ArticlePDF Available

Socio-economic performance of a novel solar photovoltaic/loop-heat-pipe heat pump water heating system in three different climatic regions

Authors:

Abstract and Figures

h i g h l i g h t s We predict the system's socio-economic performance in three climatic regions. The system achieved the highest energy efficiency in Hong Kong area. The system had the best economic revenue in London area. The system obtained the most environmental benefits in Shanghai area. The system's suitability is dependent on the priority order of the three factors. a b s t r a c t This paper aimed to study the socio-economic performance of a novel solar photovoltaic/loop-heat-pipe (PV/LHP) heat pump water heating system for application in three different climatic regions, namely, cold area represented by London, warm area represented by Shanghai, and hot (subtropical) area represented by Hong Kong. This study involved prediction of the annual fossil-fuel energy saving, investment return period and carbon emission reduction of the new system against the traditional gas-fired and electrical boilers based water heating systems. An established dynamic model developed by the authors was uti-lised to predict the system's energy performance throughout a year in the three climatic regions. A life-cycle analytical model was further developed to analyse the economic and environmental benefits of the new system relative to the traditional systems. Analyses of the modelling results drew out several conclusive remarks: (1) the system could achieve the highest energy efficiency when operating at the hot (subtropical) climatic region (represented by Hong Kong), enabling the heat output of as high as 922 kW h/m 2 yr and water temperature of above 45 °C, while the grid power input is only 59 kW h/ m 2 yr; (2) the system is worth for investment when operating at the high energy charging tariff area (rep-resented by London), with the cost payback periods of 8 and 5 years relative to the traditional gas-fired and electrical boilers based systems, respectively; (3) the system could obtain the most promising envi-ronmental benefits when operating in Shanghai where the energy quality (embodied carbon volume of per kW h energy) is relatively poor, enabling reduction in life-cycle carbon emissions of around 4.08 tons/m 2 and 17.87 tons/m 2 respectively, relative to the gas-fired and electrical boilers. Answer to such a question on which area is most suitable for the system application is highly dependent upon the priority order among the three dominating factors: (1) energy efficiency, (2) economic revenue, and (3) environmental benefit, which may vary with the users, local concerns and policy influence, etc. The research results will be able to assist in decision making in implementation of the new PV/thermal technology and analyses of the associated economic and environmental benefits, thus contributing to realisation of the regional and global targets on fossil fuel energy saving and environmental sustainability.
Content may be subject to copyright.
Socio-economic performance of a novel solar
photovoltaic/loop-heat-pipe heat pump water
heating system in three different climatic regions
Xingxing Zhang
a,
, Jingchun Shen
a
, Peng Xu
a,b
, Xudong Zhao
a,
, Ying Xu
c
a
School of Engineering, University of Hull, HU6 7RX, UK
b
Beijing University of Civil Engineering and Architecture, Beijing 100044, China
c
Shanghai Pacific Energy Centre, Shanghai 200001, China
highlights
We predict the system’s socio-economic performance in three climatic regions.
The system achieved the highest energy efficiency in Hong Kong area.
The system had the best economic revenue in London area.
The system obtained the most environmental benefits in Shanghai area.
The system’s suitability is dependent on the priority order of the three factors.
article info
Article history:
Received 19 April 2014
Received in revised form 10 July 2014
Accepted 18 August 2014
Keywords:
PV
Loop heat pipe
Simulation
Energy performance
Economic
Environment
abstract
This paper aimed to study the socio-economic performance of a novel solar photovoltaic/loop-heat-pipe
(PV/LHP) heat pump water heating system for application in three different climatic regions, namely, cold
area represented by London, warm area represented by Shanghai, and hot (subtropical) area represented
by Hong Kong. This study involved prediction of the annual fossil-fuel energy saving, investment return
period and carbon emission reduction of the new system against the traditional gas-fired and electrical
boilers based water heating systems. An established dynamic model developed by the authors was uti-
lised to predict the system’s energy performance throughout a year in the three climatic regions. A
life-cycle analytical model was further developed to analyse the economic and environmental benefits
of the new system relative to the traditional systems. Analyses of the modelling results drew out several
conclusive remarks: (1) the system could achieve the highest energy efficiency when operating at the hot
(subtropical) climatic region (represented by Hong Kong), enabling the heat output of as high as
922 kW h/m
2
yr and water temperature of above 45 °C, while the grid power input is only 59 kW h/
m
2
yr; (2) the system is worth for investment when operating at the high energy charging tariff area (rep-
resented by London), with the cost payback periods of 8 and 5 years relative to the traditional gas-fired
and electrical boilers based systems, respectively; (3) the system could obtain the most promising envi-
ronmental benefits when operating in Shanghai where the energy quality (embodied carbon volume of
per kW h energy) is relatively poor, enabling reduction in life-cycle carbon emissions of around
4.08 tons/m
2
and 17.87 tons/m
2
respectively, relative to the gas-fired and electrical boilers. Answer to
such a question on which area is most suitable for the system application is highly dependent upon
the priority order among the three dominating factors: (1) energy efficiency, (2) economic revenue,
and (3) environmental benefit, which may vary with the users, local concerns and policy influence, etc.
The research results will be able to assist in decision making in implementation of the new PV/thermal
technology and analyses of the associated economic and environmental benefits, thus contributing to
realisation of the regional and global targets on fossil fuel energy saving and environmental
sustainability.
Ó2014 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.apenergy.2014.08.074
0306-2619/Ó2014 Elsevier Ltd. All rights reserved.
Corresponding authors. Tel.: +44 (0)1482 466684; fax: +44 (0)1482 466664.
E-mail addresses: Xingxing.zhang@hull.ac.uk (X. Zhang), Xudong.zhao@hull.ac.uk (X. Zhao).
Applied Energy 135 (2014) 20–34
Contents lists available at ScienceDirect
Applied Energy
journal homepage: www.elsevier.com/locate/apenergy
1. Introduction
In contemporary energy sector, solar photovoltaic (PV) and
solar thermal are the fundamental pillars to assist in transition
from the traditional fossil fuel energy structure to a renewable
energy system. Recent governmental schemes addressed that by
2050, the solar PV will generate nearly 11% of global electricity
[1] while the solar thermal will deliver about 50% of the low and
medium temperature heat in the EU [2]. At present, the technical
drawback of the traditional PV systems lies in the relatively lower
electrical efficiency, which is in the range 10–20% [3]. Furthermore,
PVs’ electrical efficiency varies in an inversely linear trend with the
PV cells’ surface temperature, leading to around 0.5% efficiency
declining per degree rise in the cells’ temperature [4]. The PV/ther-
mal (PV/T) technology was therefore developed to control the tem-
perature of the PV cells and make advanced utilisation of the heat
trapped within the PVs simultaneously.
Technologies for this purpose have been developed substan-
tially but meanwhile exhibited by some inherent problems [5].
The most common way to cool the PV cells/modules is the one uti-
lising the naturally/mechanically ventilated air [6–10]. This
method has poor heat removal effectiveness due to the low ther-
mal mass of the air. Alternatively, PV modules could be cooled by
using the loop circulated water that runs through the backside
coils of the PVs [11–15]. This approach has also very limited
improvement in efficiency owing to a higher temperature rise of
the water within the loop and potential piping freezing occurring
in cold climatic regions. Another PV cooling approach was to place
the direct expansion evaporation coils beneath the PV module that
allow a refrigerant to pass through [16–20]. As a result, the PV cells
could be cooled to a very lower temperature, leading to a greater
increase in the PVs’ electrical efficiency. This approach, regarded
as the significant step-forward in the PVs cooling technology,
was also identified with several practical challenges: e.g., multiple
welding joints making a complex manufacturing process, high cost
by using the copper coils, and uneven refrigerant distribution
across the multiple coils in a large area [5]. In recent years, heat
pipes were applied to cool the PV cells/modules by extracting the
heat trapped within the PVs and delivering it to the passing fluid
employing the self-driven evaporation & condensation cycling of
the heat pipe working fluid [21,22]. However, traditional design
relating to this concept is to use multiple heat pipes as the thermal
absorbers, which have the relatively complex structure and higher
cost. Furthermore, the fluid passing across the heat-pipe condenser
is usually water, which has the gradually growing temperature
along the flow path thus leading to reduced heat transfer rate
between the heat pipe working fluid and passing water. It is there-
fore essential to find a solution to enhance the overall heat-transfer
efficiency of the PV/T system and meanwhile, to simplify its struc-
ture thus enabling reduced cost and broader service applications.
To remove the above addressed technical barriers remaining
with the existing heat pipe based PV/T technology, loop heat pipes
(LHP) have been introduced into the PV/T systems. A LHP is a spe-
cial type of heat pipe that combines the principles of thermal con-
ductivity and phase transition. It has a large heat transport
capacity enabling heat to be transferred along a relatively long dis-
tance by circulating the working fluid in a closed loop. The LHP has
also some particular features that makes it appropriate to cooling
of the PV cells/modules, e.g., effective heat transfer using the least
pipings, use of the anti-freezing medium within the loop fluid, her-
metically sealed loops and homogeneous capillary force [23].It
was understood that the LHPs have been widely used in thermal
controls of satellites, spacecrafts, electronics, and LEDs [23,24].
However, use of LHPs in solar collecting systems was not often
reported, with a couple of cases concerning application of the grav-
itational LHP [23]. It has also been noted that the gravitational
LHPs have the ‘dry-out’ problem caused by limited capillary effect
of the wicks within the heat pipe evaporation section, resulting in
significantly reduced heat transfer capacity.
