Sie sind auf Seite 1von 17

Solar Energy 74 (2003) 253269

The yield of different combined PV-thermal collector designs


H.A. Zondag a , *, D.W. de Vries a , W.G.J. van Helden b , R.J.C. van Zolingen c ,
A.A. van Steenhoven a
a

Eindhoven University of Technology, P.O. Box 513, 5600 MB Eindhoven, The Netherlands
Energy Research Centre of the Netherlands ECN, P.O. Box 1, 1755 ZG Petten, The Netherlands
c
Shell Solar Energy B.V, P.O. Box 849, 5700 AV Helmond, The Netherlands
Accepted 6 March 2003

Abstract
Various concepts of combined PV-thermal collectors are possible. These concepts differ in their approach to obtain the
maximum yield and it is not easy to say whether the yield of a complicated design will be substantially higher than the yield
of a simpler one. In order to obtain a clearer view on the expected yield of the various concepts, nine different designs were
evaluated. The channel-below-transparent-PV design gives the best efficiency, but since the annual efficiency of the
PV-on-sheet-and-tube design in a solar heating system was only 2% worse while it is easier to manufacture, this design was
considered to be a good alternative.
2003 Elsevier Ltd. All rights reserved.

1. Introduction
A combined PV-thermal collector, henceforth to be
called a PVT-collector, consists of a PV-laminate that
functions as the absorber of a thermal collector. In this way
a device is created that converts solar energy into both
electrical and thermal energy. The main advantages of a
PVT-collector are as follows.
1. An area covered with PVT-collectors produces more
electrical and thermal energy than a corresponding area
partially covered with conventional PV systems and
partially covered with conventional thermal collectors.
This is particularly useful when the amount of space on
a roof is limited, which will become increasingly
important in the future.
2. PVT-collectors provide architectural uniformity on a
roof, in contrast to a combination of separate PV- and
thermal-systems.
3. Due to the fact that only one type of system has to be
installed instead of two, a reduction of installation costs
is possible.
In contrast to the situation for conventional thermal
*Corresponding author. Tel.: 131-224-564941; fax: 131-224568966.
E-mail address: h.a.zondag@wtb.tue.nl (H.A. Zondag).

systems, the literature on combined photovoltaic-thermal


collector design is very limited. Research in this field was
carried out in the late 1970s and early 1980s in the USA. A
systematic investigation was made at the MIT in which
several new designs were suggested (Hendrie, 1982), while
a further optimisation study was carried out at Brown
University (Russell et al., 1981). Since then, the effort
invested in optimising the overall design of the PVTcollector has been limited; with respect to flat-plate PVTliquid almost all the work focuses on the optimisation of
sheet-and-tube designs. Theoretical work in this area was
carried out by Bergene and Lvvik (1995) while a large
amount of experimental work, also covering the possibilities of booster reflectors, was carried out at the university of Patras (Tripanagnostopoulos et al., 2002). In
Germany and Denmark, where also commercial parties
were involved in the research, the focus was likewise on
sheet-and-tube (see e.g. Rockendorf et al., 1999; Srensen,
2001). Exceptions were the work of Sandnes and Rekstad
(2002), examining a PV-thermal collector based on a
polymer channel absorber plate and the work of Bakker et
al. (2002), examining a two-absorber PV-thermal collector.
With respect to the optimisation of PV for PVT, early
theoretical investigations were carried out by Cox and
Raghuraman (1985), while more recently experimental
work was carried out by Affolter et al. (2000) and Platz et

0038-092X / 03 / $ see front matter 2003 Elsevier Ltd. All rights reserved.
doi:10.1016 / S0038-092X(03)00121-X

H. A. Zondag et al. / Solar Energy 74 (2003) 253269

254

Nomenclature
A
c
D
DAB
F
g
G
h
hAB
I
H
K
k
Lc
~
m
Nu
n
Pr
q
Q latent
R
Ra
Re
Sh
Sc
T
T red
v
Uloss
V
W
X
b
d

h
h0
u
n
ta
tPV
f

collector surface area (m 2 )


specific heat (J / kg K)
tube diameter (m)
mass diffusion coefficient (m 2 / s)
view factor ()
gravitational acceleration (m 2 / s)
irradiation (W/ m 2 )
coefficient of heat transfer (W/ m 2 )
coefficient of mass transfer (m / s)
current (A)
height of insulation air layer (m)
extinction coefficient (m 21 )
thermal conductivity (W/ m K)
length of collector surface (m)
2
mass flow (kg / s m )
Nusselt number ()
refractive index ()
Prandtl number ()
heat flux (W/ m 2 )
latent heat (J / kg)
reflection coefficient ()
Rayleigh number ()
Reynolds number ()
Sherwood number ()
Schmidt number ()
temperature (K)
reduced temperature (Km 2 / W)
velocity (m / s)
overall loss coefficient (W/ m 2 K)
voltage (V)
tube spacing (m)
vapour mass fraction ()
coefficient of expansion of air ()
thickness of layer (m)
coefficient of emissivity ()
efficiency ()
electrical efficiency at standard conditions ()
angle to perpendicular of collector
viscosity (m 2 / s)
transmission absorption coefficient ()
transmission coefficient for layers above PV ()
collector angle with the horizontal

Subscripts
a
abs
ba
ca
conv
crit
el
evap
g
in
mpp
rad
th
topglass
topglass
vap

ambient
absorber
from back to ambient
from cells to absorber
convection
critical
electrical
evaporation
glass
inflow
maximum power point
radiation
thermal
upper surface topglass
lower surface topglass
vapour

H. A. Zondag et al. / Solar Energy 74 (2003) 253269

al. (1997). Overviews of the literature on PVT-collectors


are presented by Leenders et al. (2000) and Bazilian and
Prasad (2000).
The aim of this paper is to make a comparison of the
efficiency of seven different design types of PVT-collectors to make clear what the effects of design choices are on
the electrical and thermal efficiency of the design. The
efficiency curves will be compared. In addition, a comparison will be made of the annual yield of the PVTcollectors for one specific application: a solar heater in a
domestic hot water system in the Dutch climate.

255

standard photovoltaic panel and integrate it into a thermal


collector without any modification. The sheet-and-tube
PVT-collector, that is shown in Fig. 1a, is an example of
such an approach. The thermal insulation of such a design
can be improved by increasing the number of top covers.
However, since each cover creates additional reflections,
this strategy reduces the electrical yield of the PVTcollector. Sheet-and-tube PVT-collectors are examined
with zero, one and two covers. Designs with more than
two glass covers do not seem to have practical applications, since the electrical efficiency is reduced too much.

