate (g/s) cooling
overall transfer temperature relative capacity (kW)
coefficient (oC) humidity
(kW/K) (%)
0.2 29.23 72.78 1.137 3.401
0.4 30.57 66.80 1.176 3.184
0.6 31.41 63.31 1.199 3.042
0.8 32.11 60.65 1.213 2.914

Table 6.34 indicated the effect of regenerative air heat exchanger on the
dehumidification rate and process air cooling capacity of liquid desiccant
dehumidifier without regenerative air heating and under a constant desiccant cooling
load of 4kW. The trend was similar to that using a regenerative air heater. For the

114
same reason, regenerative air heat exchanger would also not be used in the liquid
desiccant air conditioner, and Trai became Tra , ri .

6.3.2 Performance of Heat-pump-coupled Liquid Desiccant Dehumidifier


(HPCLDD)

As the finalised liquid desiccant dehumidifier included desiccant cooling and


heating, the installation of a vapour compression heat pump would meet both
requirements, as shown in Figure 6.40. The desiccant heater became a condenser of
the heat pump and the desiccant cooler worked as an evaporator.

Desiccant Desiccant
cooler heater
TL,dci TL,dhi
Process Regenerative
Desiccant
air fan heat air fan
TL,di exchanger TL,ri
Tpa,do Tra,ro

Dehumidifier Regenerator
tower tower
Weak Strong
desiccant desiccant
Tpa,di Tra,ri
pump pump

TL,do TL,ro

Compressor

Expansion
valve

Figure 6.40 Schematic diagram of heat-pump-coupled liquid desiccant dehumidifier

Comparison was made with a conventional air-cooled fresh air unit (ACFAU)
based on the weather data for 1986, as shown in Figures 6.41~43 in respect of the
results obtained from simulation carried out only when the outdoor temperature
exceeded 20oC. For ACFAU, the COP was lowest in the peak load season where the
outdoor temperature was high. On the other hand, HPCLDD tended to be more
energy efficient in the peak load season, which was beneficial in terms of the overall
energy input.

115
Outdoor temperature profile for 1986

Outdoor temperature (deg C)


40
35
30
25
20
15
10
5
0
1 731 1461 2191 2921 3651 4381 5111 5841 6571 7301 8031

Time (hour)

Figure 6.41 Outdoor temperature profile for 1986

Performance of ACFAU for 1986

4
3.8
3.6
3.4
3.2
COP

3
2.8
2.6
2.4
2.2
2
1 731 1461 2191 2921 3651 4381 5111 5841 6571 7301 8031

Time (hour)

Figure 6.42 Performance of ACFAU for 1986

By comparing Figures 6.42 and 6.43, the COP of ACFAU was still higher
than that for HPCLDD in most cases. This was probably due to the effectiveness of
the liquid desiccant cycle. By using a smaller packing size of 0.015m for dp for both
the dehumidifier and regenerator towers, the heat and mass transfer rate was
enhanced, and the simulation result was shown in Figure 6.44. The performance was
improved and generally better than that for ACFAU throughout peak load period.

116
Performance HPCLDD with dp=0.02 for 1986

4
3.8
3.6
3.4
3.2
COP

3
2.8
2.6
2.4
2.2
2
1 731 1461 2191 2921 3651 4381 5111 5841 6571 7301 8031

Time (hour)

Figure 6.43 Performance of HPCLDD with dp=0.02 for 1986

Performance of HPCLDD with dp=0.015 for 1986

4
3.8
3.6
3.4
3.2
COP

3
2.8
2.6
2.4
2.2
2
1 731 1461 2191 2921 3651 4381 5111 5841 6571 7301 8031

Time (hour)

Figure 6.44 Performance of HPCLDD with dp=0.015 for 1986

When performing the simulation of the vapour compression cycle, an


analytical model similar to that used by most previous researchers was first tried for
the compressor for which the compression process was assumed to be polytropic.
The various parameters for the compressor were then determined using parameter
estimation technique as adopted by Jin and Spitler (2002) using performance data
from the manufacturer (www.emersonclimate.com). However, the estimated
parameters were scattered at repeated trials, and the simulated performance of the
compressor using the estimated parameters gave strange trends in the isentropic
efficiency as compared with the manufacturers data. In view of this, empirical

117
formulation of compressor performance provided by the manufacturer was used
instead, which would restrict the simulation of the vapour compression cycle at the
specified degree of superheat of 5.56oC at the compressor suction, and that only
thermostatic expansion valve could be used for the expansion device.

Initially, an approximate iterative method was used for the evaporator of the
ACFAU. However, it was found that under certain outdoor conditions, the
simulation time was very long as compared with the finite difference method. This
was probably due to the double iterations in the method, leading to higher chance of
oscillation in the simulated results. To ensure stable and fast dynamic system
simulation, the finite difference method would be used for the air cooler in the
subsequent analysis.

6.3.3 Analysis of Ground-coupled Liquid Desiccant Air Conditioner (GCLDAC)


and Comparison with Conventional Ground Source Heat Pump (GSHP)

6.3.3.1 Ultimate Configuration of GCLDAC

With the fresh air being handled by the liquid desiccant cycle, auxiliary
components were needed to treat the mixed air in order to meet the complete air
conditioning requirement. This could be achieved by coupling the liquid desiccant
system with a conventional geothermal heat pump system, as shown in Figure 6.45
which indicated cooling mode operation. Two selective valves were used to control
the refrigerant flow direction for normal or reverse operation of the system to provide
cooling or heating where necessary. Two 3-way bypass valves were added so that
the refrigerant would only pass the desiccant cooler/heater in the cooling mode,
allowing the dehumidifying cycle to have the best performance as mentioned in
previous section. The desiccant pumps and regenerative air fan were turned off
when no cooling was required, and the borefield circulating pump stopped when
there was no air conditioning demand for either cooling or heating. The supply air
fan was running within the entire operating schedule.

118
Trta Desiccant Desiccant
cooler heater

Supply Supply air coil TL,dci TL,dhi Regenerative


Desiccant
air fan heat air fan
TL,di exchanger TL,ri
Tsa Tma Tpa,do Tra,ro

Dehumidifier Regenerator
tower Weak Strong tower
desiccant desiccant
Tpa,di Tra,ri
pump pump
Ground
TL,do TL,ro coupled coil

Selector
valve
Compressor

Selector
valve Expansion
valve
Tbf,f,out

Borefield
Tbf,f,in
Borefield circulating
pump

Figure 6.45 Schematic diagram of ground-coupled liquid desiccant air conditioner


(cooling mode)

6.3.3.2 Determination of Design Duty of GCLDAC/GSHP

Before the design parameters for the GSHP & GCLDAC could be set, the
loading requirement had to be estimated first. Figures 6.46&47 showed the room
loading demand based on the weather data from 1986 to 1995. The latent load was
stable throughout the period, as occupants were the only source of latent load. The
room loading demand was used to determine the supply air condition which was
assumed to be saturated moist air neglecting reheat through the process air fan and
air duct corresponding to the design indoor condition of 24oC and 54%RH. The peak
sensible and latent loads were used for this purpose as only one supply air
temperature could meet exactly the required sensible to latent load ratio. The fresh
air requirement was calculated based on ASHRAE (2005). The loading requirement
for the unit was then evaluated by adding the room loading to the ventilation load
due to fresh air based on design outdoor condition of 31oC and 83%RH as shown in
Table 6.35.

119
Figure 6.46 Building load profile for Case 1

Figure 6.46 Building load profile for Case 2

120
Table 6.35 Design requirement of GCLDAC/GSHP

Case 1 Case 2
Peak room sensible load (kW) 7.102 7.745
Peak room latent load (kW) 1.044 1.606
Supply air flow (m3/s) 0.53 0.575
Supply air temperature (oC) 13.13 12.63
Fresh air amount (m3/s) 0.035 0.070
Fresh air ratio 0.066 0.122
Unit design sensible load (kW) 7.612 9.045
Unit design latent load (kW) 2.307 4.222
Unit design total load (kW) 9.919 13.27

6.3.3.3 Setting of Design Parameters of GCLDAC/GSHP

Table 6.36 Parameters used for GSHP

Case 1 Case 2
Overall heat transfer value of supply air coil, UAsa (kW/K) 1.5 1.9
Overall heat transfer value of ground coupled coil, UAgc 1.5 1.9
(kW/K)
Volume of refrigerant in supply air coil, VOLsa (m3) 0.006 0.008
3
Volume of refrigerant in ground coupled coil, VOLgc (m ) 0.006 0.008
Volume of refrigerant in liquid line, VOLll (m3) 0.0002 0.00025
Refrigerant charge, Mr (kg) 1.5 3.0
No. of discretisation segment for supply air coil, nsa 50 50
Borefield fluid mass flowrate, mbf (kg/s) 0.5 0.6
Compressor model used ZH30K4E ZH38K4E

121
Table 6.37 Parameters used for GCLDAC

Case 1 Case 2
Overall heat transfer value of desiccant cooler, UAL,dc (kW/K) 0.1 0.2
Overall heat transfer value of desiccant heater, UAL,dh (kW/K) 0.15 0.3
Overall heat transfer value of supply air coil in cooling mode, 1.3 1.7
UAsa,cool (kW/K)
Overall heat transfer value of supply air coil in heating mode, 1.5 1.9
UAsa,heat (kW/K)
Overall heat transfer value of ground coupled coil in cooling 1.3 1.7
mode, UAgc,cool (kW/K)
Overall heat transfer value of ground coupled coil in heating 1.5 1.9
mode, UAgc,heat (kW/K)
Volume of refrigerant in desiccant cooler, VOLdc (m3) 0.0005 0.0015
Volume of refrigerant in desiccant heater, VOLdh (m3) 0.0005 0.0015
3
Volume of refrigerant in supply air coil, VOLsa (m ) 0.0055 0.0065
Volume of refrigerant in ground coupled coil, VOLgc (m3) 0.0055 0.0065
Volume of refrigerant in liquid line, VOLll (m3) 0.0002 0.00025
No. of discretisation segment for supply air coil, nsa 50 50
Height of desiccant towers (m) 0.5 0.6
No. of discretisation segment for desiccant towers 50 50
Cross-sectional area of dehumidifier tower, Sdeh (m2) 0.0225 0.0625
Packing size of dehumidifier tower, dp,deh (m) 0.01 0.015
Cross-sectional area of regenerator tower, Sreg (m2) 0.0625 0.1225
Packing size for regenerator tower, dp,reg (m) 0.01 0.015
Regenerative air flow, V&ra (m3/s) 0.07 0.15

Overall heat transfer coefficient of desiccant heat exchanger, 1.0 1.2


UAL,he (kW/K)
Liquid desiccant (LiCl) solution volume flowrate, V&L (m3/s) 0.00004 0.00009

Ratio of refrigerant to desiccant cooler 0.1 0.15


Ratio of refrigerant to desiccant heater 0.1 0.15

122
The next step involved the selection of the various coil, tower and refrigerant
properties. In order to allow the possibility of reverse cycle operation, the volume
and overall heat transfer values of the condenser and evaporator were chosen to be
the same. For GCLDAC under the cooling mode, the condenser was composed of
two coils, namely the ground-coupled coil and desiccant heater while the evaporator
consisted of the supply air coil and desiccant cooler. The overall heat transfer values
of coils reflected the mean temperature difference between the two fluids, and were
selected so as to compromise between energy efficiency and equipment cost. In this
analysis, the values were chosen so that the mean temperature difference would be
between 5 to 10oC. The parameters for the desiccant cycle were selected so that the
loading of the two systems would be similar under comparable coil properties and
with the same refrigerant charge. Tables 6.36&37 summarised the values of various
parameters to be used for GSHP and GCLDAC. A single borehole would be used in
order to be able to neglect the effect of groundwater direction as mentioned in
Section 6.1.6. To increase the capacity of the borehole, a double U-tube with the 1-
3,2-4 connection configuration as recommended by Zeng et al. (2003) was adopted
with other borehole parameters being the same as those indicated in Table 5.1 except
H, N, mf and QAC.

6.3.3.4 Performance of GCLDAC/GSHP under Design Conditions

Based on the design parameters specified in previous section, the


performance of GCLDAC/GSHP under the design indoor/outdoor conditions was
simulated and compared, with the results shown in Tables 6.38&39. The borefield
leaving fluid temperature was assumed to be 30oC. The capacities of both systems
were very close, but the load transferred to the borefield using GCLDAC was smaller.
A 9% reduction was achieved for Case 1 while there was a 14.5% reduction for Case
2 in which the fresh air ratio was higher. The potential of the new hybrid system was
thus clearly indicated. However, the system performance depended on the operation
conditions, and an annual dynamic simulation was thus needed for an overall
evaluation of the two systems for long-term use.

123
Table 6.38 Performance of GSHP/GLCDAC at design conditions for Case 1

GSHP GCLDAC
Supply air condition (DB/RH) (oC/%) 12.80/100 12.85/100
Borefield fluid entering temperature (oC) 35.87 35.26
Cooling capacity (kW) 10.07 9.98
Load transferred to borefield (kW) 12.10 10.88
COP 4.97 4.94

Table 6.39 Performance of GSHP/GLCDAC at design conditions for Case 2

GSHP GCLDAC
Supply air condition (DB/RH) (oC/%) 12.61/100 12.61/100
Borefield fluid entering temperature (oC) 36.18 35.26
Cooling capacity (kW) 12.74 12.74
Load transferred to borefield (kW) 15.29 13.08
COP 5.11 5.15

6.3.3.5 Dynamic Simulation of GCLDAC/GSHP at No Groundwater Flow

To perform dynamic system simulation, a thermostat was used to control the


operating mode of the GCLDAC/GSHP. The cooling setpoint was at 24oC and that
for heating at 20oC. The deadband was set to 2oC, meaning that the hysteresis loops
for cooling and heating were between 23~25oC and between 19~21oC respectively.
To avoid too long a computation time, the simulation time step was taken to be 15
minutes. Analysis was made using the Hong Kong weather data from 1986 to 1995.
As already mentioned in the previous sections, the borefield fluid leaving
temperature was the parameter to consider when designing the borefield. Hence, the
borehole length was selected so that the maximum borefield fluid leaving
temperature within the simulation period remained the same. The undisturbed
ground temperature was assumed to be 20oC.

