You are on page 1of 11

Solar Energy 81 (2007) 10141024 www.elsevier.

com/locate/solener

Design of a jet impingement cooling device for densely packed PV cells under high concentration
Anja Royne *, Christopher J. Dey
School of Physics A28, University of Sydney, NSW 2006, Australia Received 27 July 2005; received in revised form 20 November 2006; accepted 28 November 2006 Available online 17 January 2007 Communicated by: Associate Editor Claudio Estrada-Gasca

Abstract A cooling device based on jet impingement is proposed for cooling of densely packed photovoltaic cells under high concentration. The device consists of an array of jets where the cooling uid is drained around the sides in the direction normal to the surface. The local and average heat transfer as well as pressure drop of the device is measured. A model is proposed which predicts the pumping power required for a given average heat transfer coecient and device conguration. This model forms part of an optimised design process which is outlined. It is shown that the inherently nonuniform heat transfer distribution of jet arrays has little eect on the expected electrical performance of the PV array. 2006 Elsevier Ltd. All rights reserved.
Keywords: Concentrating PV; Cooling; Jet impingement; Pumping power; Nonuniform temperature

1. Introduction Concentrating photovoltaics is seen as one of the ways to reduce the cost of solar electricity because expensive solar cell area is replaced by less expensive concentrator material. However, under concentration the PV cells experience a high heat load that will reduce their eciency if not dissipated eciently. The progress towards more densely packed cells and higher concentrations has led to the search for innovative cooling options. Liquid jet impingement has been found to be a promising method for this purpose (Royne et al., 2005). This technology is widely used in areas such as the thermal treatment of metals, cooling of internal combustion engines, gas turbines and thermal control of high power density electronic devices. The ow structure and heat transfer characteristics of an array of impinging jets are highly dependent on the
Corresponding author. Present address: PGP, P.O. Box 1048-Blindern, 0316 Oslo, Norway. Tel.: +47 22856926; fax: +47 22855101. E-mail address: anja.royne@fys.uio.no (A. Royne). 0038-092X/$ - see front matter 2006 Elsevier Ltd. All rights reserved. doi:10.1016/j.solener.2006.11.015
*

following parameters: nozzle-to-plate spacing z/d, nozzle pitch s/d, nozzle geometry, Prandtl number and Reynolds number. Extensive literature reviews of these aspects of jet impingement are given by among others Martin (1977), Webb and Ma (1995) and Jambunathan et al. (1992) as well as Royne (2005). Following from literature (Womac et al., 1994; Garimella and Rice, 1995), submerged jets (issuing through liquid-lled cavity) were chosen in this study as opposed to free-surface jets (issuing through air). A nozzle-to-plate spacing of z/d % 34 is considered optimal in several studies (Garimella and Rice, 1995; Garrett and Webb, 1999). It is of utmost importance in high concentration photovoltaic systems to ensure a high average heat rate transfer across the entire surface. Because crossow can signicantly lower the overall heat transfer in jet impingement cooling (Garrett and Webb, 1999), an impingement based device for PV cooling should incorporate drainage of the cooling uid in a direction perpendicular to the impingement surface. Four possible back drainage congurations for were considered (Fig. 1). In setup (a), water enters into

A. Royne, C.J. Dey / Solar Energy 81 (2007) 10141024

1015

Nomenclature A C Cd d h I k L m n N Nu p P _ q Q rs Pr R R2 Re s area heat transfer correlation coecient discharge coecient nozzle diameter heat transfer coecient current thermal conductivity length Reynolds number dependence Prandtl number dependence number of nozzles Nusselt number pressure power heat ux per unit area volume ow rate series resistance Prandtl number thermal resistance correlation coecient Reynolds number nozzle pitch (distance between neighbouring nozzles in a square array) T V W X z temperature voltage pumping power optical concentration nozzle exit to impingement plate spacing