2.3. Channel PVT-collector


2. Design concepts

2.1. Introduction
Nine design concepts for water-type PVT-collectors are
evaluated here, which can be classified in four groups. An
example of each group is given in Fig. 1.
A. Sheet-and-tube PVT-collectors.
B. Channel PVT-collectors.
C. Free flow PVT-collectors.
D. Two-absorber PVT-collectors.

2.2. Sheet-and-tube PVT-collector


The simplest way to construct a PVT-collector is to rely
entirely on well-known available technology by taking a

Fig. 1b shows a channel PVT-collector with the channel


on top of the PV. Such a configuration imposes constraints
on the choice of the collector fluid; for a PVT-collector
design the absorption spectrum of the fluid should be
sufficiently different from the absorption spectrum of the
PV in order to allow the PV to receive the incoming
radiation. In the present design water is used, which has a
small overlap in absorption with the PV, resulting in 4%
relative decrease in the electrical performance. A disadvantage of this design is that, if a wide channel is used
that is covered by one large glass plate, very thick glass
may be necessary to withstand the water pressure, resulting
in a heavy but fragile construction (Bakker et al., 2002). A
variation of the channel design is obtained by letting the
water flow underneath the PV panel. Two cases are

Fig. 1. Various collector concepts: (A) sheet-and-tube PVT, (B) channel PVT, (C) free flow PVT, (D) two-absorber PVT (insulated type).

256

H. A. Zondag et al. / Solar Energy 74 (2003) 253269

examined: a channel below a conventional opaque PV


panel and a channel below a transparent PV panel with a
separate black thermal absorber underneath the channel.
The latter can be expected to have a higher thermal
efficiency, but it should be kept in mind that transparent
PV laminates are at present substantially more expensive.
In addition, the PV panel should be able to withstand the
pressure of the water. In the case of the opaque PV this
might be less of a problem since the backside could be
strengthened with a metal back. Care should be taken that
the backside of the PV laminate is sufficiently watertight.

2.4. Free flow PVT-collector


In a free flow panel unrestrained fluid flows over the
absorber, as shown in Fig. 1c. In comparison to the
channel case, this design eliminates one glass layer. Thus
reflections and material costs are reduced, while the
mechanical problem of breaking the glass cover is avoided.
A disadvantage is the increased heat loss due to evaporation. As in the case of the channel PVT-collector, the fluid
flowing over the PV panel has to be transparent for the
solar spectrum. Water seems a natural choice, but, since its
evaporation pressure is not very low, evaporation will be
shown to create problems at higher temperatures.

2.5. Two-absorber PVT-collector


The two-absorber panel uses a transparent PV laminate
as a primary absorber and a black metal plate as a
secondary absorber. The panel contains two water channels
on top of each other. The water flows in through the upper
channel and is returned through the lower channel. This
design was also examined as part of the PV-thermal
collector development program of the MIT (Hendrie,
1982), which indicated a high thermal efficiency for this
design. However, the disadvantage of the heavy channel
coverindicated previously for the channel PVTapplies
even more strongly for the two-absorber PVT (Bakker et
al., 2002). Several variations can be applied in the design.
The efficiency of the design could be improved by an
additional transparent insulating layer between the primary
and the secondary channel to reduce the heat loss further.
On the other hand, this reduces the robustness of the
collector. This insulated version is shown in Fig. 1d.
Another possible variation is obtained by replacing the
channel under the secondary absorber by a sheet-and-tube
construction, but this option is not examined here.

concept level. Of course, other materials could have been


chosen, such as transparent plastics instead of glass or
other types of PV. Highly transparent plastics would
probably give a small increase in optical efficiency, but
their material properties, such as sensitivity to UV or high
temperatures as well as limited watertightness, would
probably limit their application in PV-thermal. A different
type of PV would change the ratio of heat production over
electricity production (due to a different electrical efficiency) and change the optical efficiency of the laminate
(due to different reflection losses). In addition, a small
effect of temperature on PV-efficiency would be present, as
e.g. the yield of amorphous silicon PV is less sensitive to
temperature.

3. Optical efficiency
In order to determine the thermal efficiency of the
PVT-collector, two models are required. Whereas a thermal
model is required to determine the heat flows within the
PVT-collector, an optical model is required to determine
how much irradiation is absorbed by the PVT-collector.
The optical model is used to calculate the transmission
absorption factor of the PVT-collector ta , and this value is
then inserted as a constant into the thermal model.
The optical model presently used is based on the net
radiation method. This method solves the energy flux
balance at each interface in the PVT-collector configuration. As an example, the reflection at the glass cover is
given in Fig. 2. The energy flux equations for this example
are now presented by
qi,1 5 G
qo,1 5 R 1,2 qi,1 1s1 2 R 1,2d qi,2
qo,2 5s1 2 R 1,2d qi,1 1 R 1,2 qi,2

(1)

qi,2 5 qo,3 exp f 2 Kdtopglass / cos u g


The values for the coefficients of reflection R 1,2 are
determined from the Fresnel equations:

2.6. Materials selection


In the calculation of the various concepts, we have used
materials that are routinely applied in conventional PV
laminates and thermal collectors. In addition, we decided
to use the same materials for all concepts as much as
possible, in order to be able to make a valid comparison on

Fig. 2. The numbering of the radiative flux densities in the net


radiation method.

H. A. Zondag et al. / Solar Energy 74 (2003) 253269

1
R 5 ]sR 1 R d
2

2
2
1 sin fui 2 ur g tan fui 2 ur g
5 ] ]]]]
1 ]]]]
2
2
2 sin fui 1 ur g tan fui 1 ur g

(2)

in which the angles are determined using Snells law:


sin ur / sin ui 5 n i /n r

(3)

This method is applied to all the interfaces in the


PVT-collector, which generates a set of equations that is
solved by a matrix-solving procedure. Since both the
coefficient of extinction K and the index of refraction n
depend on the wavelength, the equations are solved for
each wavelength interval separately and then integrated
over the solar spectrum. The solar radiation is assumed to
have no nett polarisation, so the incoming light is split in
50% parallel and 50% transverse polarisation, ta is determined for each mode separately and the results are
added. The calculation is based on the assumption of
specular reflection, so diffuse reflection is not taken into
account. A complication is presented by the fact that a
PV-laminate does not present a homogeneous surface but
consists of different parts (active PV-area, the top grid and
the spacing between the cells). For each part the value for
ta is calculated separately and then ta of the entire PVTcollector is determined by taking the average of these
values, weighed with the respective surface areas. This
method results in a slight underestimation of ta due to the
fact that no exchange of light between the different
material surfaces is possible; light reflected by the top grid
area cannot be reflected back by the glass on the PV area.
In the analysis, reference is made to two types of PVpanels: either opaque or transparent. The opaque PV-panel
is a standard Shell Solar multi-crystalline PV-panel consisting of a sandwich of glass / EVA / TiO 2 / Si / EVA / PE-Altedlar. The transparent PV-panel is essentially the same but
the PE-Al-tedlar layer at the back has been exchanged for
a glass plate.