124
Table 6.40 Simulation results of GSHP with borehole length 250m at no
groundwater flow for Case 1 within 1986~1995

After 1 year After 10 years


Total cooling energy to room (kWh) 32,570.25 336,081.78
Total heating energy to borefield (kWh) 39,521.99 408,519.94
Total energy input to compressor (kWh) 6,485.50 67,631.04
Maximum borefield fluid leaving 30.48 31.67
temperature (oC)

Performance of GSHP at no groundwater flow for Case 1


35
33
Temperature (degC)

31
29
27
25 Room
23 T bf,f,out
21
19
17
15
0 20000 40000 60000 80000

Simulation time (hour)

Figure 6.47 Performance of GSHP at no groundwater flow for Case 1 within


1986~1995

The first trial was made for Case 1 using GSHP with a borehole length of
250m for ten years from 1986 to 1995 at no groundwater flow, and the results were
shown in Table 6.40 and Figure 6.47. The total energy transferred after the first year
was smaller than the average value over the entire ten-year period, and heating was
required for each year with the demand being highest in the first year as reflected by
a lower borefield leaving fluid temperature. Based on the results from the first trial,
the borehole length would be chosen to have the maximum borefield fluid leaving
temperature maintained at around 31.67oC after ten years simulation in all ongoing
simulations.

125
Table 6.41 Simulation results of GCLDAC with borehole length 225m at no
groundwater flow for Case 1 within 1986~1995

After 1 year After 10 years


Total cooling energy to room (kWh) 32,478.96 335,036.66
Total heating energy to borefield (kWh) 35,496.51 365,927.34
Total energy input to compressor (kWh) 6,522.92 67,989.45
Maximum borefield fluid leaving 30.54 31.71
temperature (oC)

Performance of GCLDAC at no groundwater flow for


Case 1
35
33
Temperature (degC)

31
29
27
Room
25
23 T bf,f,out
21
19
17
15
0 20000 40000 60000 80000

Simulation time (hour)

Figure 6.48 Performance of GCLDAC at no groundwater flow for Case 1 within


1986~1995

Table 6.41 and Figure 6.48 summarised the corresponding results for
GCLDAC with a borehole length of 225m upon repeated trials in order to maintain
the same maximum borefield fluid leaving temperature. The total cooling energy to
room was around 0.3% lower than that using GSHP after one year and ten years,
which could be considered to be the same, while the estimation based on design
conditions as mentioned in previous section indicated a reduction of less than 0.1%.
The potential saving in borehole length under the same capacity of both systems was
10%. The total energy transferred to borefield was 10.2% and 10.4% less than that
with GSHP after one year and ten years respectively. The total energy input to the
compressor was only 0.5% higher for GCLDAC which was comparable to around
0.6% difference in COP based on the design conditions as shown in Table 6.38. The

126
said difference was very small. Consequently, the new hybrid system was capable of
maintaining the energy efficiency of GSHP.

By reviewing Tables 6.40&41, the maximum borefield fluid leaving


temperatures after one year were also very close for both systems, meaning that
comparison of system performances could be made with one year simulation only.
This was definitely useful as the computation time for GCLDAC was extremely long
due to the slow convergence in the desiccant loop. It took more than 9 days to
perform the dynamic simulation for ten years using GCLDAC for Case 1, while less
than 10 hours was required with GSHP. Hence, in subsequent analysis, ten-year
simulations would be made for GSHP under different situations for determining the
borehole length, and the corresponding study for GCLDAC carried out with one-year
simulations based on the same maximum borefield leaving fluid temperature after
one year for GSHP.

Table 6.42 summarised the simulation results for Case 2. The situation was
very similar to that for Case 1. The reduction in borehole length was 13.8% while
the total heat energy to borefield was 14.6% lower for GCLDAC. The total energy
input to compressor was 0.5% lower for GCLDAC, indicating that GCLDAC was
slightly more energy-efficient, which was in line with the higher COP for GCLDAC
at the design condition listed in Table 6.39.

Table 6.42 Simulation results of GSHP/GCLDAC at no groundwater flow for Case 2

GSHP for GSHP for GCLDAC


1986 1986~1995 for 1986
Total cooling energy to room (kWh) 44,868.01 462,838.44 44,821.61
Total heating energy to borefield (kWh) 54,339.32 561,622.06 46,428.94
Total energy input to compressor (kWh) 8,686.46 90,785.95 8,639.27
Maximum borefield fluid leaving 30.47 31.72 30.49
temperature (oC)
Borehole length (m) 325 325 280

127
Figure 6.49 depicted the performance of GSHP with ten year simulation for
Case 2. The profiles were similar to those for Case 1 as shown in Figures 6.47&48.
The room temperature profiles indicated that the units were suitably sized in both
cases in order to meet the design loadings. The heating period was short in each year
which was typical for sub-tropical area and appeared to be similar in both cases.

Performance of GSHP at no groundwater flow for Case 2


35
33
Temperature (degC)

31
29
27
25 Room
23 T bf,f,out
21
19
17
15
0 20000 40000 60000 80000

Simulation time (hour)

Figure 6.49 Performance of GSHP at no groundwater flow for Case 2 within


1986~1995

6.3.3.6 Dynamic Simulation of GCLDAC/GSHP with Groundwater Flow

Performance of GCLDAC/GSHP was simulated using groundwater velocities


of 10 , 5x10-7 and 10-6m/s respectively for the two cases, and the corresponding
-7

required borehole lengths were shown in Figures 6.50&51. More detailed simulation
data were given in Appendix O. The average reduction in borehole length for
GCLDAC was 10.11% in Case 1 and 14.31% in Case 2 with very minor fluctuations
at different groundwater velocities. Clearly, the reduction in borehole length with
the new hybrid system increased with the fresh air ratio. It could be expected that up
to 20% saving in borehole length could be reached with a fresh air ratio of 0.15
which was common for applications with higher occupant density and/or indoor air
quality requirement like schools or hospitals. Most importantly, the energy
efficiency of GSHP could still be maintained, which meant that GCLDAC would be
a useful potential alternative system for installation in sub-tropical regions.

128
Comparison of required borelength between GSHP &
GCLDAC for Case 1

270
250
Bore length (m)
230
GSHP
210
GCLDAC
190
170
150
0 2 4 6 8 10 12

Groundwater velocity (1E-7m/s)

Figure 6.50 Required borehole lengths of GCLDAC/GSHP at different groundwater


velocities for Case 1

Comparison of required borelength between GSHP &


GCLDAC for Case 2

350
Bore length (m)

300

GSHP
250
GCLDAC

200

150
0 2 4 6 8 10 12

Groundwater velocity (1E-7m/s)

Figure 6.51 Required borehole lengths of GCLDAC/GSHP at different groundwater


velocities for Case 2

129
Chapter 7 : Conclusions and Recommendations

7.1 Conclusions

1. A three-dimensional numerical model was developed for the ground heat


exchanger borefield of a geothermal heat pump system with ground water
advection based on an implicit finite difference method with a rectangular
coordinate system. Each borehole was approximated by a square column
circumscribed by the borehole radius, with the gridding scheme calibrated
against the cylindrical source model (CSM) and the inclusion of a load factor
of 1.047. The coupled heat transfer between the borehole and the
surrounding ground was solved iteratively, allowing the temperature and load
variations along the borehole to be accounted for.

2. The simulation results based on the present model with no groundwater flow
showed little difference from those obtained by using analytical models for a
single borehole. However, the bigger difference in the ground temperature
rise at locations distant from the borehole resulted in a larger deviation in
performance for a large borefield between the present model and the
analytical models with superposition. The borefield thermal resistance also
showed similar discrepancies. As the analytical models assumed a constant
loading along the boreholes which might not reflect the actual situations, the
present numerical model should be used especially when designing for a large
borefield.

3. The performance of a single borehole with groundwater advection indicated


that the temperature profile along the borehole was similar at different
groundwater velocities. This meant that the borehole thermal resistance
could be assumed to be independent of the groundwater flow, which was
important when including the groundwater velocity in the parameter
estimation list for thermal resistance test analysis.

130
4. The effect of loading profiles with the same peak cooling load on the
borehole performance at different groundwater flow conditions was studied.
The borehole leaving fluid temperature rise was highest for a continuous
constant load. The provision of a daily operating schedule reduced the
temperature rise substantially. The introduction of a periodic load lowered
the temperature rise further, but to a much lesser extent. The inclusion of a
heating period had insignificant effects. The reduction in the leaving fluid
temperature rise was much smaller than the decrease in the total heat
injection to the ground for the cases corresponding to Load Profiles 1 to 4.
Indeed, the effect of loading profiles diminished with the groundwater
velocity. At high groundwater flow conditions with a groundwater velocity
greater than 10-5m/s, the leaving fluid temperature rise became independent
of the loading profiles. This reflected the importance of the peak cooling
load in determining the borehole performance. The reduction in the peak
cooling load, such as by adopting a hybrid system to share the cooling load in
the peak-load period, could significantly reduced the required borehole length
especially when the groundwater velocity was high.

5. The effect of groundwater direction on the borefield performance was


investigated by employing the percentage variation (PV) which evaluated the
maximum percentage difference in the borefield fluid leaving temperature
rise at different groundwater directions. The effect of loading profiles on PV
was similar to that on the borefield leaving fluid temperature rise. For square
borefields, PV was negligible for a groundwater velocity smaller than 10-6m/s
and generally increased with the groundwater flow. For non-square
borefields, PV first rose, reached a maximum and then reduced with
increasing groundwater flow. In both cases, PV became higher for a larger
borefield. For a periodic load (zero heating load) with a daily operating
schedule, PV could reach 5% for a 10x10 borefield, 12% for 5x10 borefield
and 10% for 6x9 borefield. The reduction in borehole length with a correct
setting of the borefield orientation was substantial.

6. Thermal response test (TRT) analysis based on parameter estimation


technique using the present model was conducted to determine the ground

131
thermal conductivity, borehole thermal resistance and groundwater velocity
simultaneously. The results obtained by using the generated test data based
on a small groundwater flow indicated that the estimated groundwater
velocity was scattered for repeated trials. This was due to the fact that the
borehole performance was very close at small groundwater velocities within
the normal thermal response test period. More confident results on the
estimated groundwater velocity could only be achieved for a groundwater
velocity higher than 2x10-7m/s. This indicated that a confident minimum
existed for the estimated groundwater velocity above which the groundwater
velocity could be determined with confidence in the TRT analysis. Moreover,
an extension of the test period could reduce this confident minimum. For a
test period of 30 days, the groundwater velocity could be estimated with
confidence down to a level of 10-7m/s.

7. Experimental thermal response tests were conducted, and the results analysed.
However, the estimated groundwater velocities deviated much from the test
values, with errors ranging from 10.54% to 160.01%. The uncontrollable
boundary temperature around the test rig was one of the major reasons for
this wide error range. The small size of the test tank also affected the
temperature profiles in the soil that should be obtained with the numerical
model having much larger grid boundaries. The estimated ground thermal
conductivities ranged from 2.391 to 2.602 W/mK, while the laboratory soil
thermal conductivity measurements yielded an upper limit of 3.731W/mK.

8. A model for the heat-pump-coupled liquid desiccant dehumidifier (HPCLDD)


was developed. Performance analysis was made with the dehumidifier used
to treat the fresh air based on Hong Kong weather data for 1986, and
compared with that using the conventional air-cooled fresh air unit (ACFAU) .
The results showed that the highest energy efficiency was obtained during the
peak-load season for the HPCLDD, while the best performance was achieved
at the low-load period for the ACFAU. This rendered the HPCLDD superior
to the ACFAU when considering the annual system energy consumption.

132
9. A new hybrid system, the ground-coupled liquid desiccant air conditioner
(GCLDAC), was designed where the fresh air was treated by the liquid
desiccant cycle in cooling mode, and the mixed air handled by a conventional
geothermal heat pump system. Only a single compressor was needed to
provide the necessary desiccant cooling/heating and air-conditioning load for
the supply air. A single-zone sample building was used to analyse the
dynamic performance of the new hybrid system. The room air conditioning
load profile was generated using the TRNSYS software package based on the
Hong Kong weather data from 1986 to 1995. Two cases were studied, the
first one with 10 occupants corresponding to a fresh air ratio 0.066 and the
second one with 20 occupants corresponding to a fresh air ratio 0.122. The
simulated results were compared with those using only the conventional
ground source heat pump (GSHP) system in terms of the required borehole
length and total energy input to the compressor at different groundwater flow
conditions.

10. With the new hybrid system, the borehole length was reduced by 10.11% on
average for Case 1, while a 14.31% saving was achieved for Case 2, as part
of condensing heat was transferred to the regenerative air stream of the liquid
desiccant loop. Hence, for a design requiring a higher fresh air ratio as in
Case 2, more beneficial results could be obtained. The total energy input to
the compressor was nearly the same for both systems, meaning that the new
hybrid system did not cause deterioration in the energy efficiency of the
conventional GSHP. The groundwater flow did not affect the borehole length
reduction significantly.

7.2 Recommendations for Further Research Work

1. The present model assumes that the ground is homogeneous. In practice, the
soil may consist of several layers, each with its own physical properties. The
ground surface temperature also changes with time although the magnitude of
the annual fluctuation decreases rapidly with depth. To account for these
factors, modification of the present numerical scheme is required.

133
2. The groundwater may not always flow in the same direction if inhomogeneity
exists inside the ground. Indeed, the groundwater flow direction needs not
be horizontal, and that the difference in the groundwater temperature may
induce a buoyancy-driven flow in the vertical direction. A more precise
manipulation of the groundwater effect in the numerical model should be
made.

3. Piles are common for buildings in Hong Kong, and the embedding of the
ground heat exchangers into the piles can reduce the installation cost
substantially. To analyse the performance of energy piles, the present model
can be modified by considering the pipes as boreholes. However, the small
pipe size may result in convergence problem if the simulation time step is not
reduced. To compromise between stability and computation time, the present
methodology used to calculate the fully implicit scheme has to be modified.