Subscripts avg average c cell fo foil heat heater net net electrical output oc open-circuit opt optimal sc short-circuit w water Greek letters g PV eciency m kinematic viscosity q density r uncertainty

Fig. 1. Example of jet congurations with drainage direction normal to the impingement surface: (a) side drainage, (b) central drainage, (c) drainage through channels in impingement plate and (d) drainage between plenum pipes.

a plenum chamber above an orice plate, through which the jets impinge onto the heated surface. The water then returns through an outlet cavity which encompasses the plenum chamber. Fig. 1b shows the opposite conguration, where the liquid is drained through a central outlet. If the number of nozzles required for the cooling module is found

to be large enough that eects of crossow become signicant, it may be necessary to have drainage exits distributed throughout the cooling device. Two such congurations are shown in Fig. 1c and d. The former consists of a thick orice plate through which long nozzles are drilled in a square conguration. Between the rows of

1016

A. Royne, C.J. Dey / Solar Energy 81 (2007) 10141024

nozzles, outlet pipes are drilled through the length of the plate, perpendicular to the nozzles. This was rst proposed for air jets by Huber and Viskanta (1994). A simplied version of this arrangement is shown in Fig. 1d. Here, water enters through parallel inlet pipes and impinges through holes along the bottom of the pipes. The water is drained through gaps between the pipes into an outlet chamber. The side drainage conguration in Fig. 1a was chosen for the rst prototype device in the present study. The experimental setup and results for these measurements are discussed in the following sections. The last part of the paper discusses a general model for optimising the design of a jet impingement cooling device, and how the inherent nonuniform heat transfer of these devices may inuence the performance of the PV array. 2. Experimental design and procedure A schematic overview of the jet testing unit is given in Fig. 2. The water ows through the inlet into a plenum chamber manufactured from a 21 mm long 20 mm wide square stainless steel tube, with a wall thickness of 1.2 mm and rounded corners. This tube was welded onto the top plate. A stainless steel orice plate was welded onto the bottom of the plenum chamber. The water is forced through the orice and impinges onto the heated surface. It then returns through a 25 29 mm2 return ow chamber, the outer walls of which consist of a 21 mm thick Perspex (trade name for polymethyl methacrylate, or PMMA) plate. The heater consists of a 31 mm 25 mm, 0.05 mm thick stainless steel foil. The foil was clamped and stretched tightly between two aluminium bus-bars. Heat losses from the bottom of the foil could be assumed to be negligible because the expected heat transfer coecient from foil to water was orders of magnitudes higher than that of natural convection and radiation under the foil. To make a water tight seal the foil was clamped between Perspex support pieces and the Perspex outlet chamber piece over a width

of about 1 mm on either side so that the resulting heater area is 29 mm 25 mm. The power dissipated in the outer 1 mm of the foil, which was clamped between the Perspex parts and not directly cooled by water, was assumed to lead to Perspex heating only, and not to have a signicant eect on the temperature measurements of the adjacent foil due to lateral conduction. The stainless steel foil temperature distribution was recorded using Thermographic Liquid Crystals (TLC) and a digital camera. The camera was placed below the jet testing unit and looked at the foil through a circular hole in the bakelite base. A sheet of R35C1WA TLC (Hallcrest Inc., 2004) was attached to the back of the heated foil. The TLC nominally turns from black to red at 35 C (in practice the colour change started at about 32 C) and then through the rest of the spectrum to blue at 36 C. The colour change was calibrated by impinging the unheated foil with water of a known temperature and recording the colour distribution of the foil. Because of the high heat transfer coecient associated with impingement, the foil could be assumed to have the same temperature as the impinging water. The bottom of the foil was illuminated by two LED lights, one on each side, through the Perspex support pieces in the direction normal to the cross-section shown in Fig. 2. For each series of measurements, a calibration picture was taken at 33.8 C, after which the camera and lighting were left undisturbed for the duration of the measurements. This ensured that the lighting and placement of the camera would remain identical for the calibration and the measurement pictures. Subsequently, the water temperature was turned down to 32.4 C and pictures were taken at increasing electrical power levels. The data from the thermal images were compared with the calibration picture using a procedure developed in Matlab software. The temperature was translated into a heat transfer coecient using h _ q ; T fo T w 1