To tackle the above addressed ‘dry-out’ problem, a novel LHP
structure comprising the top-positioned vapour–liquid separator
was developed by the authors and this new type of LHP solar ther-
mal systems have been studied using both simulation and experi-
mental methods, resulting in the dedicated parametrical
characterisation of the specific LHP and associated thermal and
power systems [25–29]. On basis of the authors’ previous research
achievement on this topic, the social economic issues of the system
for use in three different climatic regions, namely, cold area – rep-
resented by London (0.1°W, 51.3°N), warm area – represented by
Shanghai (121.8°E, 31.2°N), and hot area (subtropical) – repre-
sented by Hong Kong (114.2°E, 22.2°N), were studied in the paper.
This study involved prediction of the potential fossil fuel energy
Nomenclature
cspecific heat capacity (J/kg K)
Ccost ()
CS cost saving ()
CR carbon reduction (kg)
eroot mean square deviation
ffactor
Mmass (kg)
PP pay-back period (year)
Qenergy rate (W)
rcorrelation coefficient
ttime
Ttemperature (K)
xwidth parameter of fin sheet
X
e
experimental results
X
s
simulation results
Greek
g
energy efficiency
Subscript
au auxiliary energy
c,PV/LHP capital of PV/LHP system
eelectricity
el-CO
2
CO
2
emission of electric heater
gas-CO
2
CO
2
emission of gas boiler
LHP loop heat pipe
m,PV/LHP maintenance of PV/LHP system
m,wh maintenance of water heating system
RE renewable incentives
th thermal
ooverall
o,PV/LHP operation of PV/LHP system
o,wh operation of water heating system
wwater
wh,CO
2
CO
2
emission of water heating system
w,load water heating load
X. Zhang et al. / Applied Energy 135 (2014) 20–34 21
saving, return time on investment and carbon emission reduction
of the new system relative to the traditional gas-fired and electrical
boilers based water heating systems. The research results will be
able to assist in decision making in implementation of the pro-
posed PV/T technology and analyses of the associated economic
and environmental benefits, thus contributing to realisation of
regional and global targets on fossil-fuel energy saving and envi-
ronmental sustainability.
2. System description
Fig. 1 illustrates the target PV/LHP heat pump water heating
system, which comprises a PV/LHP module, the vapour/liquid
transportation lines, a flat-plate heat exchanger (LHP condenser
and heat pump evaporator), an electric compressor, a coil con-
denser immersed into a water tank, an expansion valve and an
electric control & storage unit (controller, inverter and battery).
The PV/LHP module is the key component of the system that con-
sists of a single tempered glazing, a PV lamination layer, the LHP
evaporator, a fin sheet and a polystyrene board for thermal insula-
tion. These layers are positioned into an aluminium-alloy casing
using the clamping and argon-arc welding methods. The LHP evap-
orator plays an essential role in transferring the heat from the PV
layers to the flat-plate heat exchanger. It is placed into an alumin-
ium
X
-shape fin sheet and then attached to the rear surface of PV
layer using thermal-conductive silicon sealants. Detailed descrip-
tions about the LHP evaporator structure, the system working prin-
ciple and the system design parameters have been fully addressed
in previous work of the authors [27,28].
Through dedicated computer numerical simulation and experi-
mental testing, the operational performance of such a novel LHP
and associated thermal and power systems has been characterised
[25–29], giving the following conclusions respectively: (1) heat
transport capacity of the new LHP was around 1.08 W/cm
2
(900 W in total), nearly 86% higher than that for common heat
pipes operated in gravitational field [30]; (2) solar thermal effi-
ciency of the new LHP based solar thermal facade system was
about 48.8%, over 18% higher than that for the conventional solar
thermal system [31]; (3) solar thermal and electrical efficiencies
of the new LHP based PV/T system were 39.25% and 9.13% respec-
tively, resulting in 15.02% exergetic efficiency, nearly 24% higher
than that for the common PV/T systems [32] and (4) Coefficient
of performance (COP) of the new LHP based solar heat pump sys-
tem was 5.51, about 1.5 times of conventional solar heat pump sys-
tems [33]. Performance characterisation results of the new LHP and
associated thermal and power systems are summarised in Fig. 2.
Such PV/LHP heat pump system may have several distinct char-
acteristics: (1) configuration of the PV/LHP module is simplified
with only one LHP, leading to a corresponding low cost; (2) the
LHP can passively transport heat for a long distance that removes
the need for a circulation pump; (3) the heat pump controls the
PV cells in a relatively low-temperature operation mode; (4)
the generated PV electricity can offset part/full power load for driv-
ing the compressor, thus creating a low/zero-carbon water heating
operation; and (5) this system can be either installed on a building
by mounting the PV/LHP module onto the facade and connecting PV
electricity to the national grid or installed as an independent heat
and power cogeneration unit to meet the energy load.
Fig. 1. Schematic of the solar PV/LHP heat pump water heating system.
22 X. Zhang et al. / Applied Energy 135 (2014) 20–34
3. Social economic performance analyses – simulation model
development
3.1. Energy-performance prediction model
The dynamic modelling had the aim of evaluating the perfor-
mance of the integrated PV/LHP heat pump system in real climatic
operational conditions, which are simultaneously affected by sev-
eral critical factors, i.e., solar radiation, air temperature, air velocity
and operating time. This modelling enabled (a) a prediction of sys-
tem performance in real climatic operational conditions; (b) a fore-
cast of seasonal system performance; (c) a recommendation for an
appropriate climate region suitable for the operation of such a PV/
LHP heat pump system; and (d) analyse the system’s economic and
environmental benefits.
For the PV/LHP heat pump water heating system, the transient
operational model involved six energy balance equations: (1) a
heat balance equation for the glazing cover; (2) a heat balance
equation for the PV layer; (3) an one-dimensional unsteady-state
heat conductance of the fin sheet; (4) a heat balance equation for
the LHP operation; (5) heat balance equations for the heat pump
evaporator and (6) the water tank [34–36]. The full mathematical
descriptions of the energy-balance equations were presented in
authors’ previous work [28]. To resolve the equation system using
a numerical method, the differential equations can be rewritten in
the formats of the MATLAB’s ‘‘ode15s’’ and ‘‘pdepe’’ solvers using an
implicit backward difference formula and the finite element meth-
ods (FEM) and [37]. The MATLAB ‘‘pdepe’’ solver is normally
applied for the initial-boundary value problems to solve the para-
bolic and elliptic partial differential equation (PDE) systems in one
space variable ‘‘x’’ and time ‘‘t’’. This solver converts the PDE into
ordinary differential equation (ODE) using a second-order accurate
spatial discretization based on a set of nodes, which is recognised
as a classical FEM method. The time integration is completed with
the ‘‘ode15s’’ solver, which is a variable order solver to resolve the
algebraic equations according to the backward differentiation for-
mulas. Such numerical method is based upon the built-in subpro-
grams in the MATLAB software, which consumed much less
computing time than the difference method in previous work [28].
3.2. Operational cost saving and payback-time prediction model
Prior to evaluating the socio-economic benefits, the annual hot
water demand, Q
w,load
, should be calculated as the baseline of
energy requirement for the comparison between such system
and traditional water heaters. The initial water temperature in
the 35-L water tank was assumed to be the same as the ground
water temperature. The eventual hot water temperature criteria
are all considered at 45 °C. It can be predicted that more or less
the energy deficiency of such PV/LHP heat pump system may
occur, which could be matched through an auxiliary gas boiler or
electric heater as the backup.
Q
w;load
¼M
w
c
w
D
T
w
ð1Þ
The cost payback period for operating such a prototype system
to replace conventional water heaters can be estimated by [38]
PP
PV=LHP
¼CapitalCost Incentives
AnnualðOperational &maintenanceÞCostSaving
¼C
c;PV=LHP
C
RE
ðC
o;wh
þC
m;wh
ÞðC
o;PV=LHP
þC
m;PV=LHP
Þð2Þ
Fig. 2. Performance characterisation results of the new LHP and associated thermal and power systems.
X. Zhang et al. / Applied Energy 135 (2014) 20–34 23
To install a new solar water heating system, it might be possible to
receive grants through the government’s renewable policy, such as
the Renewable Heat Incentive (RHI) scheme in London, which is
intended to encourage the uptake of renewable heating technolo-
gies within households, communities and businesses through the
provision of financial incentives. The financial support available
for installing such a solar thermal system are ‘‘0.24/kW h yr heat
(7 years)’’ for London and ‘‘reduction in 13% of capital cost’’ for Shang-
hai while there is no renewable tariffs found in Hong Kong right
now [39,40]. The final cost of the installation of such a solar proto-
type system equals the value achieved by subtracting the local
incentive amounts for renewable projects from the capital cost.
The maintenance cost of a solar heating system is normally esti-
mated at 2% of the initial system cost [38] due to its low mainte-
nance requirement. An electric water heater is considered free in
terms of its maintenance during its life cycle [41].
As a solar photovoltaic system is usually considered to have a
life span of 25 years [42], the life-cycle net cost saving, CS
PV/LHP
,
of this solar system in energy bills can be determined by
CS
PV=LHP
¼ðLifetime paybacktimeÞ
AnnualðOperational &maintenanceÞCostSaving
¼ð25 PP
PV=LHP
Þ½ðC
o;wh
þC
m;wh
ÞðC
o;PV=LHP
þC
m;PV=LHP
Þ ð3Þ
3.3. Environmental benefit prediction model – life cycle carbon
emission
Environmental benefits can be simply estimated by using the
annual CO
2
emission factor when operating this PV/LHP heat pump
system (including the required auxiliary energy and the net system
electricity consumption) to replace a conventional water heater
[38].