134
Appendices

A. Transformation of Thermal Interference Coefficients Between


Tubes Inside a Borehole

According to Hellstrom (1991), for a single borehole containing nt tubes lying


concentrically in a circle of radius di, the heat transfer under quasi-steady state
condition with fluid flowing through the tubes can be expressed as
N
T fl ,u Tb = Ruv qt ,v (A.1)
v =1


where Ruv can be evaluated from the formulae quoted from Hellstrom (1991) as

rb 2

Ruu =
1
ln + ln rb + R p (A.2)
2k b r 2 di 2
rpo b

buv
2 2
1 buv rb2 di 2
Ruv = ln + ln + (A.3)
2k b rb r2
b
r

b

kb k g
= (A.4)
kb + k g

buv is the centre-to-centre distance between tube u and v, and Rp is the thermal
resistance between the fluid and the grouting. Rp comprises the convective thermal
resistance between the fluid and the tube inner surface, and the conductive thermal
resistance through the tube wall, which can be calculated as
rpo
ln
r
Rp = +
pi 1
(A.5)
2k p 2hc , f rpi

where
kf
hc , f = 0.023 Re 0f.8 Pr f0.3 for cooling (A.6)
2rpi

kf
hc , f = 0.023 Re 0f.8 Pr f0.4 for heating (A.7)
2rpi

135
according to Dittus and Boelter (1930). To simplify the formulation, a mean value of
0.35 is used for the exponent of Prandtl number in the analysis. The same approach
was also adopted by Yavuzturk (1999). The fluid Reynolds and Prandtl numbers
are determined from Holman (1990) as
2m f
Re f = (A.8)
f rpi
cf f
Pr f = (A.9)
kf

Eq. (A1) is re-written in matrix form as

[T fl ,u ]
Tb = Ruv [qt ,v ]


1

[
Ruv T fl ,v Tb = [qt ,u ]

]
1

By denoting Ruv = Ruv

[q ] = R [T ]

t ,u uv fl , v Tb (A.10)

Eq. (A.10) and Eq. (3.3) represent exactly the same thing. By comparing the
coefficients of T fl ,v (v u ) and Tb from the right hand sides of both equations,
1
Ruv =
for u v (A.11)
Ruv
1
Ruu = (A.12)
nt
R
v =1
uv

For the case with no fluid flow, Rp will contain only the thermal resistance from the
pipe wall. Hence,
rpo
ln
r
R p ,nf pi (A.13)
2k p

The derivation of Rnf ,uv is similar to Ruv except by replacing Rp with Rp,nf in Eq.

(A.2).

136
B. Algorithm for Calculating Condenser Coil Performance

Step 1 Setting of initial parameter values

Calculate initial estimation for cond , sat based on the following equation

mr h f g , cond
cond , sat = Minimum ,1 e NTU dh (B.1)
CL , dhi (Trs , cond TL , dhi )
, sat using the Eqs. (3.48&49).
Calculate initial guess for xcond

Calculate initial guess for TL ,dh 2 using Eq. (3.50), assuming TL ,dh1 = TL ,dhi .

Calculate initial guess for cond , sh based on Eq. (3.46).

, sh based on Eqs. (3.44&45).


Calculate initial guess for xcond

, sh + xcond
If ( xcond , sat ) > 1 , no sub-cooling occurs

, sc = 0 and xcond
xcond , sat = 1 xcond
, sh

Else
, sc = 1 x cond
xcond , sh xcond
, sat .

Step 2 Calculation of new parameter values

Calculate new cond , sh based on Eq. (3.44).

Calculate new cond , sat based on Eq. (3.48).

Determine new TL ,dh 2 based on Eq. (3.46).

Determine new TL ,ri using Eq. (3.47).

Calculate new TL ,dh1 using Eq. (3.50).

Calculate new hr ,co using Eq. (3.51).

If sub-cooling occurs,
Calculate new cond , sc based on Eqs. (3.52&53).

Calculate another TL ,dh1 (TL ,dh1' ) using Eq. (3.55).

Calculate Tr ,co using Eq. (3.54).

Else

137
set TL ,dh1 = TL ,dhi .

set Tr ,co = Trs ,cond .

Step 3 Check of convergence

If no sub-cooling occurs,
If difference between TL ,dh1 and TL ,dhi is less than preset limit,

Proceed to step 4.
Else
, sh based on the difference between TL ,dh1 and
Revise xcond

TL ,dhi .

, sat .
Calculate new xcond

Go back to step 2.
Else
If difference between TL ,dh1 and TL ,dh1' is less than preset limit and

difference between hr ,co and h f ,cond is less than preset limit,

Proceed to step 4.
Else
, sh based on the difference between TL ,dh1 and
Revise xcond

TL ,dh1' .

, sat based on the difference between hr ,co and


Revise xcond

h f ,cond .

, sc based on Eq. (3.43).


Calculate new xcond

Go back to step 2.

Step 4 Determination of final results

Calculate leaving refrigerant enthalpy and quality.


Calculate refrigerant charge based on Eq. (3.56).

138
C. Determination of Mean Refrigerant Density along Saturated
Region in Counter-flow Condenser/Evaporator

Consider a counter-flow heat exchanger coil with heat transfer between a


refrigerant and a fluid (air or any liquid). Assume that the refrigerant lies in the
saturated region (constant temperature) throughout the entire coil and the heat
capacity of the fluid remains constant. At some characteristic length l (0=fluid inlet,
1=refrigerant inlet),
C f dT f = UA(Tr T f )dl

UA
dT f = (Tr T f )dl
Cf

UA
By setting NTU = and re-arranging,
Cf

dT f
= NTUdl
Tr T f

Integrating and applying the boundary condition that at l = 0 , T f = T f ,in ,

T T f ,in
ln r = NTU l
T T
r f
T f = Tr (Tr T f ,in )e NTU l (C.1)

By equating the energy change between refrigerant and fluid,


C f dT f = mr dhr = mr h f g d

mr h f g
dT f = d (C.2)
Cf

where is the quality of the refrigerant. Combining Eqs. (C.1) and (C.2),
mr h f g
d = NTU (Tr T f ,in )e NTU l dl
Cf

NTU (Tr T f ,in )C f


d = e NTU l dl
mr h f g

Integrating and setting = in at l = 1 and re-arranging,

= in + (e NTU e NTU l ) (C.3)

139
(Tr T f ,in )C f
where = (C.4)
mr h f g

The density of refrigerant at the saturated region is defined as


1 g f
r = = (C.5)
1 g + f g
+
g f

where f g = f g .

Substituting Eq. (C.3) into Eq. (C.5),


g f
r =
g + f g [ in + (e NTU e NTU l )
g f
= (C.6)
f g e NTU l

where = g + f g ( in + e NTU ) (C.7)

The mean refrigerant density can be determined by integrating Eq. (C.6) by dl from
0 to 1. Thus,
1
r = r dl
0

1
g f
= dl
0
f g e NTU L
1
l 1
= g f + ln( f g e NTU l )
NTU 0

1 1 f g e NTU
= g f + ln (C.8)

NTU f g

140
D. Algorithm for Calculating Evaporator Coil Performance

Step 1 Setting of initial parameter values

, sat = 1 .
Calculate hr , suc based on Eq. (3.63), assuming xevap

If (hr , suc < hg , evap ) , no superheating occurs

Set Tr , suc = Trs ,evap .

Calculate leaving refrigerant quality.


Proceed to step 4.
Else
, sh based on the following equation
Calculate initial estimation for xevap

hr , suc hg , evap
, sh =
xevap (D.1)
hr , suc hr , ei

Continue to step 2.

Step 2 - Calculation of new parameter values

, sat based on Eq. (3.59).


Calculate xevap

Calculate cond , sat based on Eqs. (3.60&61).

Calculate cond , sh based on Eqs. (3.64&65).

Calculate TL ,dc1 using Eqs. (3.66&67).

Calculate Tr , suc using Eq. (3.66).

Calculate TL ,di using Eq. (3.62).

Calculate hr ,dc1 using Eq. (3.63) by replacing hr , suc with hr ,dc1 .

Step 3 - Check of convergence

If difference between hr ,dc1 and hg , evep less than preset limit

Proceed to Step 4.
Else

141
, sh based on the difference between hr ,dc1 and hg , evap .
Revise xevap

Go back to step 2.

Step 4 - Determination of final results

Calculate leaving refrigerant enthalpy.


Calculate refrigerant charge based on Eq. (3.68).

142
E. Algorithm for Calculating Air Cooler Performance Based on
Approximate Iterative Method

E.1 Algorithm for partial coil

If entering refrigerant at saturated region,


Follow procedures in Appendix D to calculate leaving coil conditions
and refrigerant charge.
Else
Calculate leaving coil conditions based on Eqs. (3.64~67).

E.2 Algorithm for entire air cooler

Step 1 Setting of initial parameter values

Calculate Tsa using algorithm E.1 based on dry conditions for the entire air
cooler.
If (Tsa > Tadp , aci ) , no condensation occurs

sa = a , dci .
Proceed to step 4.
Else
, wet from the following equation
Calculate initial estimation of x ac

Tsa + 10
xac , wet = 1 (E.1)
Tadp , aci + 10

Continue to step 2.

Step 2 - Calculation of new parameter values

Calculate Tr ,ac1 using algorithm E.1 based on wet conditions for the wet

partial coil.
Calculate new Tadp ,dci using algorithm E.1 based on dry conditions for the dry

partial coil.

143
Step 3 - Check of convergence

If difference between new Tadp ,dci and true Tadp ,dci less than preset limit

Proceed to Step 4.
Else
Revise x ac , wet based on the difference between new Tadp ,dci and true

Tadp ,dci .

Go back to step 2.

Step 4 - Determination of final results

Calculate leaving air properties.


Calculate refrigerant charge.
Calculate cooling capacity of entire coil.

144
F. Modified Numerical Integration Method Used for FLSM

From Eq. (2.33), the determination of the ground temperature rise requires
the evaluation of a definite integral. The integrand reaches a maximum when s z ,
and drops rapidly when s differs from z like a needle hill, as shown in Figure F.1.
The integration of such kind of function can lead to a wrong result if non-uniform
step is used as is found when trying to determine Eq. (2.25) without specifying a
fixed integration step and the integration starts at the lower integration limit. A small
integration step ensures accuracy but at the expense of long computation time.
Hence, a compromise must be found.
Integrand of Eq. (2.33)

s'
0 d z H
Figure F.1 Variation of integrand of Eq. (2.33) with s

In the modified method, numerical integration using Simpsons Rule starts at


s = z if d < z < H , and spreads away with an increasing time step size by a
multiplier 1.4. The minimum integration time step used is 0.01. If z < d ,
integration will start at s = d to s = H . When z > H , integration will start at
s = H to s = d . By applying such method, the computation time is sharply reduced
as compared with that using a fixed integration time step of 0.01.

145
G. Modified Numerical Integration Method Used for LSMGA

With LSMGA, the average borehole temperature rise is given by Eqs.


(2.37&38) while borehole thermal interference determined by Eq. (2.35). Both
require the estimation of improper integrals where the integrands are undefined but
bounded and approaching zero at the integration lower limits with profiles similar
than that shown in Figure F.1. Hence, similar algorithm as presented in Appendix F
can be applied.

By denoting the integrand of Eq. (2.35) as and that of Eq. (2.37) as ,


( x V ' ) 2 + y 2 ( x V ' ) 2 + y 2
4ag (2V ' 2 2V ' x) 4ag (( x V ' ) 2 + y 2 )
e
4 a g
e 4 a g

16 a
2 2

= g
2
( x V ' ) 2 + y 2

4 a g V ' 2 2 + 4ag ( x 2 + y 2 )
= e = t t' (G.1)
4a g 3
1 R12 1 R12

16 1 R12 16
2e
2 2e 1 R12


16 16 1 R2 1
= =e
3 1 2 (G.2)
4 2 2 32 2

The value of * and * where and reach the maximum can be found by setting
Eqs. (G.1&2) to zero. Hence,
V '2 *2 + 4a g * ( x 2 + y 2 ) = 0

2a g + 4a g2 + V ' 2 ( x 2 + y 2 )
=
*
for V' 0 (G.3)
V '2
R12 *2 + 16 * 16 = 0

8(1 + 1 + R12 / 4 )
=
*
for R1 0 (G.4)
R12
The minimum integration time step and lower integration limit for Eq. (2.35) are
both set to 1, and those for Eq. (2.37) being 0.00001. The integration time step
multiplier is 1.01 for both equations.

146
H. Borehole Thermal Interference Based on FLSMGA

The ground temperature rise due to an instantaneous moving point source of


strength qb dt ' dz ' at time t ' with distance s ' along the z-axis was given by Carslaw
and Jaeger (1959) as
[( x V ' t ' ) 2 + y 2 + ( z s ' ) 2 ]
qb dt 'ds ' 4ag t '
inst , psga = e (H.1)
8 g cg (a g t ' )3 / 2

The temperature rise due to an instantaneous finite line source with insulated depth d
and effective length H below ground level is given by integrating Eq. (H.1) by ds '
from d to H. To maintain no temperature change at ground level, a negative strength
mirror finite line source is to be added above the ground level. The overall
temperature rise due to an instantaneous moving finite line source becomes
d + H [( x V t ) + y ' + ( z s ) ] [( z V t ) + y + ( z + s ) ]
' ' 2 2 ' 2 ' ' 2 2 ' 2
d +H
qb dt ' '

8 g cg (a g t ' )3 / 2 d d
4ag t '
inst , flsmga =
4ag t '
e dz e dz

[( x V ' t ' ) 2 + y 2 ] d + H ( z s ') ( z + s )


' 2 ' 2
d +H
qb dt ' '
d
'
4ag t '
=
4ag t 4a g t '
e e dz e dz (H.2)
8 g cg (ag t ' )3 / 2 d
x
2
e
2
Recalling the definition of error function that erf ( x) = d , Eq. (H.2) can
0

be rewritten as
[( x V ' t ' ) 2 + y 2 ]

qb dt ' 4ag t ' zd erf z d H
inst , flsmga = e erf 4a t '
8k g t ' 4ag t
'
g


z+d +H + erf z + d
erf (H.3)
4a g t '

'
4a g t
The temperature rise due to a continuous moving finite line source of constant load
qb at time t is then given as
{[ x V ' ( t t ' )] 2 + y 2 }
4 a g (t t ' )
zd erf z d H
t
q e
flsmga = b erf
8k g 0 t t' ' 4a (t t ' )
4a g (t t ) g

147

z+d +H z+d dt '
erf + erf
(H.4)
4a g (t t ' )
4 a (t t '
)
g

By putting = t t ' , Eq. (H.4) becomes


[( x V ' ) 2 + y 2 ]
4 a g z d
erf z d H
t
qb e
erf
8k g 0
flsmga =
4a g 4a g

z+d +H
erf + erf z + d d (H.5)
4ag 4ag

The thermal interference due to adjacent borehole distant x1 and y1 is


calculated by averaging Eq. (H.5) along the depth from d to H as
[( x1 V ) + y12 ]
' 2

d +H
4 a g zd
qb te
erf erf z d H
8k g H d 0
flsmga ,inf =
4ag 4ag

z+d +H z + d
erf + erf d dz
4a g
(H.6)
4a g

or in terms of g function as
[( x1 V ) + y12 ]
' 2

1 d + H t e
4 a g zd
erf erf z d H
4 H d 0
g flsmga ,inf =
4ag 4ag

z+d +H
erf + erf z + d d dz (H.7)
4a g 4a g

Eqs. (H.6&7) require the evaluation of a double integral, which is very time
consuming. To simplify, a weighted average is used as
[( x1 V ) + y12 ]
' 2

nz bore 1
t e 4 a g z d
qb
erf i erf zi d H
flsmga ,inf =
8k g H

4ag 4a g
i =1 0

148
z +d +H
erf i + erf zi + d d dz (H.8)
4a g 4a g i

[( x1 V ) + y12 ]
' 2

nz bore 1
t e 4 a g z d
1
erf i erf zi d H
flsmga ,inf =
4H

4ag 4ag
i =1 0

z +d +H
erf i + erf zi + d d dz (H.9)
4a g 4a g i

where z i are the mid-depths of vertical segments discretised in the same way as the
present numerical model. The trends of the integrands are similar to that for Eq.
(2.35). Unfortunately, an explicit expression for * cannot be made. Thus, the
integration will start at the lower integration limit (set as 10) with fixed time step (set
as 1) until the integrand starts to decrease. Then a multiplier of 1.01 for the
integration time step will be used for all subsequent integration steps.