257

For each of the different design concepts, a different


value is found for the absorption of the PV-combi collector. In particular, for the concepts in which water flows
over the PV laminate, both the transmissionabsorption
coefficient of the laminate, ta , and the transmissionabsorption coefficient of the water, ta ,water , have to be
distinguished. The latter represents the relative amount of
the incoming energy that is absorbed by the water layer,
while the former represents the relative amount of energy
collectively absorbed by the solid material layers in the
PVT-collector. Similarly, for the determination of the
electrical yield, the amount of radiation received by the
PV-laminate is determined by the number of glass and
water layers above the PV-laminate, the combined transmission of which is given by tPV . The calculated values for
normal incidence are presented in Table 1. Fig. 3 indicates
the normalised angular dependence of ta and tPV . This
figure shows that the curves for the one- and two-cover
designs are almost the same, while the values for the zero
cover case are somewhat larger in the interval between 50
and 80 degrees. The value of ta (u ) can be obtained by
multiplying the value from Table 1 with the angular
dependence factor in Fig. 3.

4. Thermal efficiency
The thermal model is steady-state, based on solving the
heat balance for all the layers in the PVT-collector. For
more information on the model, see Zondag et al. (2002).
The transmissionabsorption factor of the PVT-collector
was calculated with the optical model and the calculated
values are indicated in Table 1. The value of this parameter
is inserted into the corresponding thermal-yield model.
With respect to the effect of the electrical energy production, the electrical efficiency, which is a function of
temperature, is subtracted from the transmissionabsorption factor to find the amount of thermal energy that was
absorbed in the system, according to ta ,eff 5 hta 2 tPVhel j.

Table 1
Transmissionabsorption factors for the various design concepts (AM 1.5 spectrum)
Design concept

ta

ta ,water

tPV

Uncovered sheet-and-tube PVT-collector


One-cover sheet-and-tube PVT-collector
Two-cover sheet-and-tube PVT-collector
Channel above PV
Channel underneath opaque PV
Channel underneath transparent PV

0.78
0.74
0.71
0.62
0.74
0.55 first absorber
0.12 second absorber
0.62
0.52 first absorber
0.12 second absorber
0.52 first absorber
0.11 second absorber

0.16

0.10

1
0.92
0.84
0.87
0.92
0.92

0.17
0.16

0.88
0.87

0.14 upper channel


0.013 lower channel

0.87

Free flow PVT-collector


Two-absorber PVT-collector (insulated type)
Two-absorber PVT-collector (non-insulated type)

258

H. A. Zondag et al. / Solar Energy 74 (2003) 253269

Fig. 3. The normalised angular dependence of ta and tPV .

Here, tPV is obtained from Table 1. In this way one obtains


the amount of absorbed energy that contributes to the
thermal yield.
The results of the measurements and the calculations are
the yield and the efficiency of the collector. The yield of
the collector is defined as the amount of useful energy
produced by it, while the efficiency is defined as the yield
divided by the amount of solar energy received by the
collector. In this way, both an electrical and a thermal
efficiency are defined

equations relating the heat flows and the temperatures of


the various layers in the PVT-collector. These equations
are given in Appendix A (Eqs. (A.1)(A.10)) since they
are the same for all designs. These general equations are
related to each other through the heat balance equations,
which depend on the specific design that is examined.
Therefore, they will be presented below in separate
paragraphs for each design.

4.1. Sheet-and-tube collector


csT out 2 T ind
~ ]]]]
hth ; m
G

(4)

VMPP IMPP
hel ; ]]]
GA

(5)

VMPP and IMPP represent the voltage and the current in the
maximum power point, A is the panel area, c is the heat
~ the mass flow
capacity of the collector medium (water), m
per square meter in kg / s / m 2 , G the irradiation in W/ m 2
and T in and T out are the inflow temperature and the outflow
temperature in 8C, respectively. The thermal efficiency is
conventionally shown as a function of reduced temperature, which is defined as
T in 2 T a
T red ; ]]
G

(6)

in which T a represents the ambient temperature.


In the presented models, Eq. (7) is used for the electrical
efficiency of the PV

hel 5 h0s1 2 0.0045 f T 2 25 8C gd

(7)

in which h0 5 0.097 for both conventional and transparent


PV. Table 2 shows the values that are used in the models
for the characteristic collector parameters that are used in
the calculations. The values for the heat conduction
through air and water in this table are the values at 20 8C;
in the simulations the temperature dependence of these
quantities is taken into account as well. The quantities
indicated in Table 2 appear in the general heat flow

The heat flows in the sheet-and-tube PVT-collector are


indicated schematically in Fig. 4. For the sheet-and-tube
PVT-collector, the heat balance is represented by the
following equations:
qwater 5 qca 2 qba

(8)

qca 5sta 2 tPVheldG 2 qPVglass

(9)

Table 2
Values of coefficients used in simulations
Collector length
Emissivity of glass
Emissivity of PV
Heat capacity of water
Heat conduction through air
Heat conduction through glass
Heat conduction through water
Heat transfer to water for channel
Heat transfer to absorber for sheetand-tube
Heat transfer through back of the
collector
Thickness of cover glass
Thickness of PV glass
Tube diameter
Tube spacing
Width of air layer
Width of water channel

Lc
g
PV
c
k air
k glass
k water
h abstowater
h ca

1.776
0.9
0.9
4200
0.025
0.9
0.6
650
500

J / kg K
W/ m K
W/ m K
W/ m K
W/ m 2 K
W/ m 2 K

h back

W/ m 2 K

dtopglass
dPVglass
D
W
H
dwater

3.2
3.0
0.01
0.095
0.02
5

mm
mm
m
m
m
mm

H. A. Zondag et al. / Solar Energy 74 (2003) 253269

259

Fig. 4. The heat flows in the sheet-and-tube PVT-panel, together with the temperature distribution according to the HottelWhillier
equations.

qPVglass 5 qair,conv 1 qair,rad

(10)

qair,conv 1 qair,rad 5 qtopglass

(11)

qtopglass 5 qsky,conv 1 qsky,rad

(12)

