You are on page 1of 22

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/269390494

EXPERIMENTAL AND THEORETICAL ANALYSIS OF A MICROCHANNEL HEAT


EXCHANGER FOR HIGH CONCENTRATION PHOTOVOLTAIC CELLS

Conference Paper · June 2014

DOI: 10.1615/ICHMT.2014.IntSympConvHeatMassTransf.400

CITATION READ
S
1
211
2 authors:

Daduí C Guerrieri Carolina Cotta

Centro Federal de Educação Tecnológica Celso Suckow da Fonseca (CEFET/RJ) 25 Federal University of Rio de Janeiro
PUBLICATIONS 46 CITATIONS 82 PUBLICATIONS 451 CITATIONS

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Micropropulsion View project

UFRJ Collaboration View project


All content following this page was uploaded by Daduí C Guerrieri on 10 December 2014.

The user has requested enhancement of the downloaded file.


Proceedings of CONV-14: Int. Symp. on Convective Heat and Mass Transfer

June 8 – 13, 2014, Turkey

CONV-14 ―…..

EXPERIMENTAL AND THEORETICAL ANALYSIS OF A MICROCHANNEL HEAT


1 2,*
EXCHANGER
1 FOR HIGH CONCENTRATION PHOTOVOLTAIC CELLS
2

* Daduí C. Guerrieri and Carolina P. Naveira-Cotta


Mechanical Eng. Dept. - CEFET-RJ UnED Itaguaí, Rio de Janeiro, Brazil
Laboratory of Nano and Microfluidics and Micro-Systems - LabMEMS,
Mechanical Eng. Dept. - POLI & COPPE/UFRJ, Rio de Janeiro, Brazil
( Corresponding author: carolina@mecanica.coppe.ufrj.br)

ABSTRACT

HCPV (High Concentration Photovoltaic) systems operate more efficiently once the solar cell is kept
cooled by suitable heat sinks. Most of the commercial HCPV systems are based on passive air-cooling
systems, especially due to their simplicity and robustness, while here an alternative approach is
considered to remove heat more efficiently from the photovoltaic cell, replacing the passive air-cooled
system with an active water-cooled one, and also allowing the re-use of the removed heat in a secondary
circuit for other purposes, such as water desalination. In this context, this work presents the
manufacturing process, the experimental procedure and theoretical analysis of an optimized micro heat
exchanger for active cooling of a commercial HCPV module that concentrates 1200 suns, using water
as the working fluid. The optimized micro-heat exchanger, with 18 micro-channels of 400μm in width
and 945μm in depth each, was manufactured in copper using a CNC micro-milling machine and then
characterized in a 3D digital microscope. An experimental apparatus was assembled to analyze the
proposed micro-heat exchanger in both controlled and realistic situations, comprised in four distinct
experimental runs. A non-intrusive infrared camera thermography system has been employed to
measure the temperature at the external surface of the micro-heat exchanger substrate, while in the
theoretical study the commercial simulation platform COMSOL Multiphysics 4.2a was adopted,
allowing for critical comparisons of experimental and theoretical findings.

Keywords: HCPV systems, micro-heat exchanger, experimental analysis, infrared thermography.

INTRODUCTION

The solar energy market is growing rapidly, becoming increasingly more competitive especially in the
world sunnier regions. The high concentration photovoltaic technology (HCPV) is able to concentrate
the sunlight to intensities of 1000 suns or more, onto a small area of a solar photovoltaic cell. Since
smaller areas and thus reduced amount of photovoltaic material are required in such HCPV systems,
there is a significant reduction in the value of the final product compared to a non-concentrated
photovoltaic system for the same total electrical output. HCPV systems operate more efficiently once
the solar cell is kept cool enough by suitable heat sinks. Most of the commercial HCPV systems are
based on passive air-cooled heat sinks, especially due to their simplicity and robustness.