149
I. Calibration Curves for Measuring Instruments Used in Laboratory
Tests

Calibration curve of borehole temperature logger

40
38
Set temperature (degC)

36
34
32
30
28
26
24
22
20
21 26 31 36 41

Logger reading (degC)

Figure I.1 Calibration curve for borehole temperature logger

Calculation curves for temperature log buttons


Circulating water Room

30
Set temperature (degC)

28

26

24

22

20
19 21 23 25 27 29 31

Logger reading (degC)

Figure I.2 Calibration curves for temperature log buttons

150
Calculation curves for thermocouples
T _heater T _ambient

50

Set temperature (degC)


45

40

35

30
28 33 38 43 48

Thermocouple reading (degC)

Figure I.3 Calibration of thermocouples for soil thermal conductivity measurement

Calibration curve for voltmeter

50
Power analyser reading (V)

45
40
35
30
25
20
15
10
5
0 10 20 30 40 50 60

Voltmeter reading (V)

Figure I.4 Calibration curve for voltmeter

Calibration curve for ammeter

0.5
Power analyser reading (A)

0.45
0.4
0.35
0.3
0.25
0.2
0.15
0.1
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45

Ammeter reading (A)

Figure I.5 Calibration curve for ammeter

151
J. Detailed Performance Curves at Different Groundwater Directions
and Load Profiles for Various Borefields

Leaving fluid temperature rise Leaving fluid temperature rise


at DV=1E-7 at DV=2E-7
18 15
17.5 14.8
17 14.6

Tbf,f,out - To (K)
Tbf,f,out - To (K)

16.5 14.4
1 year 1 year
16 14.2
2 years 2 years
15.5 14
5 years 5 years
15 13.8
14.5 10 years 13.6 10 years
14 13.4
13.5 13.2
13 13
0 10 20 30 40 50 0 10 20 30 40 50

Groundwater direction (deg) Groundwater direction (deg)

Figure J.1 Performance of 2x2 borefield with Figure J.2 Performance of 2x2 borefield with
groundwater velocity 10-7m/s at different groundwater velocity 2x10-7m/s at different
groundwater directions under load profile 1 groundwater directions under load profile 1

Leaving fluid temperature rise Leaving fluid temperature rise


at DV=5E-7 after two years
11 9
10.9 8.5
10.8 8
Tbf,f,out - To (K)
Tbf,f,out - To (K)

10.7 7.5
1 year DV=1E-6
10.6 7
2 years DV=2E-6
10.5 6.5
5 years DV=5E-6
10.4 6
10.3 10 years DV=1E-5
5.5
10.2 5
10.1 4.5
10 4
0 10 20 30 40 50 0 10 20 30 40 50

Groundwater direction (deg) Groundwater direction (deg)

Figure J.3 Performance of 2x2 borefield with Figure J.4 Performance of 2x2 borefield with high
groundwater velocity 5x10-7m/s at different groundwater velocity after 2 years at different
groundwater directions under load profile 1 groundwater directions under load profile 1

Leaving fluid temperature rise Leaving fluid temperature rise


at DV=1E-7 at DV=2E-7
11 9.5
9.4
10.5 9.3
Tbf,f,out - To (K)

Tbf,f,out - To (K)

9.2
1 year 1 year
10 9.1
2 years 2 years
9
5 years 5 years
9.5 8.9
10 years 8.8 10 years
9 8.7
8.6
8.5 8.5
0 10 20 30 40 50 0 10 20 30 40 50

Groundwater direction (deg) Groundwater direction (deg)

Figure J.5 Performance of 2x2 borefield with Figure J.6 Performance of 2x2 borefield with
groundwater velocity 10-7m/s at different groundwater velocity 2x10-7m/s at different
groundwater directions under load profile 2 groundwater directions under load profile 2

152
Leaving fluid temperature rise Leaving fluid temperature rise
at DV=5E-7 after two years
8 7
7.9
6.5
7.8
Tbf,f,out - To (K)

Tbf,f,out - To (K)
7.7
1 year 6 DV=1E-6
7.6
2 years DV=2E-6
7.5 5.5
5 years DV=5E-6
7.4
7.3 10 years 5 DV=1E-5
7.2
4.5
7.1
7 4
0 10 20 30 40 50 0 10 20 30 40 50

Groundwater direction (deg) Groundwater direction (deg)

Figure J.7 Performance of 2x2 borefield with Figure J.8 Performance of 2x2 borefield with high
groundwater velocity 5x10-7m/s at different groundwater velocity after 2 years at different
groundwater directions under load profile 2 groundwater directions under load profile 2

Leaving fluid temperature rise Leaving fluid temperature rise


at DV=1E-7 at DV=2E-7
9 8
8.8 7.95
8.6 7.9

Tbf,f,out - To (K)
Tbf,f,out - To (K)

8.4 7.85
1 year 1 year
8.2 7.8
2 years 2 years
8 7.75
5 years 5 years
7.8 7.7
7.6 10 years 7.65 10 years
7.4 7.6
7.2 7.55
7 7.5
0 10 20 30 40 50 0 10 20 30 40 50

Groundwater direction (deg) Groundwater direction (deg)

Figure J.9 Performance of 2x2 borefield with Figure J.10 Performance of 2x2 borefield with
groundwater velocity 10-7m/s at different groundwater velocity 2x10-7m/s at different
groundwater directions under load profile 3 groundwater directions under load profile 3

Leaving fluid temperature rise Leaving fluid temperature rise


at DV=5E-7 after two years
7.5 7
7.45
6.5
7.4
Tbf,f,out - To (K)
Tbf,f,out - To (K)

7.35 6
1 year DV=1E-6
7.3
2 years DV=2E-6
7.25 5.5
5 years DV=5E-6
7.2
10 years 5 DV=1E-5
7.15
7.1
4.5
7.05
7 4
0 10 20 30 40 50 0 10 20 30 40 50

Groundwater direction (deg) Groundwater direction (deg)

Figure J.11 Performance of 2x2 borefield with Figure J.12 Performance of 2x2 borefield with
groundwater velocity 5x10-7m/s at different high groundwater velocity after 2 years at
groundwater directions under load profile 3 different groundwater directions under load
profile 3

153
Leaving fluid temperature rise Leaving fluid temperature rise
at DV=1E-7 at DV=2E-7
8.4 8
7.9
8.2
7.8

Tbf,f,out - To (K)

Tbf,f,out - To (K)
8 7.7
1 year 1 year
7.8 7.6
2 years 2 years
7.5
7.6 5 years 5 years
7.4
10 years 10 years
7.4 7.3
7.2
7.2
7.1
7 7
0 10 20 30 40 50 0 10 20 30 40 50
Groundwater direction (deg) Groundwater direction (deg)

Figure J.13 Performance of 2x2 borefield with Figure J.14 Performance of 2x2 borefield with
groundwater velocity 10-7m/s at different groundwater velocity 2x10-7m/s at different
groundwater directions under load profile 4 groundwater directions under load profile 4

Leaving fluid temperature rise Leaving fluid temperature rise


at DV=5E-7 after two years
7.5 7
7.4
7.3 6.5
Tbf,f,out - To (K)

Tbf,f,out - To (K)
7.2
1 year 6 DV=1E-6
7.1
2 years DV=2E-6
7 5.5
5 years DV=5E-6
6.9
6.8 10 years 5 DV=1E-5
6.7
4.5
6.6
6.5 4
0 10 20 30 40 50 0 10 20 30 40 50
Groundwater direction (deg) Groundwater direction (deg)

Figure J.15 Performance of 2x2 borefield with Figure J.16 Performance of 2x2 borefield with
groundwater velocity 5x10-7m/s at different high groundwater velocity after 2 years at different
groundwater directions under load profile 4 groundwater directions under load profile 4

Leaving fluid temperature rise Leaving fluid temperature rise


at DV=1E-7 at DV=2E-7
25 19
24 18.5
23
18
Tbf,f,out - To (K)

Tbf,f,out - To (K)

22 1 year 1 year
17.5
21 2 years 2 years
17
20 5 years 5 years
19 16.5
10 years 10 years
18 16
17 15.5
16 15
0 10 20 30 40 50 0 10 20 30 40 50
Groundwater direction (deg) Groundwater direction (deg)

Figure J.17 Performance of 3x3 borefield with Figure J.18 Performance of 3x3 borefield with
groundwater velocity 10-7m/s at different groundwater velocity 2x10-7m/s at different
groundwater directions under load profile 1 groundwater directions under load profile 1

154
Leaving fluid temperature rise Leaving fluid temperature rise
at DV=5E-7 after two years
13 10
12.9
12.8 9

Tbf,f,out - To (K)

Tbf,f,out - To (K)
12.7
1 year 8 DV=1E-6
12.6
2 years DV=2E-6
12.5 7
5 years DV=5E-6
12.4
12.3 10 years 6 DV=1E-5
12.2
5
12.1
12 4
0 10 20 30 40 50 0 10 20 30 40 50

Groundwater direction (deg) Groundwater direction (deg)

Figure J.19 Performance of 3x3 borefield with Figure J.20 Performance of 3x3 borefield with
groundwater velocity 5x10-7m/s at different high groundwater velocity after 2 years at
groundwater directions under load profile 1 different groundwater directions under load
profile 1

Leaving fluid temperature rise Leaving fluid temperature rise


at DV=1E-7 at DV=2E-7
14 12
13.5
11.5
13
Tbf,f,out - To (K)

1 year Tbf,f,out - To (K) 11 1 year


12.5
2 years 2 years
12 10.5
5 years 5 years
11.5
10 years 10 10 years
11
9.5
10.5
10 9
0 10 20 30 40 50 0 10 20 30 40 50

Groundwater direction (deg) Groundwater direction (deg)

Figure J.21 Performance of 3x3 borefield with Figure J.22 Performance of 3x3 borefield with
groundwater velocity 10-7m/s at different groundwater velocity 2x10-7m/s at different
groundwater directions under load profile 2 groundwater directions under load profile 2

Leaving fluid temperature rise Leaving fluid temperature rise


at DV=5E-7 after two years
8.5 8
8.45 7.5
8.4
7
Tbf,f,out - To (K)
Tbf,f,out - To (K)

8.35
1 year DV=1E-6
8.3 6.5
2 years DV=2E-6
8.25 6
5 years DV=5E-6
8.2 5.5
8.15 10 years DV=1E-5
5
8.1
8.05 4.5

8 4
0 10 20 30 40 50 0 10 20 30 40 50

Groundwater direction (deg) Groundwater direction (deg)

Figure J.23 Performance of 3x3 borefield with Figure J.24 Performance of 3x3 borefield with
groundwater velocity 5x10-7m/s at different high groundwater velocity after 2 years at
groundwater directions under load profile 2 different groundwater directions under load
profile 2

155
Leaving fluid temperature rise Leaving fluid temperature rise
at DV=1E-7 at DV=2E-7
10.5 9
8.9
10 8.8

Tbf,f,out - To (K)
Tbf,f,out - To (K)
8.7
1 year 1 year
9.5 8.6
2 years 2 years
8.5
5 years 5 years
9 8.4
10 years 10 years
8.3
8.5 8.2
8.1
8 8
0 10 20 30 40 50 0 10 20 30 40 50

Groundwater direction (deg) Groundwater direction (deg)

Figure J.25 Performance of 3x3 borefield with Figure J.26 Performance of 3x3 borefield with
groundwater velocity 10-7m/s at different groundwater velocity 2x10-7m/s at different
groundwater directions under load profile 3 groundwater directions under load profile 3

Leaving fluid temperature rise Leaving fluid temperature rise


at DV=5E-7 after two years
8 7
7.9
7.8 6.5
Tbf,f,out - To (K)

Tbf,f,out - To (K)
7.7
1 year 6 DV=1E-6
7.6
2 years DV=2E-6
7.5 5.5
5 years DV=5E-6
7.4
10 years 5 DV=1E-5
7.3
7.2
4.5
7.1
7 4
0 10 20 30 40 50 0 10 20 30 40 50

Groundwater direction (deg) Groundwater direction (deg)

Figure J.27 Performance of 3x3 borefield with Figure J.28 Performance of 3x3 borefield with
groundwater velocity 5x10-7m/s at different high groundwater velocity after 2 years at
groundwater directions under load profile 3 different groundwater directions under load
profile 3

Leaving fluid temperature rise Leaving fluid temperature rise


at DV=1E-7 at DV=2E-7
9.5 8.5
9.3 8.4
9.1 8.3
Tbf,f,out - To (K)

Tbf,f,out - To (K)

8.9 8.2
1 year 1 year
8.7 8.1
2 years 2 years
8.5 8
5 years 5 years
8.3 7.9
10 years 10 years
8.1 7.8
7.9 7.7
7.7 7.6
7.5 7.5
0 10 20 30 40 50 0 10 20 30 40 50
Groundwater direction (deg) Groundwater direction (deg)

Figure J.29 Performance of 3x3 borefield with Figure J.30 Performance of 3x3 borefield with
groundwater velocity 10-7m/s at different groundwater velocity 2x10-7m/s at different
groundwater directions under load profile 4 groundwater directions under load profile 4

156
Leaving fluid temperature rise Leaving fluid temperature rise
at DV=5E-7 after two years
7.6 7

7.5 6.5

Tbf,f,out - To (K)