Fig. 2. Schematic diagram of jet testing unit.

which assumes that all electrical power dissipated in the exposed foil is transferred to the liquid, and that the temperature of the water is the same as the inlet temperature everywhere. Finally, the heat transfer matrixes from all of the pictures were combined to produce an overall map of the heat transfer distribution for the given ow rate. Holes in the matrix due to the nite power intervals at which the pictures were taken were smoothed out using a linear interpolating algorithm. This resulting matrix was used to nd the maximum and average heat transfer coecients. The water was circulated and kept at a constant temperature using a circulating chiller. A manual ball valve at the water outlet was used to control the ow rate. The temperatures of the inlet and outlet water were measured by two platinum resistance thermometers in copper temperature pockets. The ow rate was recorded by a turbine rotor

A. Royne, C.J. Dey / Solar Energy 81 (2007) 10141024 Table 1 Overview of orice plates used in the experiments Device Nine/dense Nine/sparse Short/straight Long/straight Sharp-edged Countersunk Number of nozzles N 9 9 4 4 4 4 Nozzle diameter d (mm) 0.7 0.7 1.4 1.4 1.4 1.4 Nozzle-to-plate spacing z/d 7.14 7.14 3.57 3.57 3.57 3.57 Nozzle pitch s/d 7.14 10.0 7.14 7.14 7.14 7.14 Nozzle conguration

1017

ow transmitter, and pressure was recorded by a micro dierential pressure transducer placed between the inlet and outlet pipes of the jet testing unit. Six dierent orice plates were tested, shown in Table 1. Two had nine nozzles of small diameter, d = 0.7 mm and four had four nozzles with d = 1.4 mm. All of the arrays, except the nine/sparse array, had the same nozzle pitch to diameter ratio s/d = 7.14. The two-dimensional placement of the nozzles is shown in Fig. 3. The nozzle-to-plate spacing to diameter ratio was set to z/d = 3.57 for the fournozzle arrays and, because the distance between the orice plate and the impingement plate was kept constant at 5 mm, z/d = 7.14 for the nine-nozzle arrays. The thickness of the orice plates was 1 mm for all plates except for the one called long/straight which was 2 mm thick. The contouring of the sharp-edged and countersunk nozzles was made using a conventional 30 countersinking tool. From examining the rate of change of enthalpy in the water for a range of electrical power inputs, it was found that only 90% of the power was transferred to the water. This is taking into account any losses occurring in the pipes. By recording the temperature rise in the bus bars using thermocouples, it was found that the missing 10% were dissipated here, most likely due to contact resistance between the bus bars and the foil. The standard deviations of the ow meter, pressure transducer, PT100 temperature sensors and voltage measurements were all calculated from a range of measurements and found to be stable for measurements made on dierent days and under dierent conditions. However, because of a problem with the pressure transducer in the earlier series of measurements, only the pressure data for later measurements are included in this paper. The measured foil area was estimated to have an uncertainty of 0.5 mm in each direction which for an area of 25 mm 31 mm results in a

total uncertainty of 3.4%. The dominant uncertainties contributing to the uncertainty in havg are those of the foil area and the foil temperature. The TLC sheet did not yield a uniform colour when it was unheated and impinged with water of a known temperature because of variation in the sheet adhesive as it was supplied from the manufacturer. This was partly corrected by comparing each pixel individually against the calibration picture. The TLC colour sensitivity is very high and results in a maximum error of 0.2 C. However, the camera shutter had to be released manually, which made the camera move slightly. This meant the pixels were sometimes compared with calibration picture pixels with a slight displacement. In addition, the cameras automatic focus would sometimes focus at the bus-bars or the support instead of the TLC, so that the foil was sometimes out of focus. An analysis of the repeatability of the measurements showed that the total uncertainty in heat transfer coefcient was less than 8%. The uncertainty in the temperature distribution was inferred from this value to be about 7.2%. With the uncertainty r given in terms of a percentage of the mean value, the combined uncertainty in a variable y = f(x1, . . . , xi) is calculated from rX 2 r2 rtotal i:
i