The set of equations (8)(12) presented above is for the


case of the panel with one cover. For the zero cover case
Eqs. (10) and (11) are left out, while in Eq. (12) the
qtopglass is changed into qPVglass . For the panel with two
covers, Eqs. (10) and (11) are repeated.
In contrast to the case of the channel collector designs,
for the sheet-and-tube combi collector the PV is connected
to a separate thermal absorber plate by means of an
adhesive layer. Therefore, the heat transfer between the PV
cells and the absorber plate should be modelled as well.
This is taken into account by using Eqs. (13) and (14)
indicated below instead of Eq. (A.8) (in Appendix A)
qca 5 h casT cell 2 T absd

(13)

qba 5 h basT abs 2 T ad

(14)

T topglass ) while the set of Eqs. (8)(12) consists of only


five equations. Therefore, one more equation is required. In
order to solve this, the average absorber temperature T abs
for the sheet-and-tube design is calculated from the
temperature distribution indicated by the HottelWhiller
equations (Duffie and Beckman, 1991, pp. 252283). This
temperature distribution is also shown in Fig. 4. It results
from the fact that a temperature gradient over the absorber
is required to drive the absorbed heat to the tubes. A
mathematical representation of this temperature distribution is given by Eq. (16)

sta 2 tPVheldG
T abssxd 5 T a 1 ]]]]
Uloss
T bond 2 T a 2sta 2 tPVheldG /Uloss
1 cosh smxd]]]]]]]]]
cosh f msW 2 Dd / 2 g
(16a)
with

It is assumed that the heat resistance is minimised by the


application of a highly conductive glue of k50.85 W/ Km
in a layer of 50 mm thickness. With respect to the PVlaminate, a PE-Al-tedlar layer of 0.1 mm with conduction
k 5 0.2 W/ Km and an EVA layer of 0.5 mm and a conductivity of k 5 0.35 W/ Km are assumed (Krauter, 1993).
The heat resistance of the PV-laminate can now be
determined:

5 3 10 24 1 3 10 24 5 3 10 25
5 ]]] 1 ]]] 1 ]]]
0.35
0.2
0.85
W
5 500 ]]2
Km

(16b)

in which the heat loss coefficient and the bond temperature


are given by
qsky,rad 1 qsky,conv 1 qba
Uloss 5 ]]]]]]
T PVglass 2 T a

(17)

T bond 5 T water 1 qwater /h tube

h ca 5sR EVA 1 R tedlar 1 R glued 21

]]]
m ;Uloss /kdabs

5 T water 1 qD/Nu tube k water

21

(18)

Re , 2300 Nu tube 5 4.364


(15)

The equations for the heat flows as given in Appendix A,


together with Eqs. (13) and (14) contain six unknown
temperatures (T out , T cell , T PVglass , T abs , T topglass and

Re . 2300 Nu tube 5 0.023 Re 0.8 Pr 0.4


(Dittus 2 Boelter equation)

(19)

Eq. (16) is numerically integrated with respect to x in


order to provide the average absorber temperature.

H. A. Zondag et al. / Solar Energy 74 (2003) 253269

260

Fig. 5. The heat flows in the PVT-panel with the channel above the PV.

4.2. Channel PVT-collector


The heat flows in the channel PVT-collector are indicated schematically in Fig. 5 for the case of the channel
above the PV. For this concept a similar approach as for
the sheet-and-tube PVT-collector can be followed. However, in the set of equations for the channel, in contrast to
the case of the sheet-and-tube absorber, the water presents
a layer through which part of the heat absorbed by the PV
is lost to the collector surface again. Part of the incoming
heat qwater,in will be retained in the water and drawn off
(qwater ), while another part will be transferred from the
water layer to the glass plate on top of the water channel
(qwater,out ). In addition, the water now absorbs solar energy
directly as well, according to the transmissionabsorption
coefficient for the water ta ,water . The heat balance is
presented by

sta 2 tPVheldG 5 qPVglass 1 qba

(20)

qPVglass 5 qwater,in

(21)

qwater 1 qwater,out 5 qwater,in 1 ta ,water G

(22)

qwater,out 5 qtopglass1

(23)

qtopglass1 5 qair,conv 1 qair,rad

(24)

qtopglass2 5 qair,conv 1 qair,rad


qtopglass2 5 qsky,rad 1 qsky,conv

T topglass1 , T topglass2 and T topglass2 ) and no additional


equation is required for the temperature. The heat transfer
between the PVglass and the water is determined from
k water
Nu 5 5.385 h channel 5 Nu ]] 5 650
dwater

(27)

This value of the Nusselt number is valid for a rectangular


channel that is insulated at one side (Bejan, 1993, p. 298).
The case for the channel-below-transparent-PV (which is
assumed to consist of the package glass / EVA / PV-cell /
EVA / glass) is similar to the case for the channel-above-PV.
However, the heat flow from the PV upwards is now
entirely a loss flow, while the heat flow downwards
contributes to the collected heat. In addition, the additional
cover glass on top of the channel disappears. The heat flow
diagram is shown in Fig. 6.
The heat balance is now presented by

sta 2 tPVheldG 5 qPVglass1 1 qPVglass2

(28)

qPVglass2 5 qwater,in

(29)

qwater 1 qwater,out 5 qwater,in 1 ta ,water G

(30)

qwater,out 1 ta ,abs2 G 5 qba

(31)

(25)

qPVglass 5 qair,conv 1 qair,rad

(32)

(26)

qtopglass 5 qair,conv 1 qair,rad

(33)

qtopglass 5 qsky,rad 1 qsky,conv

(34)

The presence of the additional glass layer which acts as


the ceiling of the water channel presents two additional
equations when compared to the sheet-and-tube case, so
the full set of equations is increased to seven. Similarly,
the additional cover increases the number of temperatures
by two, but the absence of a heat transfer between cells
and absorber reduces the number of temperatures again by
one. Therefore, the amount of unknown temperatures is
increased to seven (T out , T cell , T PVglass , T topglass1 ,

Both qwater,in and qwater,out are determined from Eq. (27),


while qPVglass2 is defined by Eq. (A.9).
For the case with the opaque PV, due to the fact that the
PV is not transparent, the terms representing the absorption
by the water and the secondary absorber disappear from
Eqs. (30) and (31). However, the base set of heat balance
equations is the same as for the case with transparent PV.

H. A. Zondag et al. / Solar Energy 74 (2003) 253269

261

Fig. 6. The heat flows in the PVT-panel with the channel below the PV.