Tbf,f,out - To (K)
7.4 1 year 6 DV=1E-6
2 years DV=2E-6
7.3 5.5
5 years DV=5E-6
7.2 10 years 5 DV=1E-5

7.1 4.5

7 4
0 10 20 30 40 50 0 10 20 30 40 50

Groundwater direction (deg) Groundwater direction (deg)

Figure J.31 Performance of 3x3 borefield with Figure J.32 Performance of 3x3 borefield with
groundwater velocity 5x10-7m/s at different high groundwater velocity after 2 years at
groundwater directions under load profile 4 different groundwater directions under load
profile 4

Leaving fluid temperature rise Leaving fluid temperature rise


at DV=1E-7 at DV=2E-7
14 12.2

13.5 12

Tbf,f,out - To (K)
Tbf,fout - To (K)

13 1 year 11.8 1 year


2 years 2 years
12.5 11.6
5 years 5 years
12 10 years 11.4 10 years

11.5 11.2

11 11
0 20 40 60 80 100 0 20 40 60 80 100

Groundwater direction (deg) Groundwater direction (deg)

Figure J.33 Performance of 1x2 borefield with Figure J.34 Performance of 1x2 borefield with
groundwater velocity 10-7m/s at different groundwater velocity 2x10-7m/s at different
groundwater directions under load profile 1 groundwater directions under load profile 1

Leaving fluid temperature rise Leaving fluid temperature rise


at DV=5E-7 after two years
10 9
8.5
9.9
8
Tbf,f,out - To (K)
Tbf,f,out - To (K)

9.8 1 year 7.5


DV=1E-6
7
2 years DV=2E-6
9.7 6.5
5 years DV=5E-6
6
9.6 10 years 5.5 DV=1E-5
5
9.5
4.5
9.4 4
0 20 40 60 80 100 0 20 40 60 80 100

Groundwater direction (deg) Groundwater direction (deg)

Figure J.35 Performance of 1x2 borefield with Figure J.36 Performance of 1x2 borefield with
groundwater velocity 5x10-7m/s at different high groundwater velocity after 2 years at
groundwater directions under load profile 1 different groundwater directions under load
profile 1

157
Leaving fluid temperature rise Leaving fluid temperature rise
at DV=1E-7 at DV=2E-7
9 8.5
8.9 8.4
8.8 8.3

Tbf,f,out - To (K)

Tbf,f,out - To (K)
8.7 8.2
1 year 1 year
8.6 8.1
2 years 2 years
8.5 8
5 years 5 years
8.4 7.9
8.3 10 years 7.8 10 years
8.2 7.7
8.1 7.6
8 7.5
0 20 40 60 80 100 0 20 40 60 80 100

Groundwater direction (deg) Groundwater direction (deg)

Figure J.37 Performance of 1x2 borefield with Figure J.38 Performance of 1x2 borefield with
groundwater velocity 10-7m/s at different groundwater velocity 2x10-7m/s at different
groundwater directions under load profile 2 groundwater directions under load profile 2

Leaving fluid temperature rise Leaving fluid temperature rise


at DV=5E-7 after two years
7.5 7
7.4
7.3 6.5
Tbf,f,out - To (K)

Tbf,f,out - To (K)
7.2
1 year 6 DV=1E-6
7.1
2 years DV=2E-6
7 5.5
5 years DV=5E-6
6.9
6.8 10 years 5 DV=1E-5
6.7
4.5
6.6
6.5 4
0 20 40 60 80 100 0 20 40 60 80 100
Groundwater direction (deg) Groundwater direction (deg)

Figure J.39 Performance of 1x2 borefield with Figure J.40 Performance of 1x2 borefield with
groundwater velocity 5x10-7m/s at different high groundwater velocity after 2 years at
groundwater directions under load profile 2 different groundwater directions under load
profile 2

Leaving fluid temperature rise Leaving fluid temperature rise


at DV=1E-7 at DV=2E-7
8 7.5
7.9 7.45
7.8 7.4
Tbf,f,out - To (K)

Tbf,f,out - To (K)

7.7 7.35
1 year 1 year
7.6 7.3
2 years 2 years
7.5 7.25
5 years 5 years
7.4 7.2
7.3 10 years 10 years
7.15
7.2 7.1
7.1 7.05
7 7
0 20 40 60 80 100 0 20 40 60 80 100
Groundwater direction (deg) Groundwater direction (deg)

Figure J.41 Performance of 1x2 borefield with Figure J.42 Performance of 1x2 borefield with
groundwater velocity 10-7m/s at different groundwater velocity 2x10-7m/s at different
groundwater directions under load profile 3 groundwater directions under load profile 3

158
Leaving fluid temperature rise Leaving fluid temperature rise
at DV=5E-7 after two years
7 7
6.95
6.9 6.5

Tbf,f,out - To (K)

Tbf,f,out - To (K)
6.85
1 year 6 DV=1E-6
6.8
2 years DV=2E-6
6.75 5.5
5 years DV=5E-6
6.7
6.65 10 years 5 DV=1E-5
6.6
4.5
6.55
6.5 4
0 20 40 60 80 100 0 20 40 60 80 100
Groundwater direction (deg) Groundwater direction (deg)

Figure J.43 Performance of 1x2 borefield with Figure J.44 Performance of 1x2 borefield with
groundwater velocity 5x10-7m/s at different high groundwater velocity after 2 years at different
groundwater directions under load profile 3 groundwater directions under load profile 3

Leaving fluid temperature rise Leaving fluid temperature rise


at DV=1E-7 at DV=2E-7
7.6 7.3

7.5 7.25
Tbf,f,out - To (K)

Tbf,f,out - To (K)
7.4 1 year 7.2 1 year
2 years 2 years
7.3 7.15
5 years 5 years
7.2 10 years 7.1 10 years

7.1 7.05

7 7
0 20 40 60 80 100 0 20 40 60 80 100
Groundwater direction (deg) Groundwater direction (deg)

Figure J.45 Performance of 1x2 borefield with Figure J.46 Performance of 1x2 borefield with
groundwater velocity 10-7m/s at different groundwater velocity 2x10-7m/s at different
groundwater directions under load profile 4 groundwater directions under load profile 4

Leaving fluid temperature rise Leaving fluid temperature rise


at DV=5E-7 after two years
7 7
6.95
6.9 6.5
Tbf,f,out - To (K)

Tbf,f,out - To (K)

6.85
1 year 6 DV=1E-6
6.8
2 years DV=2E-6
6.75 5.5
5 years DV=5E-6
6.7
6.65 10 years 5 DV=1E-5
6.6
4.5
6.55
6.5 4
0 20 40 60 80 100 0 20 40 60 80 100
Groundwater direction (deg) Groundwater direction (deg)

Figure J.47 Performance of 1x2 borefield with Figure J.48 Performance of 1x2 borefield with
groundwater velocity 5x10-7m/s at different high groundwater velocity after 2 years at
groundwater directions under load profile 4 different groundwater directions under load
profile 4

159
Leaving fluid temperature rise Leaving fluid temperature rise
at DV=1E-7 at DV=2E-7
16 14
15.5
13.5
15

Tbf,f,out - To (K)
Tbf,f,out - To (K)
1 year 13 1 year
14.5
2 years 2 years
14 12.5
5 years 5 years
13.5
10 years 12 10 years
13
11.5
12.5
12 11
0 20 40 60 80 100 0 20 40 60 80 100

Groundwater direction (deg) Groundwater direction (deg)

Figure J.49 Performance of 1x3 borefield with Figure J.50 Performance of 1x3 borefield with
groundwater velocity 10-7m/s at different groundwater velocity 2x10-7m/s at different
groundwater directions under load profile 1 groundwater directions under load profile 1

Leaving fluid temperature rise Leaving fluid temperature rise


at DV=5E-7 after two years
11 10
10.8
10.6 9
Tbf,f,out - To (K)

Tbf,f,out - To (K)
10.4
1 year 8 DV=1E-6
10.2
2 years DV=2E-6
10 7
5 years DV=5E-6
9.8
9.6 10 years 6 DV=1E-5
9.4
5
9.2
9 4
0 20 40 60 80 100 0 20 40 60 80 100
Groundwater direction (deg) Groundwater direction (deg)

Figure J.51 Performance of 1x3 borefield with Figure J.52 Performance of 1x3 borefield with
groundwater velocity 5x10-7m/s at different high groundwater velocity after 2 years at different
groundwater directions under load profile 1 groundwater directions under load profile 1

Leaving fluid temperature rise Leaving fluid temperature rise


at DV=1E-7 at DV=2E-7
10 9
9.8 8.9
9.6 8.8
Tbf,f,out - To (K)
Tbf,f,out - To (K)

9.4 8.7
1 year 1 year
9.2 8.6
2 years 2 years
9 8.5
5 years 5 years
8.8 8.4
10 years 8.3 10 years
8.6
8.4 8.2
8.2 8.1
8 8
0 20 40 60 80 100 0 20 40 60 80 100

Groundwater direction (deg) Groundwater direction (deg)

Figure J.53 Performance of 1x3 borefield with Figure J.54 Performance of 1x3 borefield with
groundwater velocity 10-7m/s at different groundwater velocity 2x10-7m/s at different
groundwater directions under load profile 2 groundwater directions under load profile 2

160
Leaving fluid temperature rise Leaving fluid temperature rise
at DV=5E-7 after two years
7.6 7
7.5
7.4 6.5

Tbf,f,out - To (K)

Tbf,f,out - To (K)
7.3
1 year 6 DV=1E-6
7.2
2 years DV=2E-6
7.1 5.5
5 years DV=5E-6
7
6.9 10 years 5 DV=1E-5
6.8
4.5
6.7
6.6 4
0 20 40 60 80 100 0 20 40 60 80 100
Groundwater direction (deg) Groundwater direction (deg)

Figure J.55 Performance of 1x3 borefield with Figure J.56 Performance of 1x3 borefield with
groundwater velocity 5x10-7m/s at different high groundwater velocity after 2 years at
groundwater directions under load profile 2 different groundwater directions under load
profile 2

Leaving fluid temperature rise Leaving fluid temperature rise


at DV=1E-7 at DV=2E-7
8.2 7.8
8.1
7.7
8
Tbf,f,out - To (K)

1 year Tbf,f,out - To (K) 7.6 1 year


7.9
2 years 2 years
7.8 7.5
5 years 5 years
7.7
10 years 7.4 10 years
7.6
7.3
7.5
7.4 7.2
0 20 40 60 80 100 0 20 40 60 80 100

Groundwater direction (deg) Groundwater direction (deg)

Figure J.57 Performance of 1x3 borefield with Figure J.58 Performance of 1x3 borefield with
groundwater velocity 10-7m/s at different groundwater velocity 2x10-7m/s at different
groundwater directions under load profile 3 groundwater directions under load profile 3

Leaving fluid temperature rise Leaving fluid temperature rise


at DV=5E-7 after two years
7.2 7

7.1 6.5
Tbf,f,out - To (K)

Tbf,f,out - To (K)

7 1 year 6 DV=1E-6
2 years DV=2E-6
6.9 5.5
5 years DV=5E-6
6.8 10 years 5 DV=1E-5

6.7 4.5

6.6 4
0 20 40 60 80 100 0 20 40 60 80 100

Groundwater direction (deg) Groundwater direction (deg)

Figure J.59 Performance of 1x3 borefield with Figure J.60 Performance of 1x3 borefield with
groundwater velocity 5x10-7m/s at different high groundwater velocity after 2 years at
groundwater directions under load profile 3 different groundwater directions under load
profile 3

161
Leaving fluid temperature rise Leaving fluid temperature rise
at DV=1E-7 at DV=2E-7
7.8 7.4
7.7
7.35
7.6

Tbf,f,out - To (K)

Tbf,f,out - To (K)
1 year 7.3 1 year
7.5
2 years 2 years
7.4 7.25
5 years 5 years
7.3
10 years 7.2 10 years
7.2
7.1 7.15

7 7.1
0 20 40 60 80 100 0 20 40 60 80 100
Groundwater direction (deg) Groundwater direction (deg)

Figure J.61 Performance of 1x3 borefield with Figure J.62 Performance of 1x3 borefield with
groundwater velocity 10-7m/s at different groundwater velocity 2x10-7m/s at different
groundwater directions under load profile 4 groundwater directions under load profile 4

Leaving fluid temperature rise Leaving fluid temperature rise


at DV=5E-7 after two years
7.1 7
7.05
7
6.5
Tbf,f,out - To (K)

Tbf,f,out - To (K)
6.95
1 year 6 DV=1E-6
6.9
2 years DV=2E-6
6.85 5.5
5 years DV=5E-6
6.8
10 years 5 DV=1E-5
6.75
6.7
4.5
6.65
6.6 4
0 20 40 60 80 100 0 20 40 60 80 100
Groundwater direction (deg) Groundwater direction (deg)

Figure J.63 Performance of 1x3 borefield with Figure J.64 Performance of 1x3 borefield with
groundwater velocity 5x10-7m/s at different high groundwater velocity after 2 years at
groundwater directions under load profile 4 different groundwater directions under load
profile 4

162
K. Parameter Estimation Results with Specified Groundwater Flow
Using LSMGA on Generated TRT Data for 5 Days at Different
Groundwater Velocities

Table K.1 Parameter estimation results with specified groundwater flow using
LSMGA on generated TRT data for 5 days with groundwater velocity 10-8m/s

Specified groundwater Average analysis results based on data generated


velocity used in parameter with groundwater velocity 10-8m/s
estimation (10-8m/s) kg (W/mK) Rb (mK/W) MSD (K2)
0 3.501 0.1253 0.000000
1 3.500 0.1253 0.000000
2 3.501 0.1253 0.000000
5 3.500 0.1253 0.000000
10 3.497 0.1252 0.000000
20 3.491 0.1251 0.000001
50 3.441 0.1243 0.000042
100 3.273 0.1216 0.000696

Table K.2 Parameter estimation results with specified groundwater flow using
LSMGA on generated TRT data for 5 days with groundwater velocity 2x10-8m/s

Specified groundwater Average analysis results based on data generated


velocity used in parameter with groundwater velocity 2x10-8m/s
estimation (10-8m/s) kg (W/mK) Rb (mK/W) MSD (K2)
0 3.499 0.1253 0.000000
1 3.501 0.1253 0.000000
2 3.501 0.1253 0.000000
5 3.500 0.1253 0.000000
10 3.500 0.1253 0.000001
20 3.490 0.1251 0.000001
50 3.441 0.1243 0.000042
100 3.274 0.1217 0.000697