CR
PV=LHP
¼f
wh;CO
2
Q
w;load
f
gas-CO
2
Q
au;PV=LHP
f
el-CO
2
Q
au;PV=LHP
ð4Þ
The gas-to-CO
2
emission factor is estimated to be the same at
0.26 kg CO
2
/kW h heat for the three regions, as gas is directly
burned for heat generation [43]. While the electricity-to-CO
2
emis-
sion factor of should be different for the three regions due to the dif-
ferent efficiencies of the national power plants, which are 0.545,
0.997 and 0.840 kg CO
2
/kW h heat respectively in London, Shanghai
and Hong Kong [44–46].
3.4. Module and system performance evaluation
In this paper, performance of the PV/LHP module was examined
by both the energetic and exergetic efficiencies while the perfor-
mance of the whole system was assessed by the advanced thermal
performance coefficients (COP
PV/T
). All the mathematical descrip-
tions of these evaluation parameters could be found in authors’
previous work [28].
3.5. Simulation model operation
The corresponding initial and boundary conditions, i.e., solar
radiation, air temperature, wind speed and water temperature,
were extracted from the weather database of Energy-Plus software,
which are respectively the ‘037760_IWEC’ for London, the
‘583670_IWEC’ for Shanghai, and the ‘450070_CityUHK’ for Hong
Kong [47]. The monthly diurnal averages for solar radiation and
air temperature for these regions are respectively shown in Figs. 3
and 4.Table 1 gives the monthly average weighted wind speed for
the three climatic regions. It is seen that Hong Kong has a medium
level of solar radiation, is hot in summer and warm in winter;
Shanghai also has a medium level of solar radiation but is hot in
summer and cold in winter; while London has a low level of solar
radiation, is warm in summer and cold in winter. During the sim-
ulation, it was assumed that the system starts operation from
08:00 and ends its operation at 16:00 for a single day. The heat
pump was considered to operate at 0 °C/55 °C in winter, 10 °C/
55 °C in summer, and 5 °C/55 °C in spring and autumn. The PV/
LHP panel installation angle was set to the same level as the local
altitude in the three selected regions. The initial temperature of the
water stored in the tank was considered to be the ground water
temperature at a height of 0.5 m below ground level, as in Table 2.
An initial temperature distribution (T= 0) from the module
cover to the water in the tank is desired before starting the itera-
tion. The time step size,
D
t, and the space step size,
D
x, were,
respectively, given at 1 min and 22 mm because a smaller step
would consume much more time to calculate and would not affect
the results very much, while a larger time step would lead to an
unstable calculation process [31]. As the temperature profile of
the fin sheet is assumed to be symmetrical at two sides of the
LHP evaporator in the centre, there are then only considered to
be 11 elements from the left-hand edge to the fin centre along
the fin width. To overcome the potential energy deficiency of such
PV/LHP heat pump system, an auxiliary gas boiler or electric heater
was applied as the system backup. The socio-economic figures,
such as capital cost, renewable tariffs, system life span and gas/
electricity-to-CO
2
emission factors, were also input into the pro-
gram for simulation. The algorithm is presented in a flow diagram
in Fig. 5, which is also interpreted as follows:
(1) Assign the design, operating parameters and socio-economic
figures into program code.
(2) Input the external boundary conditions from the weather
data file.
(3) Assume the initial parameters’ values: temperature distribu-
tion and mass flow rate.
(4) Set up the time step size,
D
T, and space step size,
D
x, and
start the calculation.
(5) Carrying out heat analysis on each module/system compo-
nent and rewriting them in the required format of the
‘‘ode15s’’ solver.
(6) Rewrite the Eq. (3) in the required format of the ‘‘pdepe’’
solver.
(7) Input the boundary condition of the fin sheet into the PDE
and analyse the transient heat conductance on the fin sheet.
(8) Ensure the PDE results’ accuracy achieve the criteria of 10
3
.
(9) Calculate the hot water load by Eq. (1) based on the initial
weather data.
(10) Estimate the annual energy saving and required auxiliary
energy by Eq. (2) based on the operational results.
(11) Identify the life-cycle cost and carbon savings by Eqs. (3) and
(4).
(12) Carry out the energetic, exergetic and system performance
calculation.
(13) Complete the operation of the program until time end and
export the results.
(14) Program stops.
4. Validation of the simulation model
4.1. Validation by using the published data
The simulation model was initially validated for its suitability
and accuracy by comparing the modelling results with the pub-
lished experimental data of a PV/solar-assisted heat pump/heat
pipe (PV–SAHP/HP) system developed by Fu et al. [47]. They
attached groups of PV cells to a flat-plate heat-pipe thermal
24 X. Zhang et al. / Applied Energy 135 (2014) 20–34
collector and a direct-expansion evaporator in a heat pump for hot
water generation. There were three operating modes during their
testing, and the heat pipe operating mode with the greatest simi-
larity to the proposed PV/LHP system was selected to validate
the simulation model. The system performance indicators were
defined congruously according to the mathematical descriptions
in this section. The correlation coefficient (r) and the root mean
square percentage deviation (e) defined in below equations were
Fig. 3. Monthly averages for solar radiation in three climate regions.
Fig. 4. Monthly averages for air temperature in three climate regions.
Table 1
Monthly averages for wind speed in three climate regions (m/s).
Location January February March April May June July August September October November December
London 3.4 2.7 4.6 4.2 2.5 3.2 2.6 3.0 3.1 2.4 2.3 3.2
Shanghai 3.5 3.5 3.5 3.5 3.9 3.7 3.3 3.4 3.6 3.7 3.9 3.8
Hong Kong 3.1 3.7 3.4 3.5 2.8 3.5 3.3 3.3 4.2 3.5 3.3 3.1
Table 2
Monthly averages for ground water temperature in three climate regions (°C).
Location January February March April May June July August September October November December
London 4.2 5.3 7.5 9.6 13.6 15.7 16.2 15.2 12.7 9.7 6.7 4.7
Shanghai 5.5 7.5 11.4 15.2 22.3 26.0 27.0 25.1 20.8 15.3 10.0 6.4
Hong Kong 17.3 18.4 20.5 22.5 26.3 28.3 28.8 27.8 25.5 22.6 19.7 17.8
X. Zhang et al. / Applied Energy 135 (2014) 20–34 25
applied to analyse the difference between the theoretical and the
published experimental results.
r¼n
R
X
e
X
s
ð
R
X
e
Þð
R
X
s
Þ
ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
n
R
X
2
e
ð
R
X
e
Þ
2
q ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
n
R
X
2
s
ð
R
X
s
Þ
2
qð5Þ
e¼ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
P½100 ðX
e
X
s
Þ=X
e
2
n
sð6Þ
where nis the number of experiments implemented; and X
e
and X
s
represent the experimental and simulation results, respectively.
Fig. 5. Flow chart for the modelling set-up.
26 X. Zhang et al. / Applied Energy 135 (2014) 20–34
Fig. 6 presents the transient simulation results of the module
outputs against operating time by inputting the design, operation
and weather conditions of the referenced PV–SAHP/HP experimen-
tal rig [48]. The original experimental data in Fig. 6 was derived
from the reference [48], which was further compared with the sim-
ulation results in this paper. The correlation coefficients (r) and the
root mean square percentage deviation (e) for the modelling and
testing of the module electrical/thermal outputs were 0.9938/
0.9973 and 4.95%/12.87%, respectively. The reason for errors may
exist in the utilisation of simplified assumptions/empirical formu-
las, and an inaccurate estimation of the heat loss coefficient due to
a lack of wind data. However, the accuracy achieved by this
dynamic simulation model was acceptable from the engineering
point of view, and could therefore be applied to evaluate the sys-
tem performance in the real climates and for recommending
appropriate regions for operation.
4.2. Validation by using the experimental results
In addition, the accuracy of this simulation model was also ver-
ified using the dedicated experiments outlined in authors’ previous
work [28]. Through a parallel comparison between the modelling
and real-time test results, the established simulation model was
validated with a reasonable accuracy of mean error less than 9%.
Owing to the good level of agreement achieved, this simulation
model is appropriate for predicting the annual operational perfor-
mance of the PV/LHP heat pump system and recommending
regions for operation.
5. Results and discussion
5.1. Energy-performance prediction results
In order to establish which climate best suits the system, this
section investigates the annual operational performances of the
prototype system in three climate regions: London, Shanghai and
Hong Kong. Fig. 7 presents the monthly average temperatures of
the PV layer in the three regions. It was observed that the PV tem-
perature had a similar trend of variation to the solar radiation and
air temperature, which achieved a maximum figure in summer and
a minimum figure in winter. In London, the monthly PV tempera-
ture was in the range from 7.57 °C (in February) to 35.76 °C (in
July). In Shanghai, the monthly PV temperature was in the range
from 15.93 °C (in January) and 47.12 °C (in August). In Hong Kong,
the PV temperature was in the range from 25.23 °C (in February) to
43.41 °C (in July). The PV temperature was the lowest in London
and the highest in Hong Kong throughout the year. The annual
average PV temperatures in London, Shanghai and Hong Kong were
18.97 °C, 31.45 °C and 34.66 °C, respectively.
The monthly electrical efficiencies (
g
e
) of the PV/LHP module
varied inversely to its temperature, as shown in Fig. 8a, which were
in the range from 8.42% to 9.35% in London, 8.02% to 9.08% in
Fig. 6. Comparison of the simulation results with the published testing data.
Fig. 7. Temperatures of the PV layer in three climate regions.