The main idea of this work is to present the numerical simulation and experimental analysis of an
alternative heat sink to remove heat more efficiently from a HCPV system, changing the passive air-
cooled system to an active water-cooled one, and also allowing the re-use of the rejected heat in a
secondary thermohydraulic circuit, such as for water desalination. Tuckerman and Pease [1981] in their
early work of 1981 studied the heat exchange enhancement in micro-channel heat sinks for electronic
devices of high thermal dissipation. At that time, the so-called VLSI (very-large-scale integrated)
CONV-14 ―…..

5 2
circuits could dissipate up to 100 W in small areas, and it was a general belief that the physical limit for
7 2
heat sinks was on the order of 2x10 W/m . Using micro-channels, the authors showed that it would be
possible to dissipate heat fluxes as high as 10 W/m . Royne et al. [2005] provided an essential review
work on cooling of concentrated photovoltaics. Parida et al. [2011] present a fairly complete revision of
photovoltaic systems and show some advantages in reducing the amount of silicon, as occurs in HCPV
systems, and also present some possible applications with integrated circuits to re-use the waste heat
from the solar panels. Kermani et al. [2009] and Escher et al. [2010] introduced novel manifold micro
heat sink concepts for the cooling of concentrated solar cells. Correa et al. [2013] proposed an
optimized micro-heat exchanger for active cooling of a commercial HCPV module that concentrates
1200 suns, using water as working fluid in a closed loop. This optimized micro-heat exchanger based on
micro-channels is comprised of 18 micro-channels with 400μm in width and 945μm in depth each. This
configuration is here analyzed more closely, both numerically and experimentally, in four different
situations: three of them with a lower power dissipation and the last one for a more realistic situation
dealing with a higher power dissipation. All the numerical analysis was performed using the
commercial CFD platform COMSOL Multiphysics 4.2a. The four cases analyzed correspond to
available experimental runs, when the temperatures at the external surface of the micro-heat exchanger
were measured through infrared thermography, allowing for validation of the proposed model and
solution methodology, as shown in what follows.

2
Figure 1a shows a typical module of the commercial HCPV system (Sunflower, ATS Pyron) that
contains 30 photovoltaic cells per panel. For each cell, using Fresnel lens, as schematically shown in
Figure 1b, the system is able to concentrate 1200 suns on the photovoltaic material area of 1cm , that
needs to be actively cooled for improving its photoelectric conversion efficiency.

Figure 1a. Sunflower HCPV system module (ATS Figure 1b. Schematic drawing of the Fresnel lens
Pyron, Inc.) concentration

MICRO-HEAT EXCHANGER FABRICATION

The micro-heat exchanger thermally optimized and proposed by Correa et al. [2013], schematically
shown in Figure 2, has a total of 18 channels with dimensions of 381 x 945 μm in cross section and
10 mm in length, made of copper and designed to work with concurrent flow. The geometric shape
of the distributor was not optimized in Correa et al. [2013], only proposed as a possibility for the
inlet and outlet of the cooling fluid, but later on found to be of secondary importance in the thermal
performance of the micro-system.
CONV-14 ―…..

Figure 2. Micro-heat exchanger proposed by Correa et al. [2013]

Due to manufacturing restrictions, slight modifications on the optimum design of ref.[6] had to be
introduced. As can be seen in Figure 3a., in the micro-heat exchanger here manufactured it was
used a pressure sealing on the surfaces with low roughness, instead of using a gland of the o-ring.
Due to this modification, it was possible to increase the length of the channel from 10 to 12 mm,
increasing the overall heat exchange area. Also, the channel’s width was changed from 381 μm to
400 μm due to the available micro-tools. The fabrication was performed on a micro-milling
machine with computerized numerical control (CNC), of Minitech Machinery, Inc.. The
manufacturing process was performed in two steps, starting with the micro-heat exchanger’s base
(Figure 3a), followed by the cover (Figure 3b). Figure 3a. presents the technical details of the 18
channels with a cross section of 400 x 945 μm and 12 mm in length, which leads to a hydraulic
diameter of 562 μm. The cover of the micro-heat exchanger contains the entry and exit ports of the
cooling fluid. Figure 3b presents the technical details with the dimensions of the micro heat
exchanger’s cover.