163
Table K.3 Parameter estimation results with specified groundwater flow using
LSMGA on generated TRT data for 5 days with groundwater velocity 5x10-8m/s

Specified groundwater Average analysis results based on data generated


velocity used in parameter with groundwater velocity 5x10-8m/s
estimation (10-8m/s) kg (W/mK) Rb (mK/W) MSD (K2)
0 3.501 0.1253 0.000000
1 3.499 0.1253 0.000000
2 3.502 0.1253 0.000000
5 3.501 0.1253 0.000000
10 3.498 0.1253 0.000000
20 3.492 0.1252 0.000001
50 3.441 0.1243 0.000042
100 3.274 0.1216 0.000696

Table K.4 Parameter estimation results with specified groundwater flow using
LSMGA on generated TRT data for 5 days with groundwater velocity 10-7m/s

Specified groundwater Average analysis results based on data generated


velocity used in parameter with groundwater velocity 10-7m/s
estimation (10-8m/s) kg (W/mK) Rb (mK/W) MSD (K2)
0 3.502 0.1253 0.000000
1 3.502 0.1253 0.000000
2 3.503 0.1253 0.000000
5 3.502 0.1253 0.000000
10 3.496 0.1252 0.000000
20 3.492 0.1252 0.000001
50 3.445 0.1244 0.000039
100 3.276 0.1217 0.000683

164
Table K.5 Parameter estimation results with specified groundwater flow using
LSMGA on generated TRT data for 5 days with groundwater velocity 5x10-7m/s

Specified groundwater Average analysis results based on data generated


velocity used in parameter with groundwater velocity 5x10-7m/s
estimation (10-8m/s) kg (W/mK) Rb (mK/W) MSD (K2)
0 3.559 0.1262 0.000039
1 3.559 0.1262 0.000039
2 3.557 0.1262 0.000039
5 3.558 0.1262 0.000038
10 3.556 0.1262 0.000036
20 3.551 0.1261 0.000028
50 3.502 0.1254 0.000001
100 3.329 0.1226 0.000391

165
L. Parameter Estimation Results with Specified Groundwater Flow
Using LSMGA on Generated TRT Data for 30 Days at Different
Groundwater Velocities

Table L.1 Parameter estimation results with specified groundwater flow using
LSMGA on generated TRT data for 30 days without groundwater flow

Specified groundwater Average analysis results based on data generated


velocity used in parameter with no groundwater
estimation (10-8m/s) kg (W/mK) Rb (mK/W) MSD (K2)
0 3.499 0.1253 0.000000
1 3.501 0.1253 0.000000
2 3.500 0.1253 0.000000
5 3.498 0.1253 0.000000
10 3.486 0.1249 0.000002
20 3.446 0.1238 0.000033
50 3.177 0.1162 0.001425
100 2.557 0.0996 0.027401

Table L.2 Parameter estimation results with specified groundwater flow using
LSMGA on generated TRT data for 30 days with groundwater velocity 10-8m/s

Specified groundwater Average analysis results based on data generated


velocity used in parameter with groundwater velocity 10-8m/s
estimation (10-8m/s) kg (W/mK) Rb (mK/W) MSD (K2)
0 3.498 0.1252 0.000000
1 3.500 0.1253 0.000000
2 3.501 0.1254 0.000000
5 3.496 0.1252 0.000000
10 3.487 0.1249 0.000002
20 3.446 0.1238 0.000033
50 3.177 0.1163 0.001424
100 2.553 0.0995 0.027394

166
Table L.3 Parameter estimation results with specified groundwater flow using
LSMGA on generated TRT data for 30 days with groundwater velocity 10-7m/s

Specified groundwater Average analysis results based on data generated


velocity used in parameter with groundwater velocity 10-7m/s
estimation (10-8m/s) kg (W/mK) Rb (mK/W) MSD (K2)
0 3.514 0.1257 0.000002
1 3.513 0.1257 0.000002
2 3.513 0.1256 0.000002
5 3.510 0.1256 0.000001
10 3.500 0.1253 0.000000
20 3.459 0.1242 0.000018
50 3.190 0.1166 0.001302
100 2.569 0.1001 0.026593

Table L.4 Parameter estimation results with specified groundwater flow using
LSMGA on generated TRT data for 30 days with groundwater velocity 2x10-7m/s

Specified groundwater Average analysis results based on data generated


velocity used in parameter with groundwater velocity 2x10-7m/s
estimation (10-8m/s) kg (W/mK) Rb (mK/W) MSD (K2)
0 3.555 0.1268 0.000031
1 3.555 0.1267 0.000031
2 3.554 0.1267 0.000030
5 3.551 0.1267 0.000027
10 3.541 0.1264 0.000018
20 3.501 0.1253 0.000000
50 3.227 0.1178 0.000966
100 2.589 0.1007 0.024248

167
Table L.5 Parameter estimation results with specified groundwater flow using
LSMGA on generated TRT data for 30 days with groundwater velocity 5x10-7m/s

Specified groundwater Average analysis results based on data generated


velocity used in parameter with groundwater velocity 5x10-7m/s
estimation (10-8m/s) kg (W/mK) Rb (mK/W) MSD (K2)
0 3.846 0.1338 0.000981
1 3.847 0.1338 0.000980
2 3.845 0.1337 0.000978
5 3.842 0.1337 0.000962
10 3.832 0.1334 0.000908
20 3.788 0.1324 0.000706
50 3.499 0.1253 0.000000
100 2.771 0.1064 0.011992

Table L.6 Parameter estimation results with specified groundwater flow using
LSMGA on generated TRT data for 30 days with groundwater velocity 10-6m/s

Specified groundwater Average analysis results based on data generated


velocity used in parameter with groundwater velocity 10-6m/s
estimation (10-8m/s) kg (W/mK) Rb (mK/W) MSD (K2)
0 4.924 0.1525 0.007976
1 4.925 0.1526 0.007975
2 4.921 0.1525 0.007971
5 4.917 0.1524 0.007945
10 4.905 0.1522 0.007849
20 4.856 0.1514 0.007468
50 4.522 0.1458 0.004940
100 3.501 0.1253 0.000000

168
M. Parameter (Including Groundwater Velocity) Estimation Results
Using LSMGA on Generated TRT Data Using LSMGA for 5
Days at Different Groundwater Velocities

Table M.1 Parameter (including groundwater velocity) estimation results using


LSMGA on generated TRT data for 5 days without groundwater

Analysis results based on data generated with no


groundwater
Trial kg (W/mK) Rb (mK/W) Vgw (m/s) MSD (K2)
1 3.498 0.1253 -1.12x10-7 0.000000
2 3.494 0.1252 -1.26x10-7 0.000000
-7
3 3.499 0.1253 1.05x10 0.000000

Table M.2 Parameter (including groundwater velocity) estimation results using


LSMGA on generated TRT data for 5 days with groundwater velocity 10-8m/s

Analysis results based on data generated with


groundwater velocity 10-8m/s
Trial kg (W/mK) Rb (mK/W) Vgw (m/s) MSD (K2)
1 3.497 0.1252 -1.31x10-7 0.000000
2 3.495 0.1252 -1.46x10-7 0.000000
3 3.495 0.1252 -1.07x10-7 0.000000

Table M.3 Parameter (including groundwater velocity) estimation results using


LSMGA on generated TRT data for 5 days with groundwater velocity 2x10-8m/s

Analysis results based on data generated with


groundwater velocity 2x10-8m/s
Trial kg (W/mK) Rb (mK/W) Vgw (m/s) MSD (K2)
1 3.494 0.1252 1.14x10-7 0.000000
2 3.498 0.1253 1.16x10-7 0.000000
3 3.499 0.1253 -6.80x10-8 0.000000

169
Table M.4 Parameter (including groundwater velocity) estimation results using
LSMGA on generated TRT data for 5 days with groundwater velocity 5x10-7m/s

Analysis results based on data generated with


groundwater velocity 5x10-7m/s
Trial kg (W/mK) Rb (mK/W) Vgw (m/s) MSD (K2)
1 3.499 0.1253 -5.02x10-7 0.000000
2 3.499 0.1253 -5.08x10-7 0.000000
3 3.500 0.1253 -5.01x10-7 0.000000

Table M.5 Parameter (including groundwater velocity) estimation results using


LSMGA on generated TRT data for 5 days with groundwater velocity 10-6m/s

Analysis results based on data generated with


groundwater velocity 10-6m/s
Trial kg (W/mK) Rb (mK/W) Vgw (m/s) MSD (K2)
1 3.507 0.1254 9.88x10-7 0.000000
2 3.503 0.1253 9.93x10-7 0.000000
3 3.504 0.1254 -9.92x10-7 0.000000

Table M.6 Parameter (including groundwater velocity) estimation results using


LSMGA on generated TRT data for 5 days with groundwater velocity 2x10-6m/s

Analysis results based on data generated with


groundwater velocity 2x10-6m/s
Trial kg (W/mK) Rb (mK/W) Vgw (m/s) MSD (K2)
1 3.504 0.1254 -2.00x10-6 0.000000
2 3.503 0.1254 -2.00x10-6 0.000000
3 3.494 0.1252 2.00x10-6 0.000000

170
N. Parameter (Including Groundwater Velocity) Estimation Results
Using LSMGA on Generated TRT Data Using LSMGA for 30
Days at Different Groundwater Velocities

Table N.1 Parameter (including groundwater velocity) estimation results using


LSMGA on generated TRT data for 30 days without groundwater

Analysis results based on data generated with no


groundwater
Trial kg (W/mK) Rb (mK/W) Vgw (m/s) MSD (K2)
1 3.499 0.1253 -3.90x10-8 0.000000
2 3.501 0.1253 1.79x10-8 0.000000
-8
3 3.499 0.1253 4.45x10 0.000000

Table N.2 Parameter (including groundwater velocity) estimation results using


LSMGA on generated TRT data for 30 days with groundwater velocity 10-8m/s

Analysis results based on data generated with


groundwater velocity 10-8m/s
Trial kg (W/mK) Rb (mK/W) Vgw (m/s) MSD (K2)
1 3.496 0.1252 5.09x10-8 0.000000
2 3.500 0.1253 3.71x10-8 0.000000
3 3.499 0.1253 2.58x10-8 0.000000

Table N.3 Parameter (including groundwater velocity) estimation results using


LSMGA on generated TRT data for 30 days with groundwater velocity 2x10-8m/s

Analysis results based on data generated with


groundwater velocity 2x10-8m/s
Trial kg (W/mK) Rb (mK/W) Vgw (m/s) MSD (K2)
1 3.499 0.1253 3.73x10-8 0.000000
2 3.499 0.1253 1.67x10-8 0.000000
3 3.495 0.1252 -5.80x10-8 0.000000

171
Table N.4 Parameter (including groundwater velocity) estimation results using
LSMGA on generated TRT data for 30 days with groundwater velocity 2x10-7m/s

Analysis results based on data generated with


groundwater velocity 2x10-7m/s
Trial kg (W/mK) Rb (mK/W) Vgw (m/s) MSD (K2)
1 3.498 0.1252 -2.05x10-7 0.000000
2 3.499 0.1252 -2.02x10-7 0.000000
3 3.501 0.1253 1.97x10-7 0.000000

Table N.5 Parameter (including groundwater velocity) estimation results using


LSMGA on generated TRT data for 30 days with groundwater velocity 5x10-7m/s

Analysis results based on data generated with


groundwater velocity 5x10-7m/s
Trial kg (W/mK) Rb (mK/W) Vgw (m/s) MSD (K2)
1 3.502 0.1254 -5.00x10-7 0.000000
2 3.499 0.1252 -5.00x10-7 0.000000
3 3.499 0.1253 -5.01x10-7 0.000000

Table N.6 Parameter (including groundwater velocity) estimation results using


LSMGA on generated TRT data for 30 days with groundwater velocity 10-6m/s

Analysis results based on data generated with


groundwater velocity 10-6m/s
Trial kg (W/mK) Rb (mK/W) Vgw (m/s) MSD (K2)
1 3.495 0.1252 1.00x10-6 0.000000
2 3.500 0.1253 9.99x10-7 0.000000
3 3.499 0.1253 -9.99x10-7 0.000000

172
O. Logged Temperature Profiles in Experimental Thermal Response
Tests

Temperature profiles in Test A


Borehole Room Circulating water

45
Temperature (degC)

40

35

30

25

20
0 20 40 60 80 1 00

Test period (hour)

Figure O.1 Logged temperature profiles for Test A in experimental TRT

Temperature profiles in Test B


Borehole Room Circulating water

40
Temperature (degC)

35

30

25

20
0 20 40 60 80 1 00

Test period (hour)

Figure O.2 Logged temperature profiles for Test B in experimental TRT

Temperature profiles in Test C


Borehole Room Circulating water

40
Temperature (degC)

35

30

25

20
0 20 40 60 80 1 00

Test period (hour)

Figure O.3 Logged temperature profiles for Test C in experimental TRT

173
Temperature profiles in Test D
Borehole Room Circulating water

40

Temperature (degC)
35

30

25

20
0 20 40 60 80 1 00

Test period (hour)

Figure O.4 Logged temperature profiles for Test D in experimental TRT

Temperature profiles in Test E


Borehole Room Circulating water

45
Temperature (degC)

40

35

30

25

20
0 20 40 60 80 1 00

Test period (hour)

Figure O.5 Logged temperature profiles for Test E in experimental TRT

Temperature profiles in Test F


Borehole Room Circulating water

40
Temperature (degC)

35

30

25

20
0 20 40 60 80 1 00

Test period (hour)

Figure O.6 Logged temperature profiles for Test F in experimental TRT

174
P. Dynamic Simulation Results of GCLDAC/GSHP with
Groundwater Flow

Table P.1 Simulation results of GSHP/GCLDAC with Darcy velocity 10-7m/s of


groundwater flow for Case 1

GSHP for GSHP for GCLDAC


1986 1986~1995 for 1986
Total cooling energy to room (kWh) 32,565.62 336,056.56 32,460.11
Total heating energy to borefield (kWh) 39,588.63 408,584.38 35,477.11
Total energy input to compressor (kWh) 6,527.19 67,723.31 6,570.16
Maximum borefield fluid leaving 30.89 31.72 30.99
temperature (oC)