X. Zhang et al. / Applied Energy 135 (2014) 20–34 27
Shanghai, and 8.16% to 8.77% in Hong Kong. The annual mean elec-
trical efficiency was highest at 8.94% in London, while it was rela-
tively lower in Shanghai (8.57%) and Hong Kong (8.42%). However,
the module’s thermal efficiency (
g
th
) varied in the opposite manner
to its electrical efficiency, as displayed in Fig. 8b. The monthly ther-
mal efficiency of the PV/LHP module was in the range from 14.13%
to 34.63% in London, 18.06% to 61.41% in Shanghai, and 43.13% to
59.86% in Hong Kong. Operation of the prototype system in Hong
Kong was found to be the most stable, with the highest annual
average thermal efficiency of the module at about 51.65% owing
to the warmest air temperatures existing in this area. In Shanghai
and London, the annual average thermal efficiencies of the module
were much lower than in Hong Kong at around 38.99% and 26.99%,
respectively. After adding the electrical and thermal efficiencies
together, the overall energetic efficiencies (
g
o
) of the module in
Fig. 8c varied as a similar way to the thermal efficiency, which
had annual average values of 35.93%, 47.57% and 60.06% in London,
Shanghai and Hong Kong, respectively.
Fig. 9 illustrates the overall energy output of the module. The
quantity of the energy output was found primarily to depend on
Fig. 8. (a) Electrical energetic efficiencies, (b) thermal energetic efficiency and (c) overall energetic efficiency of the PV/LHP module in three climate regions.
28 X. Zhang et al. / Applied Energy 135 (2014) 20–34
Fig. 9. Overall energy output of the PV/LHP module in three climate regions.
Fig. 10. Condensation capacity of the heat pump in three climate regions.
Fig. 11. Overall assessment of the prototype system in three climate regions.
X. Zhang et al. / Applied Energy 135 (2014) 20–34 29
the amount of available solar radiation and the surrounding air
temperature in the three regions. The module electricity genera-
tion in London varied from 1.42 kW h/m
2
to 10.70 kW h/m
2
, with
an average value of 5.43 kW h/m
2
. The monthly electrical output
range in Shanghai was from 8.48 kW h/m
2
to 13.15 kW h/m
2
, with
an average performance of 10.76 kW h/m
2
. In Hong Kong, the sys-
tem generated electricity in the range of 7.49 kW h/m
2
to
12.56 kW h/m
2
, with an average value of 10.06 kW h/m
2
. The ther-
mal output ranges of the module were 4.07–43.99 kW h/m
2
,
19.51–89.77 kW h/m
2
and 41.04–79.50 kW h/m
2
, respectively in
London, Shanghai and Hong Kong, and their corresponding
monthly average values were 16.44 kW h/m
2
, 51.40 kW h/m
2
and
61.82 kW h/m
2
. According to these figures, Hong Kong was found
to have the highest energy output for the module, including an
almost equivalent electricity output to Shanghai and the highest
monthly average heat generation for the three regions.
By adding together the heat pump electrical consumption and
the heat output of the module (heat source), the monthly condensa-
tion heat outputs (heat sink) of the integrated PV/LHP heat pump
system are presented in Fig. 10, giving a figure in the range
from 5.37 kW h/m
2
to 54.96 kW h/m
2
in London, 25.75–
112.15 kW h/m
2
in Shanghai, and 52.54–102.28 kW h/m
2
in Hong
Kong, respectively, and the corresponding monthly average figures
are 20.89 kW h/m
2
in London, 65.59 kW h/m
2
in Shanghai, and
79.33 kW h/m
2
in Hong Kong.
An overall assessment of the PV/LHP module and the associated
heat pump system is presented in Fig. 11. From the module point of
view, the exergetic efficiency was relatively stable for all three
regions. Owing to the highest thermal output, the annual average
exergetic efficiency of the PV/LHP module was highest in Hong
Kong at 13.08%, followed by 12.35% in Shanghai, and 11.85% in
London. From the integrated system point of view, the annual aver-
age COP
PV/T
value was highest at 9.44 in London because of the
largest ratio of electricity to heat output, while relatively lower
COP
PV/T
values were found in Shanghai (8.35) and Hong Kong
(6.97). It needs to be addressed that as the evaporation and con-
densation temperatures of the heat pump operation were set up
at the same level during the simulation in all three regions, the sea-
sonal system COP
th
value will be the same as that of Shanghai at
4.57, 5.10 and 5.75, respectively, in the winter, transit, and summer
seasons.
Table 3 presents the monthly eventual water temperature after
a single day’s operation for the three climatic regions. The area
shaded grey in the table indicates the months that the temperature
of the tank water could not be heated above the hot water temper-
ature criterion of 45 °C by the prototype system alone. This system
could provide hot water service for nearly seven months per year
in Shanghai. However, there was only one month in London (July)
in which this system could achieve the water temperature crite-
rion. In Hong Kong, this system could reach the required water
temperature throughout whole year.
Table 4 gives the total annual operational output of the proto-
type system in the three areas. Hong Kong was found to have the
highest thermal energy output from the module at 741.85 kW h/
m
2
yr, which was then upgraded to 921.70 kW h/m
2
yr by input-
ting electricity at 179.85 kW h/m
2
yr into the heat pump compres-
sor. After subtracting the solar electricity generation of nearly
120.74 kW h/m
2
yr, this system would require additional grid elec-
tricity of 59.11 kW h/m
2
yr to meet the heat pump consumption in
Hong Kong, and would not need any other auxiliary heater in the
system to achieve the hot water demand. Owing to the highest
solar radiation level and conspicuous seasonal air temperature in
Shanghai, the prototype module could generate the most electric-
ity at 129.14 kW h/m
2
yr but have a lower heat output than the
operation in Hong Kong at 616.78 kW h/m
2
yr. And in Shanghai,
additional grid electricity of 16.15 kW h/m
2
yr will be required to
meet the heat pump consumption, further delivering the conden-
sation heat at 762.07 kW h/m
2
yr. In London, this module produced
the lowest energy quantity for both electricity (65.11 kW h/m
2
yr)
and heat (197.23 kW h/m
2
yr), due to the lowest level of solar radi-
ation and the coldest air temperature for the three regions.
Although this prototype system produced a net amount of electric-
ity (19.64 kW h/m
2
yr) in London, it would consume much more
energy from the back-up heater to meet the hot water demand.
The output of condensation heat was also the least at only
242.70 kW h/m
2
yr in London. The simulation results offer the
interpretation that this prototype system would perform best in
a subtropical climate, such as the Hong Kong area.
5.2. Bill saving and return time on investment prediction
5.2.1. Capital cost
5.2.1.1. Capital cost of the prototype PV/LHP heat pump system. The
capital cost of the prototype PV/LHP heat pump system was esti-
mated by adding together the individual prices of all the system
components and taking into account appropriate commercial prof-
its. Table 5 provides a list of cost breakdowns and indicates that the
initial cost of such a system is 512.69. Furthermore, the cost
details of the different system components are presented in
Fig. 12. The PV/LHP module was the most expensive of the system
components, accounting for nearly 44% of the total system cost.
It needs to be addressed that this system was a second proto-
type whose capital cost was much less than that of the previous
one [29], mainly owing to the fast reduction of PV price recently.
5.2.1.2. Renewable Earning (RE). The financial support available for
installing such a solar thermal system in the three regions is given
in Table 6 using figures extracted from Tables 4 and 5. The final
cost of the installation of such a solar prototype system equals
the value achieved by subtracting the local incentive amounts for
renewable projects from the capital cost.
Table 3
Monthly final tank water temperature in three climate regions.
Table 4
Total annual output of the prototype system in three climate regions.
Location London Shanghai Hong
Kong
Solar radiation (kW h/m
2
yr) 737.35 1515.78 1439.26
Solar electricity generation (kW h/m
2
yr) 65.11 129.14 120.74
Solar heat output of module (kW h/m
2
yr) 197.23 616.78 741.85
Heat pump work consumption
(kW h/m
2
yr)
45.47 145.29 179.85
Heat pump condensation heat
(kW h/m
2
yr)
242.70 762.07 921.70
System net electricity output
(kW h/m
2
yr)
19.64 16.15 59.11
30 X. Zhang et al. / Applied Energy 135 (2014) 20–34
5.2.2. Annual operational cost and saving
In the potential case of the unsatisfactory operation of the pro-
totype system in certain terrible weather conditions, an auxiliary
heater could be switched on to heat up the water until its temper-
ature achieves the expected 45 °C. A gas boiler and electric heater
are, respectively, considered as the auxiliary heater of the proto-
type system when comparing it with a conventional gas boiler
(efficiency of 80%) and a typical electric heater (efficiency of 90%)
[43]. The annual operating cost of these systems can be estimated
as shown in Table 7 and Fig. 13. It needs to be noted that the total
heat (required only to be 45 °C here) produced by the prototype
system was estimated to be less than the amount of heat pump
condensation heat in Table 4, because the water was heated up
to more than 45 °C in a number of circumstances. The combination
of the prototype system with an auxiliary gas boiler was the most
economical and, when replacing conventional gas/electric water
heaters, such a combined operation could save nearly 13.00/
73.44, 11.16/32.61, and 71.46/49.25 per year in the regions of
London, Shanghai and Hong Kong, resulting in an annual cost sav-
ing ratio of 34.43%/74.79%, 72.03%/88.27%, and 92.95%/90.08%,
respectively.
5.2.3. Annual maintenance cost
The maintenance cost of a solar heating system is normally esti-
mated at 2% of the initial system cost due to its low maintenance
requirement [38]. The maintenance cost of a gas boiler was esti-
mated at 31.25/yr, 10/yr and 12.50/yr, respectively, in the
regions of London, Shanghai and Hong Kong [48,49]. An electric
water heater was considered free in terms of its maintenance dur-
ing its life cycle [41].