The uncertainties in power and temperature dierence, and subsequently heat transfer coecient, are calculated in this manner. The resulting list of uncertainties is given in Table 2.

Table 2 Uncertainties of the measurements along with the associated combined uncertainties Measurement Foil area Foil temperature Flow rate Pressure Water temperature Foil voltage Shunt voltage Power Temperature dierence Heat transfer coecient Uncertainty (%) 3.4 7.2 4.2 0.6 0.1 0.1 0.1 3.4 7.2 8.0

Fig. 3. Two-dimensional placement of (a) nine/dense nozzles, (b) nine/ sparse and (c) four-nozzle arrays.

1018

A. Royne, C.J. Dey / Solar Energy 81 (2007) 10141024


nine/dense nine/sparse short/straight long/straight sharp-edged countersunk nine/dense correlation nine/sparse correlation short/straight correlation long/straight correlation sharp-edged correlation countersunk correlation

3. Results and discussion Fig. 4 shows a typical map of local heat transfer coecients under the countersunk array. As expected for impinging jets, the heat transfer coecient is high in the stagnation point directly underneath each nozzle, but drops o quickly with radial distance. The value of the heat transfer coecient in the centre of the array is less than half that at the stagnation points. Along the centrelines between the nozzles there is evidence of an interaction zone which leads to a slightly increased heat transfer coecient. This phenomenon has been described earlier by among others Pan and Webb (1994). A further discussion of the local heat transfer characteristics of the jet arrays under consideration can be found in Royne (2005). The dierent nozzle congurations were found to yield slightly dierent levels of heat transfer. Of the four-nozzle arrays, the sharp-edged nozzles were found to yield the highest heat transfer for a given ow rate (Royne and Dey, 2006). The average Nusselt numbers achieved are shown in Fig. 5. These values correspond to average heat transfer coecients in the range havg = 27 105 W m2 K1. The nine-nozzle arrays are seen to perform poorer than the four-nozzle arrays. This is partly because the former have a relatively larger heater area around the array which is cooled only by a highly weakened jet at a large distance from the stagnation point. These arrays also have a higher z/d than the latter, which could result in a lower local and average heat transfer coefcient (Garimella and Rice, 1995). Predictive correlations were made for the experimental data on the form Nuavg C Rem Prn : 3

120 100 80

Nuavg

60 40 20 0 0 2000 4000 6000 8000

Re
Fig. 5. Average Nusselt numbers versus Reynolds numbers along with correlations for the dierent orice plates.

Because the Prandtl number dependence was not investigated, the dependence n = 0.444 from the study of Li and Garimella (2001) was used. The coecients C and m for the average Nusselt number for the orices studied are given in Table 3 along with the correlation coecient R2, which gives a measure of how the correlation ts the data (R2 = 1 for a perfect t). The correlation values are shown with the experimental data in Fig. 5. Martin (1977) presented the following model for average Nusselt number under arrays of jets, which has been veried in several later studies (Huber and Viskanta, 1994; Garimella and Schroeder, 2001): Nuavg 0:5KGRe2=3 Pr0:42 ; where ( p6 )0:05  z =d f K 1 ; 0: 6 s p 1 2: 2 f p G2 f 1 0: 2 z =d 6 f 4

Table 3 Correlations for average Nusselt numbers for experimental data Orice Nine/dense Nine/sparse Short/straight Long/straight Sharp-edged Countersunk C 0.095 0.222 0.472 0.450 0.464 0.222 m 0.608 0.519 0.491 0.497 0.512 0.580 R2 0.983 0.979 0.963 0.978 0.958 0.994

Fig. 4. Heat transfer distribution for the sharp-edged orices at Re = 5380.