4.3. Free flow PVT-collector


The heat flows in the free flow PVT-collector are
indicated schematically in Fig. 7. In the free flow concept
the solving procedure is much like the PVT-collector
design with the channel above the PV. However, in the free
flow panel evaporative heat transfer takes place. The model
assumes that the evaporation process at the water film on
the PV laminate and the subsequent condensation process
at the top glass layer are much faster than the actual
transport of the water vapour through the air / vapour
mixture due to concentration differences. Therefore, this
diffusion process determines the evaporative heat transfer
coefficient. The natural convection due to the temperature
difference over the air layer enhances this process but this
effect is not accounted for in the present calculations.
Firstly, the natural convection is calculated as if no
diffusion of vapour is present. Next, the mass flux of
vapour is calculated that also represents a heat flux due to
the latent heat of the vapour.
Because of the similarity of the free flow design with the
PVT-collector design with the channel above the PV, the
equations are largely the same as in the latter case. The
difference is the absence of the topglass1, so that thermal
radiation and convection take place directly between the
water layer and the topglass2, which leads to the disappearance of Eqs. (23) and (24). Therefore, only five

equations are retained. However, the fact that air is present


above the fluid introduces a new heat transfer mechanism
due to evaporation. The equation for the corresponding
qev ap is derived in Appendix B.

sta 2 tPVheldG 5 qPVglass 1 qback

(35)

qPVglass 5 qwater,in

(36)

qwater 1 qair,rad 1 qair,conv 1 qair,evap 5 qwater,in 1 ta ,water G


(37)
qair,rad 1 qair,conv 1 qair,evap 5 qtopglass

(38)

qtopglass 5 qsky,rad 1 qsky,conv

(39)

Similar to the case of the channel design, the heat transfer


between the PVglass and the water is again provided by
Eq. (27). Since only five temperatures have to be calculated explicitly (T out , T cell , T PVglass , T topglass and
T topglass ), no additional equations are required.

4.4. Two-absorber PVT-collector


The heat flows in the insulated version of the twoabsorber PVT-collector are indicated schematically in Fig.
8. The equations are similar to those for the channel

Fig. 7. The heat flows in the free flow PVT-panel.

262

H. A. Zondag et al. / Solar Energy 74 (2003) 253269

Fig. 8. The heat flows in the two-absorber PVT-panel.

PVT-collector. However, an additional set of equations


appears that is related to the second water channel.
Therefore, we now also get qwater2,in , qwater2,out and qwater2
that are related to the heat flow from the second absorber
to the water
qba 5 qwater2,out

(40)

qwater2,in 5 qwater2 1 qwater2,out

(41)

ta 2 G 5 qair2,conv 1 qair2,rad 1 qwater2,in

(42)

qair2,rad 1 qair2,conv 5 qPVglass2

(43)

sta 2 tPVheldG 1 qPVglass2 5 qPVglass1

(44)

qwater1,in 5 qPVglass1

(45)

qwater1 1 qwater1,out 5 qwater1,in 1 ta ,water G

(46)

qwater1,out 5 qtopglass1

(47)

qtopglass1 5 qair1,conv 1 qair1,rad

(48)

qtopglass2 5 qair1,conv 1 qair1,rad

(49)

qtopglass2 5 qsky,rad 1 qsky,conv

(50)

In the equations 11 unknown temperatures appear (T back ,


T out1 , T out2 , T abs2 , T PVglass2 , T PVglass1 , T cell , T topglass1 ,
T topglass1 , T topglass2 , T topglass2 ). Since the heat balance
consists of 11 equations, no additional equations are
required.
The solving procedure for this set of equations is more
complex than for the previous cases. In the previous panels
the system is divided into a number of segments in the

direction of the flow and since the situation downstream


does not influence the situation upstream, the calculation is
straightforward. However, in the two-absorber panel the
two water layers cannot be treated independently, since the
temperature of the secondary layer influences the efficiency of the primary layer, which in its turn determines
the inflow temperature of the secondary layer. Because of
this extra interaction between the water layers, an estimate
of the temperatures of the water segments in the secondary
layer is required to calculate the temperature in the primary
layer and the top layer. With these calculated values of the
segment temperatures in the primary layer we can then
obtain a better approximation of the temperatures in the
secondary layer again. This iteration process is continued
until the solution is sufficiently converged. For three
segments in each channel 10 iterations are sufficient.
For the non-insulated version of the two-absorber design, the equations are almost the same. However, Eqs.
(42) and (43) merge due to the absence of the insulating
layer, while the absorption term moves from Eq. (42) to
Eq. (41) due to the fact that the secondary absorber is now
no longer on top of the secondary water channel but
underneath it. On the other hand, Eq. (41) is split due to
the presence of the secondary absorber at the bottom of the
channel. This results in Eqs. (51)(53) coming in place of
Eqs. (41)(43)
qwater2,in 1 ta ,water2 G 5 qwater2 1 qwater2,out

(51)

ta 2 G 1 qwater2out 5 qback

(52)

qPVglass2 5 2 qwater2,in

(53)

Note that qPVglass , as defined in Fig. 8, has become


negative due to the fact that the heat flows from the PV to
the water in the secondary channel.

H. A. Zondag et al. / Solar Energy 74 (2003) 253269

263

Fig. 9. Thermal efficiency for the case with production of electricity of the various PVT-panels.

Fig. 10. Electrical efficiency of the various PVT-panels. The lines for the electrical efficiency of the channel beneath opaque and transparent
PV coincide in the figure.

H. A. Zondag et al. / Solar Energy 74 (2003) 253269

264
Table 3
Ambient conditions used in simulations
Ambient temperature
Irradiance
Wind speed
Mass flow
Sky temperature (clear sky)
Collector angle

T amb
G
vwind
~
m
T sky
w

20 8C
800 W/ m 2
1 m/s
76 kg / m 2 h
4 8C
458

higher reduced temperatures, water is not a good choice for


a free flow collector. An extra complication is caused by
the fact that the condensate on the top glass layer will
cause additional reflection of the radiation (not taken into
account in the numerical model), which will reduce the
efficiency at high reduced temperatures even faster than
shown in Fig. 9. An advantage of the free flow concept is
its intrinsic prevention against overheating because of the
strong increase in the rate of evaporation towards higher
temperatures.