5.2.4. Cost payback period and life-cycle net cost saving
Table 8 gives the estimated payback period and life-cycle net
cost saving of such a prototype PV/LHP heat pump system when
supported by a conventional gas boiler. When replacing a typical
gas heater, this system has the shortest cost payback period of
up to 7 years and the highest life-cycle net cost saving of nearly
2174 per m
2
in Hong Kong. In London, the cost payback period
will be around 8 years with a life-cycle net cost saving of about
985 per m
2
after considering the renewable award of installing
a new solar thermal system. About 41 years (more than its life
span) would be required in Shanghai to reclaim the initial invest-
ment due to the current lowest gas charging tariff, which indicates
that it would be uneconomical to replace a gas water heater with
the proposed PV/LHP heat pump system in this area at the
moment. When replacing a conventional electric water heater,
the system’s payback periods were estimated at nearly 5, 20 and
14 years, respectively, in London, Shanghai and Hong Kong owing
to the different electricity charging tariffs and governmental sup-
port policies. The net cost saving would be around 2151, 184
and 756 per m
2
accordingly throughout the system life span.
Table 5
Capital cost breakdown of the prototype PV/LHP heat pump system.
No. Component Quantity/size Unit price () Cost ()
PV/LHP module
1 PV layer 1 [89 Wp] 106.25 106.25
2 Loop heat pipe 3.2 [m] 20.00 20.00
3 Treated Al-alloy sheet 1 [piece] 18.75 18.75
4 Tempered glazing 1 [piece] 7.50 7.50
5 Aluminium frame 1 [piece] 8.75 8.75
6 Aluminium fin sheet 1 [piece] 13.75 13.75
Other components
7 Flat-plate heat exchanger 1 [1 kW] 18.75 18.75
8 Heat pump compressor 1 [1 HP] 125.00 125.00
9 Expansion valve 1 [piece] 6.25 6.25
10 Refrigerant 1 [300 g] 7.50 7.50
11 Water tank 1 [35 L] 6.88 6.88
12 Solar controller 1 [12V10 A] 10.00 10.00
13 Electric wire 1 [coil] 5.00 5.00
14 Battery 1[12V100 AH] 18.75 18.75
Other accessories
15 Thermal insulation 4 [piece] 7.50 7.50
16 Module bracket 1 [piece] 10.00 10.00
17 Silicon seal 2 [bottle] 3.75 3.75
Subtotal
Total system fabrication cost () 394.38
Additional profit (30% of total fabrication cost) 118.31
Capital cost () 512.69
Fig. 12. Cost breakdown of the PV/LHP heat pump prototype system.
Table 6
Renewables tariffs for the installation of a solar thermal system.
Location London Shanghai Hong
Kong
Tariff 0.24/kW h yr
heat [39]
Reduction in 13% of
capital cost [40]
Not found
Years 7 1 N/A
Total earning () 249.54 66.25 N/A
X. Zhang et al. / Applied Energy 135 (2014) 20–34 31
The analytical results in Table 8 demonstrate that it would be
most cost-effective for such a PV/LHP heat pump system with a
backup gas boiler to replace a gas water heater in Hong Kong
and an electric water heater in London. It would be very uneco-
nomical for the PV/LHP heat pump system to replace either type
of conventional water heater in the Shanghai area.
5.3. Environmental benefit prediction – life cycle carbon emission
reduction
The gas-to-CO
2
emission factor was estimated to be the same
for the three regions, as gas is directly burned for heat generation.
The CO
2
emission factor of electricity was considered different for
Table 7
Annual operating costs of different water heating systems.
Location London Shanghai Hong Kong
Energy price and demand
Gas price (/kW h) [50–52] 0.06 0.03 0.19
Electricity price (/kW h) [53–55] 0.18 0.08 0.15
Feed-in tariff (/kW h) [39,40] 0.19 0.13 0.18
Total heat demand for hot water (kW h/yr) 519.91 431.13 328.04
PV/LHP heat pump system operational performance
Heat produced from system (kW h/yr) 148.02 317.58 328.04
Energy required from auxiliary heater (kW h/yr) 371.89 113.55 0.00
Total electricity output from module (kW h/yr) 39.85 79.03 73.89
Electricity consumed by heat pump (kW h/yr) 27.84 82.30 110.04
Net electricity output (kW h/yr) 12.01 3.27 36.15
Gas water heater (efficiency 80%)
Required gas energy (kW h/yr) 649.88 538.91 410.05
Operational cost (/yr) 37.50 15.00 77.50
Electric water heater (efficiency 90%)
Required electricity energy (kW h/yr) 577.68 479.03 364.49
Operational cost (/yr) 98.75 37.50 55.00
PV/LHP heat pump system + auxiliary gas boiler
Gas required from auxiliary gas boiler (kW h/yr) 464.86 141.94 0.00
Operational cost (/yr) 24.76 4.34 5.43
Annual saving by replacing gas boiler (/yr) 13.00 11.16 71.46
Annual saving ratio by replacing gas boiler (%) 34.43 72.03 92.95
Annual saving by replacing electric heater (/yr) 73.44 32.61 49.25
Annual saving ratio by replacing electric heater (%) 74.79 88.27 90.08
PV/LHP heat pump system + auxiliary electric heater
Electricity required from auxiliary electric heater (kW h/yr) 413.21 126.17 0.00
Operational cost (/yr) 67.99 9.99 5.43
Annual saving by replacing gas boiler (/yr) 30.23 5.51 71.46
Annual saving ratio by replacing gas boiler (%) 0.00 35.57 92.95
Annual saving by replacing electric heater (/yr) 30.21 26.96 49.25
Annual saving ratio by replacing electric heater (%) 30.76 72.98 90.08
Fig. 13. Annual operating costs of different water heating systems.
32 X. Zhang et al. / Applied Energy 135 (2014) 20–34
the three regions due to the different efficiencies of the national
power plants, given in Table 9.
Shanghai was found to have the highest life-cycle CO
2
emission
saving of about 4.08 ton/m
2
and 17.87 ton/m
2
when replacing gas
and electric water heaters with the prototype system respectively,
which is mainly owing to the lowest efficiency of its national
power plant. In London and Hong Kong, the prototype system
had a relatively lower carbon emission reduction of around
1.97 ton/m
2
& 7.92 ton/m
2
and 3.11 ton/m
2
& 11.27 ton/m
2
to
replace gas and electric water heaters, respectively. The analytical
results illustrate that the maximum environmental benefits would
be achieved by operating the PV/LHP heat pump prototype system
in the Shanghai area at the present time.
6. Conclusions
This paper reported a dedicated socio-economic performance
study of a novel LHP based PV/T heat pump hot-water generation
system for application in three different climatic regions: namely,
cold area (represented by London), warm area (represented by
Shanghai), and hot area (represented by Hong Kong). This involved
the prediction of the fossil fuel energy saving, return time on
investment and life cycle carbon emission reduction of the new
system, relative to the traditional gas-fired and electrical boilers
based water heating systems.
The energy performance of the proposed system was simulated
using the established dynamic model that delivered the monthly
thermal and power performance of the system over a typical oper-
ational year. Summary of the annual heat & power outputs from the
prototype panel in three typical climatic regions were 741.85 kW h/
m
2
yr & 120.74 kW h/m
2
yr in Hong Kong, 616.78 kW h/m
2
yr &
129.14 kW h/m
2
yr in Shanghai and 197.23 kW h/m
2
yr &
65.11 kW h/m
2
yr in London, respectively. It was seen that this pro-
totype system obtained the highest heat output in Hong Kong,
which could provide sufficiently high water temperatures (above
45 °C) throughout a year. In Shanghai, the prototype system
obtained a higher volume of electricity but a lower volume of heat
compared to Hong Kong, which could provide hot water service
up to 7 months. In London, the prototype module provided the low-
est volume of electricity and heat, which could only deliver hot
water service for 1 month independently. As a result, the system,
if operated in London, would consume a higher volume of additional
energy provided by the backup heaters. From energy efficiency
point of view, the system is most applicable to the hot climatic
region.
Instead of the energy outputs, the local energy charging rates
and renewable incentives were found to be the critical factors that
impacted on the investment return time. To replace a typical gas-
fired boiler, this prototype system (with a backup gas boiler) has
the shortest cost payback period of 7 years and the highest life-
cycle net cost saving of nearly 2174 per m
2
in Hong Kong. In Lon-
don, the cost payback period would be 8 years with a life-cycle net
cost saving of about 985 per m
2
after considering the renewable
award. It was found to be uneconomical to invest in the prototype
system in Shanghai, with a payback period of 41 years (more than
its life span) due to the low gas charging tariff. To replace a conven-
tional electric water heater, the system’s payback periods were
estimated at nearly 5, 20 and 14 years, respectively, in London,
Shanghai and Hong Kong. The life-cycle net cost saving would,
accordingly, be around 2151, 184 and 756 per m
2
. From the
economic point of view, this system seems most applicable to Lon-
don or Hong Kong where either the energy charging rates are high
or the governmental financial support is positive.
Apart from the energy output, the local carbon conversion fac-
tors of gas and electricity were found to be the most critical param-
eters that affected on the environmental benefits. The system
could obtain the highest life-cycle carbon reduction volume at
4.08 tons/m
2
and 17.87 tons/m
2
of in Shanghai when using it to
replace gas and electric water heaters respectively. In London
and Hong Kong, this system would have relatively lower life-cycle
carbon emission reduction of around 1.97 ton/m
2
& 7.92 ton/m
2
and 3.11 ton/m
2
& 11.27 ton/m
2
by replacing gas and electric water
heaters with it, respectively. This phenomenon could be inter-
preted in such a way: as the higher carbon conversion factor
implies the poorer energy quality (more carbon emission volume
of per kW h energy generation), the system appeared to be more
environmentally benefiting in a place where the energy quality is
lower, e.g. Shanghai.