(a) (b)
Figure 3. Technical views of the micro-heat exchanger

The manufactured micro heat exchanger is shown in Figures 4, with the external dimensional
measurements. The internal dimensional measurements were obtained with the digital microscope
HIROX KH-8700. The widths of the micro-channels were measured at 30 different points, resulting
in an average size of 401.0 μm with a standard deviation of 3.6 μm. To measure the depth of the
channels it was necessary to generate a 3D image in the Hirox Digital Microscope. Using this
function, a total of nine measurements were performed at different points of the micro-channels,
leading to an average channel depth of 938.7 μm, with a standard deviation of 2.6 μm, which
represents a difference of only 6.3 μm in relation to the previously optimized depth of 945 μm.
CONV-14 ―…..

(a) (b) (c) (d)


Figure 4. External dimensional measurements of the micro-heat exchanger

EXPERIMENTAL ANALYSIS

An experimental workbench for the thermal analysis of the micro-heat exchanger was then assembled.
The experimental analysis was divided into two different setups, for either low power or high power
dissipation by the micro-heat exchanger. The low power analysis was used to characterize the micro-
heat exchanger in controlled conditions and different heat fluxes. The high power analysis was aimed at
studying the behavior and efficiency of the micro-heat exchanger in conditions closer to those to be
encountered when installed on the HCPV panel.

Figure 5 shows the configuration for the low power setup, which includes the following devices:
the infrared thermography camera, FLIR SC645; the data acquisition system (Agilent 34970-A); five
thermocouples; the acquisition computer; the syringe pump (New Era NE-1000); the precision scale
(Mars AS5500C); a DC current-controlled voltage source; the electric resistance (constantan skin
heater); and the micro-heat exchanger manufactured for this study.

Thermographic camera

DC Current Source

Syringe pump

Micro-heat exchanger

Acquisition

system

Scale

Figure 5. Experimental setup for low power dissipation


Figure 6 shows in detail the micro-heat exchanger already installed in the apparatus and how it was
assembled in this low power analysis setup. It can be seen that the micro-heat exchanger is painted
with a graphite ink that provides uniform emissivity of the surface to a known value of 0.97,
reported by the paint manufacturer (Graphit 33, Kontact Chemie).

For this low power characterization of the micro-heat exchanger, the employed skin heater has an
electrical resistance of 15.6 ohms, capable of dissipating power in the range of 2-20W. This
resistance has a limitation on its maximum operating temperature at 150°C. Thus, for the initial
characterization of the experimental apparatus, a maximum power of 12.6 W has been adopted.
CONV-14 ―…..

Figure 7 illustrates the hydraulic circuit and connections of the micro to the macro scales for the
low power experimental setup. The syringe pump conveys the energy required to transport the fluid
through the micro-heat exchanger and then the fluid is stored in a container on the precision scale,
so as to confirm the mass flow rate against the volumetric flow rate prescribed on the syringe pump
controls.

Inlet Micro-heat exchanger


Outlet

Thermocouples

Resistance
Resistence
Insulating wool

Acrylic
Rock wool

Figure 6. Instrumentation of micro-heat exchanger for characterization

Syringe pump

Micro-heat Container
exchanger

Precision Scale

Figure 7. Hydraulic circuit and connections scheme for the low power setup.
5

In order to analyze conditions closer to the operating conditions of the micro-heat exchanger once
installed in the actual HCPV cell, a high power setup was also assembled. Whereas the peak
irradiation on Fundão Island, Rio de Janeiro, Brazil, in the month of January, is of the order of 600
W/m², according to Correa et al. [2013], and considering a photovoltaic cell with concentration
factor of 1200 times, we may reach a heat flux of 7.2 x10 W/m² on this cell of 0.3 cm², which
represents a total power of 21.6 W to be dissipated. To achieve this power level within the
experimental setup, it was necessary to use an electrical resistance cartridge of 259.2 ohms and
maximum 46.6 W under AC voltage of 110 V, positioned inside a specially prepared copper device
to direct the heat flow to an area of 10 x 10 mm where the micro heat exchanger is positioned.