Table P.2 Simulation results of GSHP/GCLDAC with Darcy velocity 5x10-7m/s of


groundwater flow for Case 1

GSHP for GSHP for GCLDAC


1986 1986~1995 for 1986
Total cooling energy to room (kWh) 32,542.99 336,094.84 32,459.10
Total heating energy to borefield (kWh) 39,604.32 408,414.91 35,471.20
Total energy input to compressor (kWh) 6,602.40 67,583.57 6,636.52
Maximum borefield fluid leaving 31.46 31.68 31.46
temperature (oC)

Table P.3 Simulation results of GSHP/GCLDAC with Darcy velocity 10-6m/s of


groundwater flow for Case 1

GSHP for GSHP for GCLDAC


1986 1986~1995 for 1986
Total cooling energy to room (kWh) 32,550.63 336,115.53 32,446.64
Total heating energy to borefield (kWh) 39,604.61 408,299.22 35,449.02
Total energy input to compressor (kWh) 6,603.24 67,547.20 6,649.20
Maximum borefield fluid leaving 31.46 31.65 31.60
temperature (oC)

175
Table P.4 Simulation results of GSHP/GCLDAC with Darcy velocity 10-7m/s of
groundwater flow for Case 2

GSHP for GSHP for GCLDAC


1986 1986~1995 for 1986
Total cooling energy to room (kWh) 44,855.52 462,832.34 44,810.99
Total heating energy to borefield (kWh) 54,364.38 561,528.63 46,419.65
Total energy input to compressor (kWh) 8,725.58 90,697.45 8,679.83
Maximum borefield fluid leaving 30.77 31.60 30.84
o
temperature ( C)

Table P.5 Simulation results of GSHP/GCLDAC with Darcy velocity 5x10-7m/s of


groundwater flow for Case 2

GSHP for GSHP for GCLDAC


1986 1986~1995 for 1986
Total cooling energy to room (kWh) 44,841.37 462,896.82 44,798.82
Total heating energy to borefield (kWh) 54,468.42 561,578.88 46,424.90
Total energy input to compressor (kWh) 8,849.47 90,707.00 8,804.38
Maximum borefield fluid leaving 31.54 31.70 31.71
temperature (oC)

Table P.6 Simulation results of GSHP/GCLDAC with Darcy velocity 10-6m/s of


groundwater flow for Case 2

GSHP for GSHP for GCLDAC


1986 1986~1995 for 1986
Total cooling energy to room (kWh) 44,831.63 462,929.16 44,805.61
Total heating energy to borefield (kWh) 54,438.52 561,359.88 46,415.61
Total energy input to compressor (kWh) 8,833.55 90,513.59 8,787.97
Maximum borefield fluid leaving 31.44 31.61 31.55
temperature (oC)

176
References

Abramowitz, M., and I.A. Stegun (1964). Handbook of Mathematical Functions,


National Bureau of Standards, Applied Mathematics Series 55, U.S.
Department of Commerce.

Arundel, A.V., E.M. Sterling, J.H. Biggin and T.D. Sterling (1992). Indirect health
effects of relative humidity in indoor environments, Desiccant Cooling and
Dehumidification, Atlanta: AHSRAE, 3-12.

ASHRAE (1995). Commercial/Institutional Ground-Source Heat Pump Engineering


Manual, Atlanta: ASHRAE.

ASHRAE (1998). Operating Experiences with Commercial Ground-Source Heat


Pump Systems, Atlanta: ASHRAE.

ASHRAE (2003). ASHRAE Handbook: Heating, Ventilation and Air-Conditioning


Applications, Atlanta: ASHRAE.

ASHRAE (2004). ASHRAE Handbook: HVAC Systems and Equipment, Atlanta:


ASHRAE.

ASHRAE (2005). ASHRAE Handbook: Fundamental, Atlanta: ASHRAE.

ASHRAE (2006). ASHRAE Handbook: Refrigeration, Atlanta: ASHRAE.

Austin, W.A. (1998). Development of an In-situ System for Measuring Ground


Thermal Properties, Master Thesis, Oklahoma State University.

Austin, W.A., C. Yavuzturk and J.D. Spitler (2000). Development of an in-situ


system and analysis procedure for measuring ground thermal properties,
ASHRAE Transactions, 106(1), 365-379.

177
Bejan, A. (1993). Heat Transfer, Wiley.

Bennet, J., J. Claesson and G. Hellstrom (1987). Multipole method to compute the
conductive heat flows to and between pipes in a composite cylinder, Notes on
Heat Transfer 3-1987, Department of Building Physics and Mathematical
Physics, Lund Institute of Technology, Lund, Sweden.

Bernier, M.A. (2000). A review of the cylindrical heat source method for the design
and analysis of vertical ground-coupled heat pump systems, Fourth
International Conference on Heat Pumps in Cold climates, Aylmer, Quebec.

Bernier, M.A. (2001). Ground-coupled heat pump system simulation; ASHRAE


Transactions, 107(1), 605-616.

Bernier, M.A., P. Pinel, R. Labib and R. Paillot (2004). A multiple load aggregation
algorithm for annual hourly simulations of GCHP systems, HVAC&R
Research, 10(4), 471-487.

Bernier, M.A. (2006). Closed-loop ground-coupled heat pump systems, ASHRAE


Journal, September, 13-24.

Bose, J.E., J.D. Parker and F.C. McQuiston (1985). Design/Data Manual for Closed-
Loop Ground-Coupled Heat Pump Systems, Altanta, ASHRAE.

Bose, J.E., M.D. Smith and J.D. Spitler (2002). Advances in ground source heat
pump systems - an international overview, Proceedings of the Seventh
International Energy Agency Heat Pump Conference, Beijing, 313-324.

Bradley, D. (2007). Personal communication.

Brandl, H. (2006). Energy foundations and other thermo-active ground structures,


Geotechnique, 56(2), 81-122.

178
Cane, R.L.D., and D.A. Forgas (1991). Modeling of ground source heat pump
performance, ASHRAE Transactions, 97(1), 909-925.

Carslaw, H.S., and J.C. Jaeger (1959). Conduction of Heat in Solids, Oxford:
Claremore Press.

Cecchini, C., and D. Marchal (1991). A simulation model of refrigerant and air-
conditioning equipment based on experimental data, ASHRAE Transactions,
97(2), 388-393.

Chiasson, A.D., S.J. Rees and J.D. Spitler (2000). A preliminary assessment of the
effects of groundwater flow on closed-loop ground-source heat pump systems,
ASHRAE Transactions, 106(1), 380-393.

Chung, T.W. (1994). Predictions of moisture removal efficiencies for packed-bed


dehumidification systems, Gas Separation and Purification, 8(4), 265-268.

Chung, T.W., T.K. Ghosh and A.L. Hines (1996). Comparison between random and
structured packings for dehumidification of air by lithium chloride solutions
in a packed column and their heat and mass transfer correlations, Industrial
and Engineering Chemistry Research, 35(1), 192-198.

Claesson, J., and G. Hellstrom (1987). Thermal Resistances to and between Pipes in
a Composite Cylinder, Department of Mathematical Physics and Building
Technology, University of Lund, Lund, Sweden.

Claesson, J., and G. Hellstrom (2000). Analytical studies of the influence of regional
groundwater flow on the performance of borehole heat exchangers,
Proceedings of the 8th international Conference on Thermal Energy Storage,
Terrastock 2000, Stuttgart, Germany, 195-200.

Conde, M.R. (2004). Properties of aqueous solutions of lithium and calcium


chlorides: formulations for use in air conditioning equipment design,
International Journal of Thermal Sciences, 43(4), 367-382.

179
Cooper, L.Y. (1976). Heating of a cylindrical cavity; International Journal of Heat
and Mass Transfer, 19(7), 575-577.

Deerman, J.D., and S.P. Kavanaugh (1991). Simulation of vertical U-tube ground
coupled heat pump systems using the cylindrical heat source solution,
ASHRAE Transactions, 97(1), 287-295.

Diao, N., Q. Li and Z. Fang (2004). Heat transfer in ground heat exchangers with
groundwater advection, International Journal of Thermal Sciences, 43(12),
1203-1211.

Diao, N.R., H.Y. Zeng and Z.H. Fang (2004). Improvement in modeling of heat
transfer in vertical ground heat exchangers, HVAC&R Research, 10(4), 459-
470.

Diersch, H.J.G. (2002). FEFLOW Reference Manual, Institute of Water Resources


Planning and System Research Limited.

Dittus, F.W., and L.M.K. Boelter (1930). University of California (Berkeley) Pub.
Eng., 2, 443.

Dobson, M.K., D.L. O'Neal and W. Aldred (1995). A modified analytical method for
simulating cyclic operation of vertical U-tube ground-coupled heat pumps,
Proceedings of the 1995 ASME/JSME/JSES International Solar Energy
Conference, 1, 69-76.

Domanski, P., and D. Didion (1984). Mathematical model of an air-to-air heat pump
equipped with a capillary tube, International Journal of Refrigeration, 7(4),
249-255.

Eklof, C., and S. Gehlin (1996). TED - a mobile equipment for thermal response test,
Masters Thesis, Lulea University of Technology, Sweden.

180
Elsarrag, E., E.E.M. Magzoub and S. Jain (2004). Mass-transfer correlations for
dehumidification of air by triethylene glycol in a structured packed column,
Industrial and Engineering Chemistry Research, 43(23), 7676-7681.

Elsarrag, E. (2006). Dehumidification of air by chemical liquid desiccant in a packed


column and its heat and mass transfer effectiveness, HVAC&R Research,
12(1), 3-16.

Elsayed, M.M. (1994). Analysis of air dehumidification using liquid desiccant


system, Renewable Energy, 4(5), 519-528.

Eskilson, P. (1987). Thermal Analysis of Heat Extraction Boreholes, Doctoral Thesis,


264p, Department of Mathematical Physics and Building Technology,
University of Lund.

Eskilson, P., and J. Claesson (1988). Simulation model for thermally interacting heat
extraction boreholes, Numerical Heat Transfer, 13, 149-165.

Factor, H.M. and G. Grossman (1980). A packed bed dehumidifier/regenerator for


solar air conditioning with liquid desiccants, Solar Energy, 24(6), 541-550.

Fromentin, A., D. Pahud and G. Sarlos (1998). Heating and cooling systems with
heat exchanger piles, Proceedings of the International Conference on Energy
and Environment, ICEE, 230-234.

Fujii, H., R. Itoi, and T. Ishikami (2004a). Improvements on analytical modeling for
vertical U-tube ground heat exchangers, Geothermal Resources Council
Transactions, 28, August 29 September 1, 73-77.

Fujii, H., R. Itoi, J. Fujii, Y. Uchida and T. Ishikami (2004b). Numerical simulation
of large-scale GCHP systems in the presence of groundwater flow,
Geothermal Resources Council Transaction, 28, August 29 September 1,
79-84.

181
Fumo, N., and D.Y. Goswami (2002). Study of an aqueous lithium chloride desiccant
system: air dehumidification and desiccant regeneration, Solar Energy, 72(4),
351-361.

Gandhidasan, P., C.F. Kettleborough and M.R. Ullah (1986). Calculation of heat and
mass transfer coefficients in a packed tower operating with a desiccant-air
contact system, Transactions of the ASME Journal of Solar Energy
Engineering, 108, 123-128.

Gandhidasan, P. (2004). A simplified model for air dehumidification with liquid


desiccant, Solar Energy, 76(4), 409-416.

Gasparella, A., G.A. Longo and R. Marra (2005). Combination of ground source
heat pumps with chemical dehumidification of air, Applied Thermal
Engineering, 25(2-3), 295-308.

Gehlin, S.E.A., and G. Hellstrom (2003a). Comparison of four models for thermal
response test evaluation, ASHRAE Transactions, 109(1), 131-142.

Gehlin, S.E.A., and G. Hellstrom (2003b). Influence on thermal response test by


groundwater flow in vertical fractures in hard rock, Renewable Energy,
28(14), 2221-2238.

Georgiev, A., A. Busso and P. Roth (2006). Shallow borehole heat exchange:
Response test and charging-discharging test with solar collectors, Renewable
Energy, 31(7), 971-985.

Ghaddar, N., K. Ghali and A. Najm (2003). Performance of solar-assisted hybrid air-
conditioning liquid desiccant system in Beirut, Proceedings of Energy and
the Environment 2003, Halkidiki, Greece, 241-251.

Goulburn, J.R., and J. Fearon (1983). Domestic heat pump with deep hole ground
source evaporator, Applied Energy, 14(2), 99-113.

182
Griffiths, W.C. (1992). Use of liquid sorption dehumidification to improve energy
utilisation of air systems, Desiccant Cooling and Dehumidification, Atlanta:
AHSRAE, 33-39.

Gu, Y., and D.L. O'Neal (1995). An analytical solution to transient heat conduction
in a composite region with a cylindrical heat source, Transactions of the
ASME Journal of Solar Energy Engineering, 117, 242-248.

Gu, Y., and D.L. O'Neal (1998a). Development of an equivalent diameter expression
for vertical U-tubes used in ground-coupled heat pumps, ASHRAE
Transactions, 104(2), 347-355.

Gu, Y., and D.L. O'Neal (1998b). Modeling the effect of backfills on U-tube ground
coil performance, ASHRAE Transactions, 104(2), 356-365.

Hamada, Y., H. Saitoh, M. Nakamura, H. Kubota and K. Ochfuji (2007). Field


performance of an energy pile for spacing heating, Energy and Buildings,
39(5), 517-524.

Hart, D.P., and R. Couvillion (1986). Earth Coupled Heat Transfer, Prepared for the
National Water Well Association, Dublin.

He, M.M. (2007). Analysis of Underground Thermal Energy Storage Systems with
Ground Water Advection in Subtropical Regions, Master Thesis, Department
of Mechanical Engineering, University of Hong Kong.

Hellstrom, G. (1989). Duct Ground Heat Storage Model, Manual for Computer Code,
Department of Mechanical Physics, University of Lund, Sweden.

Hellstrom, G. (1991). Ground Heat Storage. Thermal Analysis of Duct Storage


Systems: Part I Theory, Doctoral Thesis, Department of Mathematical
Physics, University of Lund, Sweden.

183
Hellstrom, G., L. Mazzarella and D. Pahud (1996). Duct Ground Heat Storage
Model, Lund DST TRNSYS13.1 Version January 1996, Department of
Mathematical Physics, University of Lund, Sweden.

Hellstrom, G., and B. Sanner (2000). Earth Energy Designer, User Manual, Version
2.0.

Hepbasli, A., and O. Akdemir (2004). Energy and exergy analysis of a ground source
(geothermal) heat pump system, Energy Conversion and Management, 45(5),
737-753.

Holman J.P. (1990). Heat Transfer, New York: McGraw-Hill.