5. Results

5.1. Efficiency curves

5.2. Experimental verification of the results for the one


cover sheet-and-tube PVT-collector

With the set of equations obtained in paragraph 3, the


efficiency of the various designs is calculated as a function
of the reduced temperature. The thermal efficiency curves
are shown in Fig. 9 and the electrical efficiencies in Fig.
10. The ambient conditions used in the simulations are
presented in Table 3. Table 4 shows the thermal and
electrical efficiencies at zero reduced temperature.
The uncovered sheet-and-tube collector is obviously
performing poorest at zero reduced temperature because of
its large heat losses. The sheet-and-tube collectors with
one or two covers have a substantially larger efficiency at
these conditions. It seems that a sheet-and-tube collector
with two covers is only useful for high temperature
applications since the electrical efficiency strongly deteriorates due to the second cover, while the thermal
efficiency does not show a significant increase at modest
values of the reduced temperature. In addition, it can be
observed that the channel concepts (the channel concept,
the free flow panel concept and the two-absorber concept)
all have a substantially higher efficiency than the sheetand-tube collectors because of the excellent heat transfer
properties of a channel.
For the case of the free flow panel, Fig. 9 shows that
evaporation strongly reduces the thermal efficiency at
higher reduced temperatures. It can be concluded that, at

In order to be able to verify the results, a test facility


was built (de Vries, 1998). Since it was not feasible to
build all the designs described, a choice had to be made
which design would be realised. The single cover sheetand-tube design was chosen. It was preferred over the
channel designs because it could be built entirely from
components that were based on well-known technology
and that were commercially available. The single cover
sheet-and-tube was preferred over the other two sheet-andtube designs because, as can be seen in Fig. 9, the thermal
performance strongly decreases if the number of covers is
less, while the electrical performance decreases too much
if more than one cover is used.
A non-optimised first prototype was constructed by
connecting a conventional PV-laminate, containing multicrystalline silicon cells, to the absorber plate of a conventional glass-covered sheet-and-tube collector, as shown
previously in Fig. 1. The panel was then integrated into a
test rig, next to a conventional sheet-and-tube thermal
collector and a multi-crystalline silicon PV-panel of the
same length and width. A photograph of the test rig is
shown in Fig. 11. For more extensive information on the
experiments and experimental conditions, the reader is
referred to de Vries (1998) and Zondag et al. (1999).
The results of the measurements are presented in Fig.

Table 4
Thermal efficiency at zero reduced temperature with simultaneous production of electricity and corresponding electrical efficiency at zero
reduced temperature for various PVT-collector design concepts
Panel type

Thermal
efficiency

Electrical
efficiency

PV laminate
Sheet and tube PVT-collector 0 cover
Sheet and tube PVT-collector 1 cover
Sheet and tube PVT-collector 2 covers
PVT-collector with channel above PV
PVT-collector with channel below opaque PV
PVT-collector with channel below transparent PV
Free flow PVT-collector
Two-absorber PVT-collector (insulated type)
Two-absorber PVT-collector (non-insulated type)
Thermal collector

0.52
0.58
0.58
0.65
0.60
0.63
0.64
0.66
0.65
0.83

0.097
0.097
0.089
0.081
0.084
0.090
0.090
0.086
0.085
0.084

H. A. Zondag et al. / Solar Energy 74 (2003) 253269

265

Fig. 11. The test rig. Left to right: a conventional thermal collector, the PVT-collector and a conventional PV-laminate.

12, together with the results of the simulations. The


temperature difference over the laminate was measured
and for h ca a value of 45 W/ Km 2 was found, which was
substantially below the expected value. The low heat
transfer coefficient between cells and absorber was found
to be due to air enclosure in the glue layer, as well as a
heat conduction coefficient of the glue that turned out to be
substantially below the specified value. The value of 45 W/
Km 2 is used in the simulations shown in Fig. 12. The
thermal efficiency was determined as a function of reduced
temperature for the 2D steady state model. The agreement
between the model and the experiments is considered to be
well within the range of the experimental data.

5.3. Annual yield of a PV-thermal domestic hot water


system
For the calculation of the annual yield, a shell is built
around the program for the calculations of the efficiency
curves, in which the meteorological data and the system
characteristics are specified. The meteorological data are
obtained from the KNMI test reference year. In addition,
the sky temperature is calculated for every hour with
TRNSYS by means of a correlation that determines the
cloudiness factor and the clear sky emittance from the
parameters in the test reference year. The system consists
of two 1.75 m 2 PVT-collectors connected in parallel

Fig. 12. Left, measured thermal efficiency (3) thermal collector, (s,1) PVT-collector either without or with electricity production. The
uncertainties (least square fits) in the measurements are presented by the bar lengths. Right, least squares fit of the measurements of the
thermal efficiency (solid) compared to the results obtained with the 2D model (dashed). Upper line, conventional thermal collector; middle
line, PVT-collector not producing electricity; lower line, PVT-collector producing electricity.

H. A. Zondag et al. / Solar Energy 74 (2003) 253269

266

Table 5
ISSO warm water withdrawal schedule, (2) no withdrawal, (1) 175 / 8 l withdrawal
Hour

10

11

12

13

14

15

16

17

18

19

20

21

22

23

24

Tapping

together with a storage vessel of 175 l. A conventional


mass flow of 50 l /(m 2 h) is assumed, leading to a total
mass flow of 175 l per hour. It is assumed that tapping
takes place every day of the year according to the same
pattern. For the tapping during the day the ISSO tapping
scheme is used, that is displayed in Table 5.
During a day, the angle of the irradiation changes. It was
assumed that all radiation (direct and diffuse) comes out of
the direction of the sun. In Fig. 3 it was found that the
normalised angular dependence of the one-cover and the
two-cover panels is almost the same. Since the new
designs can be expected to have an absorption that follows
the absorption of the one-cover case even more closely
than the two-cover panel, the one-cover normalised angular dependence curve is used to determine the angular
dependence of the channel-, free flow- and two-absorber
designs as well.
In the calculation of the thermal yield, heat losses from
the storage vessel and the tubing to the ambient are
included in the model. For the calculation of the heat loss,
a constant local temperature of 10 8C is assumed around
the vessel and the tubing. The heat loss coefficients for the
vessel and the combined tubing are taken to be 0.4 W/ Km 2
and 3 W/ K, respectively. Mixing within the vessel is not
taken into account. The vessel is replenished with water of
10 8C. These calculations led to the results presented in
Table 6.
This table shows that the insulated two-absorber design,
the two-cover sheet-and-tube design and the channelabove-PV design have the highest thermal yield for this
domestic heating configuration, while the uncovered sheetand-tube design performs poorest. Second in line with

respect to the thermal yield are the non-insulated twoabsorber design and the channel-below-transparent-PV
design. The lower electrical efficiency of the first four
designs makes the channel-below-transparent-PV design
the most attractive from the efficiency point of view.
However, the basic one-cover sheet-and-tube design performs only 2% worse while it is substantially easier to
manufacture, which seems to make it the most promising
scheme from a marketing point of view. The table also
shows that the free-flow panel performs much poorer than
the channel designs for the present configuration, although
it has roughly the same thermal efficiency as the channel
designs at zero reduced temperature. This is due to the
large thermal loss at higher temperatures due to evaporation.
The presented thermal efficiencies show a substantial
reduction when compared to the thermal efficiency at zero
reduced temperature. In order to understand this reduction,
the efficiency was simulated again in which the loss
mechanisms were removed one by one. The results are
summarised in Table 7. With respect to the thermal losses,
the largest loss is due to the reduced efficiency of the
PVT-collector at higher reduced temperatures. However,
the tubing heat loss also gives a substantial contribution of
5%. Small losses result from the vessel heat loss and the
increased reflection losses due to oblique radiation.
With respect to the calculation of the electrical yield, a
similar procedure can be followed. The results are indicated in Table 8. The largest losses are those due to
inverter losses (0.9%) and top cover reflection (0.8%).
This again underscores the importance of the glass cover in
reducing the PV-efficiency. The next largest losses are