Figure 8 shows the configuration of the high power setup, which includes the following equipment:
the infrared thermography camera, FLIR SC645; the data acquisition system, Agilent 34970-A;
three thermocouples; the acquisition computer; a centrifugal pump, Rs-360sh; a type T connection;
a valve; the precision scale, Mars AS5500C; the controlled voltage source; the electrical resistance
cartridge with AC, 110v; the micro-heat exchanger.

The cartridge resistance was thermally insulated with fiberglass, TERMOVID 901, and the heat
transfer surface of 10 x 10 mm was covered with thermal compound to reduce contact resistance
with the micro-heat exchanger. Figure 9 shows a picture of the assembled structure, showing details
CONV-14 ―…..

of the hydraulic connections and of the acrylic plate fixed by four screws, positioned to control the
pressure on the contact surface between the cartridge resistance and the micro-heat exchanger.

Figure 10 shows the hydraulic scheme for the high power setup. The centrifugal pump moves the
fluid which is withdrawn from the reservoir, while part of the fluid returns to the reservoir and part
passes through the needle valve that controls the mass flow rate, and after passing through the heat
exchanger, is dumped on a container that sits on the precision scale.

Container

Thermography camera

Centrifugal pump

Micro-heat exchanger

Scale

Acquisition

DC Current Source system

Figure 8. Experimental setup for high power dissipation.

Micro-heat

exchanger

Cartridge

resistance

Fiberglass

insulation
Figure 9. Cartridge resistance and micro-heat exchanger assembly in high power setup.
CONV-14 ―…..

Needle Valve

Bypass

Reservoir

Container

Micro-heat

exchanger

Centrifugal Pump Precision

scale

Figure 10. Hydraulic scheme for the high power setup.

Table 1 summarizes the four cases of low and high power analyzed in this study, showing the
number of repetitions for each experiment, besides volumetric flow rates and dissipated power in
each case.
Table 1. Experimental cases
Case Repeatitions Flow rate (ml/min) Power (W)

Low Power 1 3x 10 6,6


2 3x 10 9,5

High Power 3 3x 10 12,6


4 3x 19,5 46,6

NUMERICAL ANALYSIS

Numerical simulations were performed for each of the cases presented in Table 1, making use of the
COMSOL Multiphysics 4.2a software, which is based on the finite element method. In the description
of the physical-mathematical problem, the conjugate problem package of the software was employed,
including the equations of continuity, Navier-Stokes and energy (both solid and fluid), considering a
three-dimensional transient problem of internal flow and heat transfer under the assumptions of
incompressible fluid, laminar regime, non-slip velocity boundary condition, continuity conditions of
heat flow and temperature at the solid-fluid interfaces and temperature dependent thermophysical
properties. For the input data, some experimentally determined values fed the proposed model, such as
room temperature, inlet fluid temperature and dissipated power in the electrical resistance. Judiciously
meshes were generated and a mesh convergence analysis was undertaken to ensure accurate results.

RESULTS

The inlet and outlet cooling fluid temperatures and the average external surface temperature evolutions,
as obtained from the thermocouple readings and infrared thermography recordings, respectively, are
now employed in the validation of the constructed model and solution methodology inherent to the
COMSOL platform. Figure 11 shows comparison the numerical results, in red, with the experimental
results, in blue, for the difference between the outlet and inlet water temperatures as obtained from the
thermocouple readings along the transient state. The thermocouple readings, after taking the repetitions
into consideration, present a standard deviation of 0.057 °C, 0.064 °C and 0.057°C for cases 1, 2 and 3,
respectively, and a standard deviation of 1.33°C for the high power case. The higher standard deviation
achieved in case 4 is due to the variation of the volumetric flow rate observed during the experiment,
with an average flow rate of 19.47 mL/min and a standard deviation of 2.08 mL/min, accounting for the
three repetitions. The average difference between experimental and numerical water outlet and inlet
temperature results in steady state, observed in Figures 11, is 0.18°C, 0.36°C and 0.72°C, respectively,
CONV-14 ―…..