Ingersoll, L.R., O.J. Zobel and A.C. Ingersoll (1954). Heat Conduction: with
Engineering, Geological and Other Applications, Madison: University of
Wisconsin Press.

Jain, N.K. (1999). Parameter Estimation of Ground Thermal Properties, Masters


Thesis, Oklahoma State University.

Jin, H., and J.D. Spitler (2002). A parameter estimation based model of water-to-
water heat pumps for use in energy calculation programs, ASHRAE
Transactions, 108(1), 3-17.

Jung, D., C. Park and B. Park (1999). Capillary tube selection for HCFC22
alternatives, International Journal of Refrigeration, 22(7), 604-614.

Katsura, T., K. Nagano, S. Takeda and K. Shimakura (2006). Heat transfer


experiment in the ground with ground water advection, Proceedings of
Ecostock 2006, New Jersey.

Kavanaugh, S.P. (1985). Simulation and Experimental Verification of Vertical


Ground-Coupled Heat Pump Systems, PhD Dissertation, Oklahoma State
University.

184
Kavanaugh, S.P. (1992). Simulation of ground-coupled heat pumps with an
analytical solution, ASME-JSES-KSES International Solar Energy
Conference, 395-400.
Kavanaugh, S.P. (1995). A design method for commercial ground-coupled heat
pumps, ASHRAE Transactions, 101(2), 1088-1094.

Kavanaugh, S.P. (1998). A design method for hybrid ground-source heat pumps,
ASHRAE Transactions, 104(2), 691-698.

Kavanaugh, S.P., and K. Rafferty (1997). Ground-Source Heat Pumps: Design of


Geothermal Systems for Commercial and Institutional Buildings, Altanta:
ASHRAE.

Kepinska, B. (2003). Current state and prospects of geothermal energy


implementation in Poland, Applied Energy, 74(1-2), 43-51.

Khan, A.Y., and H.D. Ball (1992). Development of a generalised model for
performance evaluation of packed-type liquid sorbent dehumidifiers and
regenerators, ASHRAE Transactions, 98(1), 525-533.

Kasuda, T., and P.R. Archenbach (1965). Earth temperature and thermal diffusivity
at selected stations in the United States, ASHRAE Transactions, 71(1), 61-75.

Kyriakis, N., A. Michopoulos and K. Pattas (2006). On the maximum thermal load of
ground heat exchangers; Energy and Buildings, 38(1), 25-29.

Laloui, L., M. Nuth and L. Vulliet (2006). Experimental and numerical investigations
of the behaviour of a heat exchanger pile, International Journal for
Numerical and Analytical Methods in Geomechanics, 30(8), 763-781.

Lam, H.N., and H.M. Wong (2005). Geothermal heat pump system for air
conditioning in Hong Kong, Proceedings of World Geothermal Congress
2005, Antalya, Turkey, article no. 1475.

185
Lau, K.F., and M.T. Suen (2003). Geothermal heat pump air-conditioning system for
the Hong Kong International Wetland Park, Proceedings of Shandong-Hong
Kong Joint Symposium 2003, A1-A9.

Lamarche, L., and B. Beauchamp (2007). A new contribution to the finite line-source
model for geothermal boreholes, Energy and Buildings, 39(2), 188-198.

Lazzarin, R.M., A. Gasparella and G.A. Longo (1999). Chemical dehumidification


by liquid desiccants: theory and experiment, International Journal of
Referigeration, 22(4), 334-347.

Lazzarin, R.M., and F. Castellotti (2007). A new heat pump desiccant dehumidifier
for supermarket application, Energy and Buildings, 39(1), 59-65.

Li, X., J. Zhao and Q. Zhou (2005). Inner heat source model with heat and moisture
transfer in soil around the underground heat exchanger, Applied Thermal
Enginering, 25(10), 1565-1577.

Li, X., Z. Chen and J. Zhao (2006). Simulation and experiment on the thermal
performance of U-vertical ground coupled heat exchanger, Applied Thermal
Engineering, 26(14-15), 1564-1571.

Liu, X.H., K.Y. Qu and Y. Jiang (2006). Empirical correlations to predict the
performance of the dehumidifier using liquid desiccant in heat and mass
transfer, Renewable Energy, 31(10), 1627-1639.

Lund, J.W. (2003). Direct-use of geothermal energy in the USA, Applied Energy,
74(1-2), 33-42.

Mago, P.J., L. Chamra and G. Steele (2006). A simulation model for the performance
of a hybrid liquid desiccant system during cooling and dehumidification,
International Journal of Energy Research, 30(1), 51-66.

186
Martin, V., and D.Y. Goswami (2000). Effectiveness of heat and mass transfer
processes in a packed bed liquid desiccant dehumidifier/regenerator,
HVAC&R Research, 6(1), 21-39.

Mei, V.C. (1988). Heat pump ground coil analysis with thermal interference,
Transactions of the ASME Journal of Solar Energy Engineering, 110, 69-73.

Mei, V.C. (1991). Heat transfer of buried pipe for heat pump application,
Transactions of the ASME Journal of Solar Energy Engineering, 113, 51-55.

Mogensen, P. (1983). Fluid to duct wall heat transfer in duct system heat storages,
Proceedings of the International Conference on Subsurface Heat Storage in
Theory and Practice, Stockholm, Sweden, 652-657.

Morino, K., and T. Oka (1994). Study on heat exchanged in soil by circulating water
in a steel pile, Energy and Buildings, 21(1), 65-78.

Moujaes, S.F. (1990). Cyclic simulation of a model describing heat transfer from a
ground-coupled water source heat pump, considering transient effects on both
soil and water sides, International Journal of Refrigeration, 13(5), 330-335.

Muraya, N.K., D.L. O'Neal and W.M. Heffington (1996). Thermal interference of
adjacent legs in a vertical U-tube heat exchanger for a ground-coupled heat
pump, ASHRAE Transactions, 102(2), 12-21.

Nelder, J.A., and R. Mead (1965). A simplex method for function minimization,
Computer Journal, 7(1), 308-313.

Nield, D.A., and A. Bejan. (1992). Convection in Porous Media, Springer-Veriag,


New York.

Niibori, Y., Y. Iwata, S. Ichinose and G. Fukaya (2005). Design of the BHP system
considering the heat transport of groundwater flow, Proceedings of World
Geothermal Congress 2005, Antalya, Turkey, article no. 1422.

187
Ozgener, O., and A. Hepbasli (2006). An economical analysis on a solar greenhouse
integrated solar assisted geothermal heat pump system, Journal of Energy
Resources Technology, Transactions of the ASME, 128(1), 28-34.

Pahud, D. (1999). PILESIM - LASEN: Simulation Tool for Heating/Cooling Systems


with Heat Exchanger Piles or Borehole Heat Exchangers, User Manual;
Laboratory of Energy Systems, Swiss Federal Institute of Technolgy in
Lausanne, Switzerland.

Pahud, D. (2001). Two response tests of two <<identical>> boreholes drilled to a


depth of 160m near Lazern, Proceedings of the workshop Tests de reponse
geothemiques, EPFL Lausanne, 37-47.

Pahud, D. (2006). Private communication.

Pahud, D., A. Fromentin and J.C. Hadorn (1996). The Duct Ground Heat Storage
Model (DST) for TRNSYS Used for the Simulation of Heat Exchanger Piles.
User Manual, December 1996 Version, Internal Report, Laboratory of
Energy Systems (LASEN), Swiss Federal Institute of Technology (EPFL),
Lausanne, Switzerland.

Pahud, D., and G. Hellstrom (1996). The new duct ground heat model for TRNSYS,
Proceedings of Eurotherm Seminar No49, Eindhoven, The Netherlands, 27-
136.

Pahud, D., A. Formentin and M. Hubbuch (1999). Heat Exchanger Pile System of the
Dock Midfield at the Zurich Airport - Detailed Simulation and Optimization
of the Installation, Final report, Swiss Federal Office of Energy, Switzerland.

Pahud, D., and B. Mathey (2001). Comparison of the thermal performance of double
U-pipe borehole heat exchangers measured in situ, Energy and Buildings,
33(5), 503-507.

188
Patek, J., and J. Klomfar (2006). A computationally effective formulation of the
thermodynamic properties of LiBr-H2O solutions from 273 to 500K over full
composition range, International Journal of Refrigeration, 29(4), 556-578.

Petit, P.J., and J.P. Meyer (1998). Economic potential of vertical ground-source heat
pumps compared to air-source air conditioners in South Africa, Energy, 23(2),
137-143.

Parise, J.A.R. (1986). Simulation of vapour-compression heat pumps, Simulation,


46(2), 71-76.

Poulikakos, D. (1994). Conduction Heat Transfer, Englewood Cliffs, Prentice Hall.

Ren, C., Y. Jiang and Y. Zhang (2006). Simplified analysis of coupled heat and mass
transfer processes in packed bed liquid desiccant-air contact system, Solar
Energy, 80(1), 121-131.

Rottmayer, S.P., W.A Beckman and J.W. Mitchell (1997). Simulation of a single
vertical U-tube ground heat exchanger in an infinite medium, ASHRAE
Transactions, 103(2), 651-659.

Sachs, H.M. (2002). Geology and Drilling Methods for Ground-Source Heat Pump
System Installations: An Introduction for Engineers, Atlanta: ASHRAE.

Sanner, B., C. Karytsas, D. Mendrinos and L. Rybach (2003). Current status of


ground source heat pumps and underground thermal energy storage in Europe,
Geothermics, 32(4-6), 579-588.

Smith, M.D., R.L. Perry and W.A. Holloway (1999). Development of a system for
verification of transient in-situ testing models and development of a testing
standard, Research Update Report for the Department of Energy, Oklahoma
State University, Division of Engineering Technology.

189
Shonder, J.A., and J.V. Beck (1999). Determining effective soil formation thermal
properties from field data using a parameter estimation technique, ASHRAE
Transactions, 105(1), 458-466.

Shonder, J.A., V. Baxter, J. Thornton and P. Hughes (1999). A new comparison of


vertical ground heat exchanger design methods for residential applications,
ASHRAE Transactions, 105(2), 1179-1188.

Shonder, J.A., V. Baxter, J. Thornton and P. Hughes (2000). A new comparison of


vertical ground heat exchanger design software for commercial applications,
ASHRAE Transactions, 106(1), 831-842.

Sliwa, T., and J. Kotyza (2003). Application of existing wells as ground heat source
for heat pumps in Poland, Applied Energy, 74(1-2), 3-8.

Sliwa, T., and A. Gonet (2005). Theoretical model of borehole heat exchanger,
Transactions of the ASME Journal of Energy Resources Technology, 127(2),
142-148.

Spiegel, M.R., and J. Liu (1999). Mathematical Handbook of Formulas and Tables,
New York: McGraw-Hill.

Spitler, J.D. (2000). GLHEPRO - A design tool for commercial building ground loop
heat exchangers, Proceedings of the Fourth International Heat Pumps in
Cold Climates Conference, Aylmer, Quebec.

Spitler, J.D., S.J. Rees and C. Yavuzturk (2000b). Recent developments in ground
source heat pump system design, modeling and application, Proceedings of
CIBSE/ASHRAE joint conference "20 20 Vision", Dublin, Session 9a, paper
A28.

Spitler, J.D. (2005). Ground-source heat pump system research-past, present and
future, HVAC&R Research, 11(2), 165-167.

190
Spitler, J.D. (2006). Private communication.

Stefanuk, N.B.M., J.D. Aplevich and M. Renksizbulut (1992). Modeling and


simulation of a superheat-controlled water-to-water heat pump, ASHRAE
Transactions, 98(2), 172-184.

Stoecker, W.F., and J.W. Jones. (1982). Refrigeration and Air-Conditioning, New
York: McGraw-Hill.

Studak, J.W., and J.L. Peterson (1992). A preliminary evaluation of alternative


liquid desiccants for a hybrid desiccant air conditioner, Desiccant Cooling
and Dehumidification, Atlanta: AHSRAE, 129-133.

Sutton, M.G., R.J. Couvillion, D.W. Nutter and R.K. Davis (2002). An algorithm for
approximating the performance of vertical bore heat exchangers installed in a
stratified geological regime, ASHRAE Transactions, 108(2), 177-184.

Thornton, J.W., T.P. McDowell and P.J. Hughes (1997a). Comparison of practical
vertical ground heat exchanger sizing methods to a Fort Polk data/model
benchmark, ASHRAE Transactions, 103(2), 675-683.

Thornton, J.W., T.P. McDowell, J.A. Shonder, P.J. Hughes, D. Pahud and G.A. J
Hellstrom (1997b). Residential vertical geothermal heat pump system
models: Calibration to data, ASHRAE Transactions, 103(2), 660-674.

Treybal, R.E. (1969). Adiabatic gas absorption and stripping in packed towers,
Industrial and Engineering Chemistry, 61(7), 36-41.

Wageningen, N., M. Waegemasekers, B. Brunekreef and J. Boleij (1992). Health


complaints and indoor molds in relation to moisture problems in homes,
Desiccant Cooling and Dehumidification, Atlanta: AHSRAE, 16-18.

191
Wagner R., and C. Clauser (2005). Evaluating thermal response tests using
parameter estimation for thermal conductivity and thermal capacity, Journal
of Geophysics and Engineering, 2(4), 349-356.

Yavuzturk, C. (1999). Modeling of Vertical Ground Loop Heat Exchangers for


Ground Source Heat Pump Systems, Doctoral Thesis, Oklahoma State
University.

Yavuzturk, C., and A.D. Chiasson (2002). Performance analysis of U-tube,


concentric tube, and standing column well ground heat exchangers using a
system simulation approach, ASHRAE Transactions, 108(1), 925-938.

Young, T.R. (2004). Development, Verification, and Design Analysis of the Borehole
Fluid Thermal Mass Model for Approximating Short Term Borehole Thermal
Response, Master Thesis, Oklahoma State University.

Yuan, Y.J., and V. Alberce (2004). Gas heat and mass transfer method based on
self-microcirculation of caloric infinitesimal liquid, China Patent
PRC200410015955.

Zeng, H.Y., N.R. Diao and Z.H. Fang (2002). A finite line-source model for
boreholes in geothermal heat exchangers, Heat Transfer - Asian Research,
31(7), 558-567.

Zeng, H., N. Diao and Z. Fang (2003). Heat transfer analysis of boreholes in vertical
ground heat exchangers, International Journal of Heat and Mass Transfer,
46(23), 4467-4481.

192

Anda mungkin juga menyukai