A question may arise from the research: which area is most
suitable for the system application? Answer to this question is
highly dependent upon the priority order among the three factors:
(1) energy efficiency, (2) economic revenue, and (3) environmental
benefit. This is thought to vary with the users, local concerns and
policy influence, etc.
The research outcomes will be able to assist in decision making
in implementation of the new PV/thermal technology and analyses
of the associated economic and environmental benefits, thus con-
tributing to realisation of the regional and global targets on fossil
fuel energy saving and environmental sustainability.
Acknowledgements
The authors would acknowledge our sincere appreciation to the
financial supports from the University of Hull, Shanghai Pacific
Energy Centre, and EU Marie Curie International Research Staff
Exchange Scheme (R-D-SBES-R-269205).
References
[1] Technology roadmap-solar photovoltaic energy. International Energy Agency;
2010. <http://www.iea-pvps.org> [accessed 11.04.11].
[2] Solar heating and cooling for a sustainable energy future in Europe (Revised).
European Solar Thermal Technology Platform (ESTTP); 2009.
[3] Zondag et al. A roadmap for the development and market introduction of PVT
technology. In: Proceeding 15th symposium thermische solarenergie, Bad
Staffelstein, Germany; 27–29 April 2005.
[4] Chow TT. A review on photovoltaic/thermal hybrid solar technology. Appl
Energy 2010;87(2):365–79.
Table 8
Cost payback period and life-cycle net cost saving.
Location London Shanghai Hong Kong
Replacing a gas water heater
Payback period (yrs) 8 41 7
Life-cycle cost saving (/m
2
) 958 2174
Replacing an electric water heater
Payback period (yrs) 5 20 14
Life-cycle cost saving (/m
2
) 2,151 184 756
Table 9
Environmental benefits of the prototype system.
Location London Shanghai Hong
Kong
Replacing a gas water heater
CO
2
emission factor (kg CO
2
/kW h heat)
[43]
0.260 0.260 0.260
CO
2
emission reduction (kg/m
2
) 78.61 163.32 124.59
Life cycle CO
2
emission reduction (ton/m
2
) 1.97 4.08 3.11
Replacing an electric water heater
CO
2
emission factor (kg CO
2
/kW h heat)
[44–46]
0.545 0.997 0.840
CO
2
emission reduction (kg/m
2
) 316.94 714.75 450.67
Life cycle CO
2
emission reduction (ton/m
2
) 7.92 17.87 11.27
X. Zhang et al. / Applied Energy 135 (2014) 20–34 33
[5] Zhang X et al. Review of R&D progress and practical application of the solar
photovoltaic/thermal (PV/T) technologies. Renew Sustain Energy Rev
2012;16(01):599–617.
[6] Tonui JK, Tripanagnostopoulos Y. Improved PV/T solar collectors with heat
extraction by forced or natural air circulation. Renew Energy
2007;32(04):623–37.
[7] Sarhaddi et al. An improved thermal and electrical model for a solar
photovoltaic thermal (PV/T) air collector. Appl Energy 2010;87(07):2328–39.
[8] Hussain et al. Design development and performance evaluation of
photovoltaic/thermal (PV/T) air base solar collector. Renew Sustain Energy
Rev 2013;25:431–41.
[9] Sukamongkol et al. Condenser heat recovery with a PV/T air heating collector
to regenerate desiccant for reducing energy use of an air conditioning room.
Energy Build 2010;42(03):315–25.
[10] Delisle Véronique, Kummert Michaël. A novel approach to compare building-
integrated photovoltaics/thermal air collectors to side-by-side PV modules
and solar thermal collectors. Sol Energy 2014;100:50–65.
[11] Huang BJ et al. Performance evaluation of solar photovoltaic/thermal systems.
Sol Energy 2001;70:443–8.
[12] Aste Niccolò, del Pero Claudio, Leonforte Fabrizio. Water flat plate PV–thermal
collectors: a review. Sol Energy 2014;102:98–115.
[13] Fraisse G et al. Energy performance of water hybrid PV/T collectors applied to
combisystems of direct solar floor type. Sol Energy 2007;81(11):1426–38.
[14] Kalogirou SA, Tripanagnostopoulos Y. Hybrid PV/T solar systems for domestic
hot water and electricity production. Energy Convers Manage
2006;47:3368–82.
[15] Herrando María et al. A UK-based assessment of hybrid PV and solar-thermal
systems for domestic heating and power: system performance. Appl Energy
2014;122:288–309.
[16] Ji et al. Performance analysis of a photovoltaic heat pump. Appl Energy
2008;85(08):680–93.
[17] Chow TT et al. Modeling and application of direct-expansion solar-assisted
heat pump for water heating in subtropical Hong Kong. Appl Energy
2010;87(02):643–9.
[18] Zhao X et al. Theoretical investigation of a novel PV/e roof module for heat
pump operation. Energy Convers Manage 2011;52:603–14.
[19] Omojaro Peter, Breitkopf Cornelia. Direct expansion solar assisted heat pumps:
a review of applications and recent research. Renew Sustain Energy Rev
2013;22:33–45.
[20] Liu et al. Performance study of a photovoltaic solar assisted heat pump with
variable-frequency compressor – a case study in Tibet. Renew Energy
2009;34(12):2680–7.
[21] Wu SY et al. A heat pipe photovoltaic/thermal (PV/T) hybrid system and its
performance evaluation. Energy Build 2011;43(12):3558–67.
[22] Gang Pei et al. Performance study and parametric analysis of a novel heat pipe
PV/T system. Energy 2012;37(01):384–95.
[23] Reay David, Kew Peter. Heat pipes: theory, design and applications. 5th
ed. Elsevier Science Ltd.; 2006.
[24] Maydanik YF. Loop heat pipes. Appl Therm Eng 2005;25:635–57.
[25] He Wei et al. Theoretical investigation of the thermal performance of a novel
solar loop-heat-pipe façade-based heat pump water heating system. Energy
Build 2014;77:180–91.
[26] Zhang X et al. Study of the heat transport capacity of a novel gravitational loop
heat pipe. Int J Low Carbon Technol 2013;8(3):210–23.
[27] Zhang X et al. Characterization of a solar photovoltaic/loop-heat-pipe heat
pump water heating system. Appl Energy 2013;102:1229–45.
[28] Zhang X et al. Dynamic performance of a novel solar photovoltaic/loop-heat-
pipe heat pump system. Appl Energy 2014;114:335–52.
[29] Zhang X et al. Design, fabrication and experimental study of a solar
photovoltaic/loop-heat-pipe based heat pump system. Sol Energy
2013;97:551–68.
[30] Riffat SB, Zhao X, Doherty PS. Analytical and numerical simulation of the
thermal performance of ‘mini’ gravitational and ‘micro’ gravitational heat
pipes. Appl Therm Eng 2002;22:1047–68.
[31] Solar water heating. <http://www.bigginhill.co.uk/solar.htm> [accessed
12.02.14].
[32] Sobhnamayan F et al. Optimization of a solar photovoltaic thermal (PV/T)
water collector based on exergy concept. Renew Energy 2014;68:356–65.
[33] Huang BJ, Chyng JP. Performance characteristics of integral type solar assisted
heat pump. Sol Energy 2001;71(06):403–14.
[34] Incropera Frank P, Lavine Adrienne S, DeWitt David P. Fundamentals of heat
and mass transfer. John Wiley & Sons; 2011.
[35] Duffle JA, Beckman WA. Solar engineering of thermal processes. 2nd ed. New
York: John Wiley and Sons Inc.; 1991.
[36] Eastop TD, Mcconkey A. Applied thermodynamics for engineering
technologists. 5th ed. Pearson Education; 1993.
[37] Jianfei Jiang, Liangjian Hu, Jian Tang. Numerical analysis and MATLAB
experiment. Science Press; 2004.
[38] Wang Zhangyuan. Investigation of a novel façade-based solar loop heat pipe
water heating system. PhD thesis. University of Nottingham; 2012. p. 131–5.
[39] Renewable Heat Incentive (RHI). <www.energysavingtrust.org.uk> [24.07.13].
[40] Solar thermal tariff in China. <www.sdpc.gov.cn> [24.07.13].
[41] Benefits of electric heating. <www.dimplex.co.uk> [26.11.13].
[42] Kannan R et al. Life cycle assessment study of solar PV systems: an example of
a 2.7 kWp distributed solar PV system in Singapore. Sol Energy
2006;80:555–63.
[43] Energy. <www.greenspec.co.uk> [24.07.13].
[44] 2010 Guidelines to Defra/DECC’s GHG conversion factors for company
reporting. <www.carbontrust.co.uk> [26.11.13].
[45] Electricity-to-CO
2
conversion ratio in China. <www.china5e.com> [26.11.13].
[46] Electricity-to-CO
2
conversion ratio in Hong Kong. <www.clp.com.hk>
[26.11.13].
[47] Fu et al. Experimental study of a photovoltaic solar-assisted heat-pump/heat-
pipe system. Appl Therm Eng 2012;40:343–50.
[48] Maintenance cost of gas boiler in UK. <www.britishgas.co.uk> [28.11.13].
[49] Maintenance cost of gas boiler in China. <www.sdpc.gov.cn> [28.11.13].
[50] UK gas price. <www.britishgas.co.uk> [26.06.13].
[51] Shanghai gas price. <www.shgas.com.cn> [26.06.13].
[52] Hong Kong gas price. <www.towngas.com> [26.06.13].
[53] UK electricity price. <www.britishgas.co.uk> [26.06.12].
[54] Shanghai electricity price. <www.sh.sgcc.com.cn> [26.06.12].
[55] Hong Kong electricity price. <www.hkelectric.com/> [26.06.12].