Table 6
Annual average efficiencies for the presented PVT-collector design concepts
System

Annual
thermal
efficiency

Annual
electrical
efficiency

PV
Sheet and tube PVT-collector uncovered
Sheet and tube PVT-collector 1 cover
Sheet and tube PVT-collector 2 covers
PVT-collector with channel above PV
PVT-collector with channel below opaque PV
PVT-collector with channel below transparent PV
Free flow PVT-collector
Two-absorber PVT-collector (insulated type)
Two-absorber PVT-collector (non-insulated type)
Thermal collector

0.24
0.35
0.38
0.38
0.35
0.37
0.34
0.39
0.37
0.51

0.072
0.076
0.066
0.058
0.061
0.067
0.065
0.063
0.061
0.061

H. A. Zondag et al. / Solar Energy 74 (2003) 253269


Table 7
Loss mechanisms affecting the annual thermal yield (values for
one cover sheet-and-tube PVT-collector)
Loss mechanism

Thermal
efficiency

Base case: one cover sheet-and-tube


No oblique irradiation loss
No vessel heat loss
No tubing heat loss
No reduced efficiency at increasing
reduced temperature

0.35
0.37
0.38
0.43
0.58

those due to low (0.6%) and oblique irradiation (0.4%).


The increased laminate temperature accounts for only
0.2%. However, one should realise that this accounts for
the present solar heater configuration only and is linked to
the fact that the average water outflow temperature is only
28 8C.

6. Concluding discussion

267

design. However, the electrical performance of these three


systems is somewhat lower due to the additional glass and
water layers on top of the PV. Taking this into account, it
seems that the channel-below-transparent-PV seems the
best option from the efficiency point of view, while the
one-cover sheet-and-tube design is a good alternative since
its efficiency is only 2% less. Since the latter design is by
far the easiest to manufacture, the single cover sheet-andtube design seems the most promising of the examined
concepts for domestic hot water production.
However, a low-temperature application may show a
different picture. For this case, the uncovered PVT-collector will probably come out better, since the reflection
losses at the cover are foregone in this design, while the
heat losses will remain low because of the low temperature
level required. This gives an interesting perspective for the
combination of an uncovered PVT-collector with a heatpump.

Appendix A. Overview of the heat transport


equations

The presented numerical analysis shows that for a


combined photovoltaic-thermal collector, the total efficiency at zero reduced temperature is over 50%. Therefore, it produces a higher yield per unit area than a thermal
collector and a PV laminate placed next to each other
under these conditions.
Firstly, efficiency curves were calculated and it was
shown that for zero reduced temperature the thermal
efficiency of the uncovered collector is 52% and the
thermal efficiency of the single cover sheet-and-tube
design is 58%, while the channel above PV design
typically has 65% thermal efficiency.
Next, a prototype PVT-collector was built for the single
cover sheet-and-tube collector. The measurements indicated that the model agreed with the experimental results
well within the range of the experimental data.
Finally, the annual yield was calculated for a domestic
hot water system. The best thermal annual yield is
provided by the channel-above-PV design, the two-cover
sheet-and-tube design and the insulated two-absorber
Table 8
Loss mechanisms affecting the annual electrical yield (values for
one cover sheet-and-tube PVT-collector)
Loss mechanism

Electrical
efficiency

Base case: one cover sheet-and-tube


No increased laminate temperature
No low irradiation
No oblique irradiation
No inverter losses
Ideal power-point tracking
No top cover reflection

0.066
0.068
0.074
0.078
0.087
0.089
0.097

~ sT out 2 T ind
qwater 5 mc

(A.1)

4
qsky,rad 5 Fsky g ssT topglass
2 T 4skyd

1 Fearth g ssT 4topglass 2 T a4d

(A.2)

qsky,conv 5 h windsT topglass 2 T ad


Nu wind k air
5 ]]]sT topglass 2 T ad
Lc

(A.3)

Nu wind 5 0.56sRa crit sin wd 0.25 1 0.13sRa 0.333 2 Ra 0.333


crit d
gbsT topglass 2 T adL c3
Ra 5 Pr ]]]]]]
n2
8
Ra crit 5 10 (Fujii and Imura, 1972)

(A.4)

g pv
4
4
qair,rad 5 ]]]]ssT PVglass
2 T topglass
d
g 1 pv 2 g pv

(A.5)

qair,conv 5 h airsT PVglass 2 T topglassd


Nu air k air
5 ]]sT PVglass 2 T topglassd
H

(A.6)

1708ssin 1.8wd 1.6


Nu air 5 1 1 1.44 1 2 ]]]]]
Ra cos w
3

1708
1 2 ]]]
Ra cos w

G FS

qba 5 h basT cell 2 T ad

Ra cos w
]]]
5830

0.333

21

(A.7)
(A.8)

268

H. A. Zondag et al. / Solar Energy 74 (2003) 253269

Table B.1
Temperature dependency of properties in the airwater vapour mixture
Property of saturated mixture

Least squares fit for property as a function of


temperature in 8C

Binary mass diffusion coefficient D


Kinematic viscosity n
Thermal conductivity k
Specific heat c
Density r
Water vapour mass fraction X
Latent heat of water vapour Q
Prandtl number Pr
Schmidt number Sc

D 5 1.5 3 10 207 T 1 2.2 3 10 205 m 2 s 21


n 5 8.1 3 10 208 T 1 1.35 3 10 205 m 2 s 21
k 5 1.0 3 10 204 T 1 2.45 3 10 202 W m 21 K 21
c 5 0.1T 2 2 3T 1 1057 J kg 21 K 21
r 5 2 5.3 3 10 203 T 1 1.297 kg m 23
X 5 4.76 3 10 203 T 3 exp [5.57 3 10 202 T ]
Q 5 2 2.43 3 10 03 T 1 2.50 3 10 06 J kg 21
Pr 5 2.48 3 10 205 T 2 2 1.30 3 10 203 T 1 0.732
Sc 5 2 3.23 3 10 204 T 1 0.612

k glass
qPVglass 5 ]]sT cell 2 T PVglassd
dPVglass
k glass
qtopglass 5 ]]sT topglass 2 T topglassd
dtopglass

(A.9)

(A.10)

In Eq. (A.7), the correlation found by Hollands (Duffie and


Beckman, 1991, p. 160), the 1 signs indicate that these
terms have the indicated value only if they are positive. If
the term between the brackets assumes a negative value
these terms become zero.