A. Royne, C.J. Dey / Solar Energy 81 (2007) 10141024


1000

1019

___Nuavg__

0.5KGPr0.42

nine/dense nine/sparse short/straight long/straight sharp-edged countersunk


100 1000 10000

Re
Fig. 6. Comparison of average Nusselt versus Reynolds number for nine and four nozzle jet arrays with the correlation from Martin (1977).

and  2 p d : f 4 s 7

4. Descriptive model of jet impingement cooling device 4.1. Correlation for pumping power The pumping power W required for any forced convection device is given as the product of ow rate and pressure drop W DpQ: 9

Fig. 6 shows a comparison of the experimental data with this correlation. There is good agreement, although the correlation tends to overestimate the values for the ninenozzle arrays and underestimate those of the four-nozzle arrays. The overestimation of the nine-nozzle arrays is probably due to the fact that the correlation is based on a square area with side lengths s around the central jet, while in the present study the area around the array is also included, with a larger area of low heat transfer taken into account. The fact that the orice plates of the four-nozzle arrays all had short developing length probably contributes to the underestimation of these values, because a short developing length yields a higher Nusselt number for a given Reynolds number (Garimella and Nenaydykh, 1996). The Martin correlation is made on the basis of several experimental studies with a range of nozzle congurations but has no correction for developing length. It is shown in Royne and Dey (2006) that the pressure drop through a jet impingement device is given by Dp qQ2 8
4 N 2 p2 C 2 dd

where Dp is the pressure drop, q is the coolant density, Q is the volumetric ow rate, N is the number of nozzles, Cd is the nozzle discharge coecient and d is the nozzle diameter. The discharge coecients obtained for the four-nozzle arrays in this study are listed in Table 4 (Royne and Dey, 2006).

The correlation for pressure drop (Eq. (8)) can be substituted directly into Eq. (9) to eliminate Dp. In the subsequent sections it will be assumed that Cd is independent of nozzle diameter. This is consistent with the theory from textbooks such as Walshaw and Jobson (1972), however some studies suggest that small diameter nozzles show a slightly dierent behaviour (Reader-Harris et al., 1995). The next objective is to eliminate the ow rate Q from the equation. This can be done by including the correlation for average heat transfer coecient havg in terms of Q, solving for Q, and substituting this into Eq. (9). It was decided to use two dierent models with dierent s/d dependence. The rst is the correlation given by Martin (1977). The second model, which will be referred to as the Huber model, has a better t to the experimental results from the current study. It incorporates the constant C and Reynolds number dependence m from the experimental data with the Prandtlnumber dependence from Li and Garimella (2001) and the s/d dependence from Huber and Viskanta (1994). As shown in Fig. 7, the two chosen models are qualitatively dierent with respect to their s/d dependence. The Martin model has a negative second derivative, while the Huber model has the opposite shape. This will be shown to result in quite dierent behaviours for the pumping power correlations made using these models. 4.1.1. The Martin model Rewritten in terms of havg and Q and solved for Q, the Martin model becomes  3=2 N pm 5=2 havg 0:42 d Pr Q : 10 4 0:5KGk

Table 4 Coecients of discharge for dierent nozzle congurations Nozzle conguration Short/straight Long/straight Sharp-edged Countersunk Cd 0.582 0.613 0.520 0.653

1020
80 70 60 50 Nu 40 30 20 10 0 4 5 6 s /d M artin Huber

A. Royne, C.J. Dey / Solar Energy 81 (2007) 10141024

For simplicity, in the following sections only square arrays of jets are considered so that the number of jets is restricted to N = n2 where n is an integer, and the heated surface is supposed to be square with sides Lheat. The jet pitch, s, is then given as a function of N and Lheat by Lheat s p : N 4.2. Model predictions
7 8

15

Fig. 7. Predicted Nusselt number as a function of s/d using the correlations from Martin (1977) and Huber and Viskanta (1994) for an arbitrary jet conguration.