34 X. Zhang et al. / Applied Energy 135 (2014) 20–34
... Only about 15-20% of the received irradiance by rooftop c-Si PV panel is converted into electricity. The remaining 80-85% is lost as heat energy, which causes the working TEMP of PV panels to increase [45], and leads to the system's voltage drop [46]. The impact of temperature on PV panel production can be calculated and the result is called temperature coefficient (TEMP coef ), used to estimate the performance of the PV system during summer. ...
... The PV panels are rapidly heated to high temperatures by the unconverted 80-85% of the solar irradiance absorbed by the c-Si PV panels, as waste heat energy. High temperature triggers the system's voltage drop, PV cell degradation and the service life reduction of both solid and thin film c-Si PV modules [45,46]. Literature has it that the TEMP coef of PV panels is between 0.1-0.5% per degree Celsius. ...
... Since elevated temperature impedes the flow of current, the wind will serve as a coolant and lowers the PV panel's temperature. Wind with moderate speed lowers the panel's temperature significantly [45] and then improves the PCE of the PV cells [54]. ...
Article
Full-text available
Despite the successes recorded over the years, photovoltaic (PV) cells’ power conversion efficiency (PCE) of commercially available crystalline silicon (c-Si) PV panels still hovers between 10 and 21%. For optimal performance at 17–21% PCE, certain factors need to be understood and addressed. This study estimates the solar PV potential of selected cities across Africa, using computational modelling. The selected sites’ cities are Abuja, Addis Ababa, Kinshasa, Pretoria, and Tripoli. Sites’ coordinate systems will be exploited to generate data from meteorological databases of the selected locations needed for the PV potential assessment. This information coupled with PV system configuration will be used as inputs for PV design and simulation. The PV potential of the selected location will be extracted from the resulting simulation reports in terms of irradiance, possible power output generation, performance ratio (PR) and capacity factor (CF). The study results and analysis as extracted from the reports of the modelled hypothetical 10-kWhp c-Si rooftop PV systems at the selected sited locations, show that—Pretoria possesses the highest GTI (2234.4 kWh/m ² ) and the lowest GTI (1766.7 kWh/m ² ) was observed in Kinshasa; Pretoria has the highest PV power output (PVOUT) (17.292 MWh/), and the least (13.678 MWh) in Kinshasa; the highest PR (77.4%) was observed in Kinshasa and Pretoria and the lowest PR (76.4%) in Tripoli; Pretoria and Kinshasa recorded the highest CF (19.7%) and lowest CF (15.6%), respectively. The results indicate that the examined locations are technically viable for the PV system schemes, and therefore, massive deployment of this technology in these areas is advised.
... The ASC tool is able to provide access to accurate solar energy information mainly for urban planning, energy management and grid stability. Furthermore, due to the large cost reductions of solar panels, photovoltaic systems can be considered a profitable green investment throughout Europe, by contributing to the realization of the regional and global targets on fossil fuel energy saving and environmental sustainability [94]. The information from this climate change mitigation application will also support the sustainable urban development pilot of the EU-funded Project Eiffel [95]. ...
Article
Full-text available
The traditional Radiative Transfer Modelling solutions for Solar Energy monitoring and forecasting often provide outputs for a single point location or an area location. However, for high resolution representation of areas these solutions suffer due to low simulation speeds. This approach makes it difficult for decision-makers to accurately estimate the solar potential of the administrative area and plan installations accordingly. In this direction, the study introduces three-dimensional Ray-Tracing based radiative modeling which is a high-speed area-based solution for solar energy monitoring. The three-dimensional ray-tracing was simulated by using advanced graphic creation platforms and cloud computing in conjunction with satellite data of the clouds, aerosols, building shadows effects and three-dimensional representations of the city using Cesium 3D tiles and Unreal Engine ®. The entire system was developed in a hybrid model to be exploited by urban planners for solar PV installations and by electricity distribution system operators for energy management and efficient incorporation of the produced energy into the regular and smart grids. This study implements and analyses this Ray-Tracing model for solar photovoltaic energy potential estimation at a rooftop level for the city of Athens, Greece. The total rooftop exploitable area in Athens was found to be close to 34 km 2 , which is able to massively host distributed PVs followed by almost 4.3 TWh of annually produced energy, whilst Penteli (a Municipality in Athens) possessed a potential of 96.8 GWh with an exploitable area of just 0.8 km 2. This amount of energy, in a hypothetical full coverage scenario, is able to provide for 48.7% of Athens's total energy requirement. Similar year-long simulations were conducted using the EU's largest rooftop solar installation at Stavros Niarchos Foundation Cultural Center and randomly selected rooftops having solar installations in different municipalities in Athens. With these estimated solar potential values, the gross savings in natural gas consumption and hence the CO 2 equivalent emissions can be computed. With the current estimated solar potential of Athens, the analyzed savings accounted of nearly 2.43 billion euros and 18 MT CO 2 equivalent emissions. These computed annual savings are capable of covering installation costs for nearly 100,000 new solar installations. The end-product of this study is the development of a solar cadastre web tool which will support the decision-makers in the energy transition policies and the solar PV penetration into the urban environment and eventually drive the effort into renewable energy transition across the globe.
... Obalanlege et al. (2022) steered an economic and environmental analysis of the PV/T HP system under UK climatic conditions and reported that the recurring cost of the proposed system is low and destroys CO 2 by 610 kg annually throughout its lifetime. Zhang et al. (2014) investigated the socio-economic assessment of the IE PV/T HP system with a loop heat exchanger. They found that it is economically suitable for London's climatic conditions and environmentally most suitable for Shanghai's climatic conditions. ...
Article
Full-text available
A sustainable, affordable, and eco-friendly solution has been proposed to address water heating, electricity generation, space cooling, and photovoltaic (PV) cooling requirements in scorching climates. The photovoltaic thermal system (PV/T) and the direct expansion PV/T heat pump (PV/T DXHP) were numerically studied using MATLAB. A butterfly serpentine flow collector (BSFC) and phase change material (PCM) were assimilated in the PV system and MATLAB model was developed to evaluate the economic and enviroeconomic performance of the PV/T water system (PV/T-W), PV/T PCM water system (PV/T PCM-W), the PV/T DXHP system, and the PV/T PCM heat pump system (PV/T-PCM-DXHP). In this study, annual energy production, socioeconomic factors, enviro-economic indicators, and environmental characteristics are assessed and compared. Also, an economic, environmental, and enviro-economic analysis was conducted to assess the commercial viability of the suggested system. The PV/T PCM-DXHP demonstrated the highest electrical performance of 53.69%, which is comparatively higher than the other three configurations. The discounted levelized cost of energy (DLCOE) and payback period (DPP) of the PV/T PCM-DXHP were ₹2.87 per kW-h and 3–4 years, respectively, resulting in a total savings of ₹67,7403 over its lifetime. Furthermore, installing this system mitigated 280.72 tonnes of CO2 emissions and saved the mitigation cost by ₹329,700 throughout its operational lifecycle. Graphical Abstract
... For instance, previous authors [39] concluded that almost all the required heating loads in the winter months in Nanjing and Hong Kong (China) could be met using a PV-T-HP system with a variable speed compressor. Zhang et al. [40] analysed the economic performance of a solar photovoltaic/ loop-heat-pipe and HP system for domestic heating and concluded that local utility prices and renewable incentives are critical for the financial feasibility of these systems. Recently, some authors proposed the integration of a roll-bond PV-T collector as the evaporator of a single-stage compression HP for the provision of heat and power in northern China during summer [41]. ...
... In summary, microchannel FHPS are still mostly used in battery thermal management systems [23][24][25][26][27][28][29] and solar collectors [30][31][32][33][34][35], and there is still relatively little research into their application in building envelopes, and when heat pipes are used in buildings, some bending of the heat pipes is inevitable to adapt to the room structure, however, the combined heat transfer properties of the bent FHP and the PCM are yet unknown, despite current research on bending heat pipes. The objective of this paper, which also suggests a novel type of FHP phase change radiation ceiling structure, is to investigate the influence of bending on the thermal performance of the microchannel FHP and the coupled heat transfer mechanism of the FHP phase change heat storage module. ...
Article
Using renewable energy, especially solar energy, is essential to achieve a low-carbon society. PCMs suffer from low thermal conductivity, which hinders the efficiency of phase change thermal storage systems. Heat pipes exhibit vastly superior thermal conductivity, making them a promising candidate for enhancing PCM-based systems. To solve the contradiction between building insulation and importing solar energy into buildings in winter, this study proposes a new building heating ceiling system consisting of a flat-plate heat pipe (FHP) coordinated with a phase change material (PCM), and establishes a platform for a heat stockpiling device integrating the FHP and PCM. This research investigates the impact of heating power and bending angle on the heat transfer efficiency of the heat pipe and phase change heat storage system. The experimental findings demonstrate a direct relationship between increasing heating power and the enhancement of the FHP's heat transfer efficiency. When the FHP is bent at a 5° angle and heating power is increased from 20 W to 80 W, to store heat energy into the PCM cavities, the effective thermal conductivity rises significantly, escalating from 7.26 × 104 W/(m·°C) to 15.1 × 104 W/(m·°C), marking a remarkable 108 % rise. Moreover, although increasing the bending angle leads to heightened effective thermal conductivity of the FHP under constant heat input, and the heat transfer in PCM cavities will be weakened within a bending angle of 15°, experimental results demonstrate that the melting time of PCM still decreases with the increase of bending angle; the heat transfer inertia of phase change materials is dominant.
... To reduce this voltage, drop, thermoelectric devices could be a better option. In this study, the Maximum Power Point Tracking Algorithm Perturb and Observe (P&O) is also applied because its inclusion in the electric charge controllers could be used for extracting maximum available power from the PV module under certain conditions [7]- [9]. This maximum power could be filtered to DC through a DC-DC converter for the battery charging and ultimately utilized by DC loads. ...