Appendix B. Evaporation loss


In order to calculated the evaporation loss, the diffusive
mass flow is determined by
X
Mvap 5 2 r DAB ] 5 r hAB DX.
y

(B.1)

Here, DX is the difference in mass fraction of the vapour


over the air layer, d is the width of the air layer and DAB is
the mass diffusion coefficient. Introducing the Sherwood
number, which is defined as
hABd
Sh ; ]]
DAB

(B.2)

Eq. (B.1) can be rewritten in the form


DAB DX
Mvap 5 Shr ]]
d

(B.3)

By multiplication with the latent heat of the vapour, Q latent ,


the evaporative heat flow qev ap across the air layer between
the PV and the cover is calculated from
DX
qevap 5 Shrvap DAB ]Q latent
d

(B.4)

The gradient in the vapour mass fraction, which drives


the evaporative heat flow, results from the temperature
difference between the PV and the glass cover, since it is

assumed that the vapour at both sides of the layer is fully


saturated. The Sherwood number Sh is determined from
the empirical relation

S D

Sc
Sh Nu ]
Pr

1/3

(B.5)

which is valid over the range 0.6 , Pr , 60 and 0.6 , Sc ,


3000. In this equation the Schmidt number and the Prandtl
number appear, which are defined as

n
Sc ; ]
DAB

n
Pr ; ]
a

(B.6)

Here, n represents the momentum diffusivity (viscosity)


and a the thermal diffusivity. Tabulated values of Q latent ,
DAB and rv ap can be found in various handbooks. The
quantities are evaluated for a mixture of vapour and air.
Firstly, the mass fraction of the vapour is determined and
then the mixture properties can be found. For all the
quantities involved the temperature dependence is taken
into account by fitting the data to a function over the
required temperature range. The fits are presented in Table
B.1. In the calculation, the equations are evaluated at the
average layer temperature.

References
Affolter, P., Haller, A., Ruoss, D., Toggweiler, P., 2000. A new
generation of hybrid solar collectorsabsorption and high
temperature behaviour evaluation of amorphous modules. In:
Proceedings of 16th European Photovoltaic Solar Energy
Conference, 15 May, Glasgow, UK.
Bakker, M., Zondag, H.A., van Helden, W.G.J., 2002. Demonstration on a dual flow photovoltaic / thermal combi panel. In:
Proceedings of PV in EuropeFrom PV Technology to Energy
Solutions Conference and Exhibition, October 711, Rome,
Italy.
Bazilian, M.D., Prasad, D., 2000. A holistic approach to photovoltaic / thermal / daylight (PV/ T / L) cogeneration. In: Proceedings of Eurosun 2000, Copenhagen, Denmark.
Bejan, A., 1993. Heat Transfer. Wiley, New York.

H. A. Zondag et al. / Solar Energy 74 (2003) 253269


Bergene, T., Lvvik, O.M., 1995. Model calculations on a flatplate solar heat collector with integrated solar cells. Solar
Energy 55 (6), 453462.
Cox, C.H., Raghuraman, P., 1985. Design considerations for flatplate photovoltaic / thermal collectors. Solar Energy 35 (3),
227241.
Duffie, J.A., Beckman, W.A., 1991. In: 2nd Edition. Solar Engineering of Thermal Processes. Wiley Interscience, New York.
Fujii, T., Imura, H., 1972. Natural convection heat transfer from a
plate with arbitrary inclination. Int. J. Heat Mass Transfer 15,
755767.
Hendrie, S.D., 1982. Photovoltaic / thermal Collector Development
ProgramFinal Report, Report MIT. Lincoln laboratory.
Krauter, S., 1993. Betriebsmodell der optischen, thermischen und
elektrischen Parameter von photovoltaischen Modulen. Berlin
University of Technology, Ph.D. Thesis.
Leenders, F., Schaap, A.B., van der Ree, B.C.G., van Helden,
W.G.J., 2000. Technology review on PV/ thermal concepts. In:
Proceedings of Eurosun 2000, Copenhagen, Denmark.
Platz, R., Fischer, D., Zufferey, M.-A., Anna Selvan, J.A., Haller,
A., Shah, A., 1997. Hybrid collectors using thin-film technology. In: Proceedings of 26th IEEE Photovoltaic Specialists
Conference, Anaheim.
Rockendorf, G., Sillman, R., Podlowski, L., Litzenburger, B.,
1999. PV-hybrid and thermo-electric-collectors. In: Proceedings
of ISES 1999 Solar World Congress, July 49, Jerusalem,
Israel.

269

Russell, T., Beall, J., Loferski, J.J., Roessler, B., Dobbins, R.,
Shewchun, J., Krikorian, J., Case, C., Doodlesack, G., Oates,
W., 1981. Combined photovoltaic / thermal collector panels of
improved design. In: Proceedings of IEEE Photovoltaic Specialists Conference 1981.
Sandnes, B., Rekstad, J., 2002. A photovoltaic / thermal (PV/ T)
collector with a polymer absorber plateexperimental study
and analytical model. Solar Energy 72 (1), 6373.
Srensen, B., 2001. Modelling of hybrid PV-thermal systems. In:
Proceedings of 17th European Photovoltaic Solar Energy
Conference, October 2226, Munich, Germany.
Tripanagnostopoulos, Y., Nousia, T., Yianoulis, P., 2002. Hybrid
photovoltaic / thermal solar systems. Solar Energy 72 (3), 217
234.
de Vries, D.W., 1998. Design of a PV/ Thermal Combi Panel.
Eindhoven University of Technology, Ph.D. Thesis.
Zondag, H.A., de Vries, D.W., van Steenhoven, A.A., van Helden,
W.G.J., van Zolingen, R.J.C., 1999. The efficiency of PV/ Tcombi energy production. In: Proceedings of ISES 1999 Solar
World Congress, July 49, Jerusalem Israel.
Zondag, H.A., de Vries, D.W., van Helden, W.G.J., van Zolingen,
R.J.C., van Steenhoven, A.A., 2002. The thermal and electrical
yield of a PV-Thermal collector. Solar Energy 72 (2), 113128.

Das könnte Ihnen auch gefallen