Substituting (8) and (10) into (9) gives us a correlation for pumping power W:  9=2 N pm3 d 7=2 havg 0:42 Pr : 11 W q 0:5KGk 8C 2 d 4.1.2. The Huber model The second model, which incorporates the experimental correlations from this study with the Huber s/d dependence, is given by Nuavg C Rem Pr0:444 s=d
0:725

12

Rewriting in terms of havg and Q and solving for Q yields  1=m   havg d N pd m 0:725=m Q : 13 Pr0:444=m s=d Ck 4 The resulting pumping power correlation becomes " #3 1=m   8 havg d N pd m 0:725=m 0:444=m W q 2 2 2 4 s=d : Pr Ck 4 N p Cd d
14

Fig. 8 shows a major dierence between the predictions from the Martin model and the Huber model. While the former predicts a denite optimal nozzle diameter, dopt, for each set of conditions, the latter recommends always using the smallest possible nozzles. This discrepancy arises from the dierence in s/d dependency for the Martin and Huber correlations shown in Fig. 7. The existence of a dopt seems intuitively correct. When the nozzle diameter is decreased, the increase in jet velocity leads to a higher heat transfer coecient. This, however, comes at the cost of a highly increased pressure drop. Moreover, the heat transfer distribution drops o more rapidly away from the impingement point. At some specic diameter, the negative eects would be expected to become dominant and to lead to an increased pumping power for a given havg, as the Martin model predicts. Both models predict a lower pumping power for a higher number of nozzles, independent of other variables. This result is contrary to the conclusions from several studies that optimise against ow rate. Brevet et al. (2002) found that a decrease in the number of nozzles, and the subsequent increase in Reynolds number, would result in a higher havg. However, the pressure drop, which was not taken into account in this optimisation, increases drastically when the total orice cross-section is decreased. Garimella and Schroeder (2001) also found the heat transfer to be highest for a single jet when optimised for ow rate, but noted that the pressure drop, and thus the pumping power, would be lower for a high number of nozzles.

a
100

N=1 N=4 N=9 N=16

b
10

W (W)

W (W)

10

0.1

0.1 0 2 4 6

0.01 0 2 4 6
d (mm)

d (mm)

Fig. 8. Varying nozzle diameter for dierent numbers of nozzles (Lheat = 80 mm, havg = 10,000 W m2 K1) using (a) the Martin model and (b) the Huber model.

A. Royne, C.J. Dey / Solar Energy 81 (2007) 10141024

1021

When the jets are too closely spaced, the negative eects of jet interaction before impingement (San and Lai, 2001) become increasingly signicant. Beyond a certain s/d, the benet gained by adding nozzles is lost to increased jet interference. Garimella and Schroeder (2001) found the average heat transfer to level o for s/d < 4. The Huber and Martin models are also only valid down to s/d = 4 and s/d = 4.43, respectively. A spacing of s/d = 4 is therefore a reasonable lower limit for design purposes. The Huber model predicts the smaller nozzles to be superior under all conditions, and shows no dierence in trend for the dierent numbers of nozzles. The Martin model, on the other hand, predicts an optimum nozzle diameter which shifts towards smaller nozzles for increasing N. The latter makes sense because with fewer nozzles, a larger area has to be covered by each jet. Increasing the nozzle diameter makes the local heat transfer distribution fall o more slowly away from the stagnation point in terms of absolute distance, so that a larger area is covered by the central high heat transfer region. The optimal nozzle diameter, dopt, is found to be independent of havg, Cd and Pr. However it depends on N and Lheat as shown in Fig. 9. This gure also includes the required pumping power per area, W/A, for havg =