Preprint
Full-text available
p>Solar energy being the cheapest source of renewable energy is abundantly available on earth. The energy produced by solar panels is conditioned by several circumstances in photovoltaic (PV) systems, that is, solar radiation, temperature, angle of the incident light, etc. With rising panel temperature, solar panel efficiency and photovoltaic module output power fall by 0.5% per °C. This study analyzes a hybrid solar thermoelectric system using Thermoelectric Generator (TEG) and Thermoelectric Cooler (TEC) to boost PV system efficiency. Peltier and Seebeck effects of thermoelectric devices in PV systems cool and generate voltage. By utilizing these effects thermoelectric devices improve the efficiency and output power of PV systems by cooling them and generating voltage from their excess heat. Another important aspect to maximize the efficiency of the hybrid solar thermoelectric system is the continuous monitoring of the maximum power point (MPP). In this study, the Maximum Power Point Tracking (MPPT) algorithm is applied to the hybrid solar thermoelectric system to obtain maximum power and enhance the efficiency of the system in unfavorable conditions. The use of thermoelectric devices based on silicone material can lead to increase the efficiency of the normal solar system by 9% to 4.1% at the ideal temperature of 25°C to the extreme temperature of 55°C when thermoelectric module will be work as a generator and during hot weather conditions IoT based thermoelectric module which will work as a cooler of solar panel can increase the efficiency of system by 0% to 4.1% at the ideal temperature of 25°C to the extreme temperature of 55°C by just keeping down the temperature of solar panel respectively. This study is also tested via IoT based controls of loads as well as thermoelectric coolers to make overall system more advance and efficient to control. </p
Article
Full-text available
The main obstacle that greatly affects solar power output systems is the low conversion efficiency of solar panels, which is greatly influenced by their operating temperature. Failure to consider solar panel temperature increases the financial risks of installing the installation system. In this research, the output performance of solar panels was examined in outdoor conditions. All data was measured and recorded from 09.00 to 17.00, over 30 minute intervals. Panel temperature measurements were carried out using a Ditec C355 infrared thermometer. And then compared with using Pvsyst software. The output power of a solar panel is highly dependent on the solar radiation that falls on its surface. The amount of incoming sunlight is much higher during the hours from 11.00 to 13.00, which can be determined as the peak of the sun during the day. So it can be concluded that the ideal climate for setting up a large solar installation system is a cool and sunny climate.
Article
Full-text available
This article presented a theoretical analysis of the heat transfer limits associated with a gravitational loop heat pipe (LHP), which involves the utilization of an innovative liquid feeding/distributing and vapour/liquid-separating structure. The mathematical equations governing the heat transport capacity were applied to simulate several commonly known heat transfer limits of the pipe, namely, viscous, sonic, entrainment, capillary, boiling and liquid filling mass limits. This will allow the determination of the actual figure of the limitation and analyses of the factors effecting the limits, including the loop operational temperature, wick type, evaporator diameter/length, evaporator inclination angle, vapour column diameter in the three-way fitting, liquid filling mass and evaporator-to-condenser height difference. During the study, the heat-transfer limits associated with the three-way fitting for liquid feeding/distribution and vapour/liquid separation were given particular attention. The results derived from the analytical model indicated that the compound screen mesh wick can achieve better thermal performance over the sintered powder and open rectangular groove wicks. It was also found that the heat transport capacity of such LHP operation is positively proportional to the operational temperature, evaporator diameter, evaporator inclination angle, vapour column diameter within the three-way fitting, liquid filling mass and evaporator-to-condenser height difference, and in a reciprocal order to the evaporator length. With the specified loop configuration and operational conditions, the LHP can achieve a high heat transport capacity of around 900 W. Overall, the work presented in this article provided an approach to determine the heat transfer limitations for such a specific LHP operation that will be of practical use for the associated system design and performance evaluation.
Article
Full-text available
In this paper, a novel solar photovoltaic/loop-heat-pipe (PV/LHP) module-based heat pump system was designed and fabricated for both electricity and hot water generation. A coated aluminium-alloy (Al-alloy) sheet was applied as the baseboard of PV cells for enhanced heat dissipation to the surroundings, which was characterised by a series of laboratory-controlled conditions over the conventional Tedlar–Polyester–Tedlar (TPT) baseboard. The whole prototype system was subsequently evaluated in outdoor weather conditions throughout a consecutive period for about one week. Impact of several external parameters to the PV panel with different baseboards was discussed and the results showed that weaker incident radiation, lower air temperature, higher wind speed, and ground mounting solution, were propitious to the PV electrical performance. Given the specific indoor testing conditions, temperature of the Al-alloy based PV cells was observed at about 62.4 °C, which was 5.2 °C lower than that of the TPT based PV cells, and its corresponding PV efficiency was about 9.18%, nearly 0.26% higher than the TPT based type. During the outdoor testing, the mean daily electrical, thermal and overall energetic and exergetic efficiencies of the PV/LHP module were measured at 9.13%, 39.25%, 48.37% and 15.02% respectively. The basic-thermal system performance coefficient (COP th) was found at 5.51 and the advanced system performance coefficient (COP PV/T) was nearly 8.71. A simple comparison was also conducted between the PV/LHP based heat-pump system and those conventional solar/air energy systems, which indicated that this advanced system harvests larger amount of solar energy and therefore enables enhanced solar efficiency and system performance. Basic analysis into the economic and environmental benefits of this prototype system further demonstrated such technology will be competitive in the future energy supply industry with a payback period of 16 (9) years and a life-cycle carbon reduction of 12.06 (2.94) tons in Shanghai (London).
Article
Solar thermal is one of the most cost-effective renewable energy technologies, and solar water heating is one of the most popular solar thermal systems. Based on the considerations on the existing barriers of the solar water heating, this research will propose a novel façade-based solar water heating system employing a unique loop heat pipe (LHP) structure with top-level liquid feeder, which will lead to a façade-integrated, low cost, aesthetically appealing and highly efficient solar system and has considerable potential to provide energy savings and reduce carbon emissions to the environment. The research initially involved the conceptual design of the proposed system. The prefabricated external module could convert the solar energy to heat in the form of low-temperature vapour. The vapour will be transported to indoors through the transport line and condensed within the heat exchanger by releasing the heat to the service water. The heated water will then be stored in the tank for use. An analytical model was developed to investigate six limits to the loop heat pipe’s operation, i.e., capillary, entrainment, viscous, boiling, sonic and filled liquid mass. It was found that mesh-screen wick was able to obtain a higher capillary (governing) limit than sintered-powder. Higher fluid temperature, larger pipe diameter and larger exchanger-to-pipes height difference would lead to a higher capillary limit. Adequate system configuration and operating conditions were suggested as: pipe inner diameter of 16 mm, mesh-screen wick, heat transfer fluid temperature of 60oC and height difference of 1.5 m. This research further developed a computer model to investigate the dynamic performance of the system, taking into account heat balances occurring in different parts of the system, e.g., solar absorber, heat pipes loop, heat exchanger, and tank. Data extracted from two previously published papers were used to compare with the established model of the same setups, and an agreement could be achieved under a reasonable error limit. This research further constructed a prototype system and its associated testing rig at the SRB (Sustainable Research Building) Laboratory, University of Nottingham and conducted testing through measurement of various operational parameters, i.e., heat transfer fluid temperature, tank water temperature, solar efficiency and system COP (Coefficient of Performance). Two types of glass covers, i.e., evacuated tubes and single glazing, were applied to the prototype, and each type was tested on two different days of 8 hours from 09:00:00 to 17:00:00. By comparison of the measurement data with the modelling results, reasonable model accuracy could be achieved in predicting the LHP system performance. The water temperature remained a steady growth trend throughout the day with an increase of 13.5oC for the evacuated tube system and 10.0oC for the single glazing system. The average testing efficiencies of the evacuated tube system were 48.8% and 46.7% for the two cases with the testing COPs of 14.0 and 13.4, respectively. For the single glazing system, the average testing efficiencies were 36.0% and 30.9% for the two cases with the COPs of 10.5 and 8.9, respectively. Experimental results also indicated that the evacuated tube based system was the preferred system compared to the single glazing system. This research finally analysed the annual operational performance, economic and environmental impacts of the optimised evacuated tube system under real weather conditions in Beijing, China by running an approved computer model. It was concluded that the novel system had the potential to be highly-efficient, cost-effective and environmentally-friendly through comparison with a conventional flat-plate solar water heating system.
Article
Heat Pipes, 6th Edition, takes a highly practical approach to the design and selection of heat pipes, making it an essential guide for practicing engineers and an ideal text for postgraduate students. This new edition has been revised to include new information on the underlying theory of heat pipes and heat transfer, and features fully updated applications, new data sections, and updated chapters on design and electronics cooling. The book is a useful reference for those with experience and an accessible introduction for those approaching the topic for the first time. © 2014 David Reay, Peter Kew, Ryan McGlen Published by Elsevier Ltd All rights reserved.
Article
Direct expansion solar assisted heat pump systems have widely been used in solar drying, water heating, space heating, space air conditioning and cold storage applications. With the high potentials of its efficient heat retrieving unit to effectively use low temperature solar energy, the investigation of the heating and cooling applications of the system has advanced within the past decade. Also, large numbers of refrigerants with high potentials for use in direct expansion solar assisted heat pump have been explored. Investigation results have shown ways of evaluation and various factors determining the performance of the system. These investigations are necessary in order to extend design and performance knowledge and to improve the technology in general. This paper summarizes various investigations and analysis of direct expansion solar assisted heat pump systems. The review rightly point out the obvious shortage of investigation in the cooling application area of the technology and suggest possible investigation area for improving its cooling applications.