for cases 1, 2 and 3, and 2.07°C for case 4, showing a good and consistent validation of the model and
numerical results obtained.
Next, the average external surface temperature in the cover of the exchanger was evaluated, both
from the thermographic images and from the COMSOL outputs, in the region of the microchannels,
which corresponds to an area of 10x10 mm. The pixel-to-pixel average based on the repetitions was
evaluated to estimate measurement uncertainty, yielding for cases 1 to 4, respectively, the
maximum combined standard uncertainties of 0.62, 1.00, 1.47 and 6.10 °C. Figure12 brings the
comparison of the average temperature on the10 x 10 mm surface area of the channels, along time.
The curves in blue represent the experimentally determined data through the infrared camera, while
the red curves represent the numerical simulation results. The average difference between
experimental and numerical temperatures is 0.39°C , 0.60°C and 0.10 °C, respectively, for cases 1,
2 and 3 , and 0.78°C for case 4. Again, the adherence between theoretical and experimental data is
indeed good.
Figures 13 shows steady state temperature results, for case 4, along the two central lines over the
10x10 mm surface, as presented in Figure 14, that cross transversally (Fig.13a) and longitudinally
(Fig.13b) the micro-heat exchanger. Figure 13a shows an average temperature difference between
the numerical and experimental results of 1.55°C and Figure 13b shows an average temperature
difference of 1.71 °C.

T ºC Case 3 T ºC

14

25
Case 2
12
Case 4
20

10
Case 1

15

10

0 20 40 60 80 100 t s 0 500 1000 1500 ts

0
40
Figure 11a. Comparison of the output and input water
temperature difference for cases 1, 2 and 3: numerical 38

(red) and experimental (blue) T ºC

T ºC 60
Case 3
Fi perimental (blue).
55 gu
re
11
b.
C
o
m
pa
ris
on
of
th
e
ou
tp
ut
an
d
in
pu
t
w
at
er
te
m
pe
ra
tu
re
di
ff
er
en
ce
fo
r
ca
se
4:
n
u
m
er
ic
al
(r
ed
)
an
d
e
x
Case 2 50

36
Case 4
Case 1 45

34

40

32

35

30

30

28 ts ts

0
20 40 60 80 100 120 140 0 500 1000 1500

Figure 12a. Evolution of average surface Figure 12b. Evolution of average surface
temperature of the cover for cases 1, 2 and 3: temperature of the cover for case 4: numerical (red)
numerical (red) and experimental (blue). and experimental (blue).
CONV-14 ―…..

T °C T °C

80 80

70 70

60 60

50 50

40 y mm 40 x mm
2 4 6 8 10 2 4 6 8 10
0 0

Figure 13a. Surface temperature along Figure13b. Surface temperature along transversal
longitudinal centerline for case 4: numerical (red) centerline for case 4: numerical (red) and
and experimental (blue) experimental (blue)

Figure 14. Typical Infrared image for case 4 with the two central lines that cross transversally (red-1
line) and longitudinally (green line) the external surface of the-1micro heat exchanger.

Table 2 shows some flow and heat transfer parameters calculated using experimental results for the
four different cases presented in Table 1. One may observe that the overall heat balance percentage
deviations were of the order of 15%-17% for cases 1 to 3, and increases to about 25% for case 4.
Furthermore, the micro-heat exchanger presents an area/volume ratio of microchannels of 7.1 m
and an area/volume ratio of the whole micro-heat exchanger of 7.7 m .

Table 2. Global analysis of micro-heat exchanger


Symbol Case 1 Case 2 Case 3 Case 4

Re 17.43+0.004 17.84+0.006 18.14+0.006 41.30+0.20


Reynolds number in the channel
ΔP (Pa) 23.47+0.005 19.10+0.007 18.78+0.007 39.50+0.19
Pressure drop in the channel

Hydrodynamic development length Lh (mm) 0.587+0.0001 0.601+0.0002 0.611+0.00


02
Thermal development length (mm) 3.151+0.0001 3.141+0.0001
3.134+0.00
Lth 01
1.39+0.006 6.15+0.0 04

Mean temperature between inlet and outlet T (°C) 30.26+0.01 31.37+0.02 32.20+0.02 39.22+0.28

(Tin+Tout)/2

Thermal load in the fluid Q (W) 4.70+0.01 6.80+0.02 9.18+0.02 31.40+0.79

Power dissipated in electrical resistance Qele(W) 6.6 9.5 12.6 46,6

Heat loss by natural convection Qair(W) 0.48+0.03 0.65+0.04 0.90+0.08 2.92+0.04

Heat loss by radiation Qrad (W) 0.39+0.05 0.40+0.06 0.41+0.13 0.55+0.05

Percentage deviation of the overall heat Perc (%) 15 17 17 25

balance

Thermal resistance of the heat exchanger Rth (°C/W) 0.897+0.05 0.958+0.05 0.981+0.07 1.140+0.008

Overall heat transfer coefficient U(W/°C.m²) 11145+618 10441+492 10193+756 8770+68.6

Asymptotic Nusselt number in the channel Nu 5.64+0.01 5.48+0.01 5.25+0.005 4.32+0.02
CONV-14 ―…..

CONCLUSIONS

A theoretical-experimental analysis of a manufactured micro-heat exchanger was performed,


considering the power to be dissipated and the mass flow rate as the major parameters to be
considered. Experimental and theoretical results were obtained for the temperature difference
between the output and input cooling fluid temperatures, as well as for the external surface average
temperatures, overall heat balance deviations, among others. The experimental and numerical
results were in excellent agreement and were satisfactory in characterizing the micro-heat
exchanger and validating the model and solution methodology employed in the simulations. The
natural sequence of this study shall be the installation of the micro-heat exchanger here produced
and analyzed, on a panel of the high concentration photovoltaic system to evaluate its performance
in loco.

REFERENCES

Correa, M.; Guerrieri, C. D.; Naveira-Cotta, C. P. and Colman, J. [2013], "Design and Manufacture
of Microchannel Heat Sinks for High Concentration Photovoltaic Cells", 22nd International
Congress of Mechanical Engineering (COBEM 2013), Ribeirão Preto, SP, Brazil, November 2013.

Escher, W., Michel, B., and Poulikakos, D. [2010], “A novel high performance, ultra thin heat sink
for electronics,” Int. Journal of Heat and Fluid Flow, vol. 31, nº 4, pp. 586-598.

Kermani, E., Dessiatoun, S., Shooshtari, A., and Ohadi, M.M. [2009], "Experimental Investigation
of Heat Transfer Performance of a Manifold Microchannel Heat Sink for Cooling of Concentrated
Solar Cells", 2009 Electronic Components and Technology Conference, IEEE. pp.453-459.

Parida, B., Iniyanb, S. and Goicc, R. [2011], "A Review of Solar Photovoltaic Technologies",
Renewable and Sustainable Energy Reviews, no. 15, pp. 1625-1636.

Royne, A., Dey, C.J., and Mills, D.R. [2005], “Cooling of photovoltaic cells under concentrated
illumination: a critical review,” Solar Energy Materials & Solar Cells, pp. 451-483.

Tuckerman, B. D. and Pease, R. F. W. [1981], “High-Performance Heat Sinking for VLSI”, IEEE
Electron Device Letters, vol. 2, nº 5, pp. 126-129.
View publication stats

You might also like