104 W m2 K1 to illustrate the large variation in pumping power for an increasing number of nozzles. Note that W/A also increases with the heater size. This reects back on the benet of a high number of nozzles per area. As N is kept constant but the area is increased, W/A increases as well. The optimal nozzle diameter is also dependent on the nozzle-to-plate spacing z/d as shown in Fig. 10. It can be seen that dopt decreases linearly with increasing z/d. The slope of the graph is dependent on N but independent of Lheat. In the Martin model, the nozzle-to-plate separation, z/d, has an eect in that it shifts the size of dopt. It also predicts a higher havg at low z/d. In the majority of the measurements presented here the parameter z/d was kept constant at z/d = 3.57 because several studies had observed the maximum havg to occur at 3 < z/d < 4 (Womac et al., 1993; Garimella and Rice, 1995). These studies generally found very little change in havg when reducing z/d below this value. However, other studies including Martin (1977) have found low z/d to be favourable, especially at low s/d. 4.3. Net PV output: determining the optimal operating conditions In Royne et al. (2005) a one-dimensional thermal model for typical silicon PV cells was presented which can be used to predict the electrical output from the cell as well as the cell temperature as a function of illumination level and cooling system thermal resistance, R, which is the inverse of the heat transfer coecient, h. This model can be used together with the correlation for pumping power as a function of havg presented in this paper to nd the optimal cooling performance for a given illumination level. Fig. 11 shows typical plots of the total and net (pumping power subtracted) PV output for dierent conditions. As expected, a low average heat transfer coecient results in a high cell temperature and a subsequent low cell output. The pumping power is slightly underestimated because only the mechanical, not electrical, power requirement is calculated. The graphs in Fig. 11 are based on an area of 50 mm 50 mm, and the parameters N = 4, d = 1.4 mm,
7

12 N=1 10 8 N=4 N=9 N = 16

d opt (mm)

6 4 2 0 0 20 40 60 80 100

Lheat (mm)

10000 1000

W/A (W m -2)

100 10 1 0.1 0 20 40 60 80 100 Lheat (mm)

6 5 Lheat = 100 mm Lheat = 80 mm 4 3 2 1 0 2 3 4 5 Lheat = 60 mm Lheat = 40 mm Lheat = 20 mm

Fig. 9. Optimal nozzle diameter as a function of heated area, with the corresponding power requirement per area for havg = 104 W m2 K1 using the Martin model. The results for dopt are independent of havg.

d opt (mm)

z /d
Fig. 10. dopt as a function of z/d for a range of Lheat.

1022

A. Royne, C.J. Dey / Solar Energy 81 (2007) 10141024

125 120 115

70 60 50

b
T (oC)

125 120 115

70 60 50 40 30 20 10 0 0 20000 40000 60000

P (W)

110 105 100 95 90

105 100 95 90 0 20000 40000

30 20 10 0 60000

h avg (W m -2 K-1)

h avg (W m -2 K-1)
140 120 100

c
P (W)

350 300 250 200 150 0 20000 40000

d
T (oC)

350 300

140 120 100 80 60 40 20 0 0 20000 40000 60000

P (W)

60 40 20 0 60000

250 200 150

h avg (W m -2 K-1)
Pc Pnet Tc

h avg (W m -2 K-1)

Fig. 11. Cell and total power output and cell temperature plotted for cooling system average heat transfer coecient for the two models and concentration levels: (a) 200 suns, Martin model; (b) 200 suns, Huber model; (c) 500 suns, Martin model; (d) 500 suns, Huber model.

Cd = 6.1, C = 1.96 and m = 0.491. The cell properties are given in Royne et al. (2005). From Fig. 11 it can be seen that there is a denite optimal, although broad, havg at which the net electrical output reaches a maximum. The predictions from both the Martin and the Huber models are shown for concentration levels for 200 and 500 suns. In this range there is not much dierence between the two models. The Huber model predicts a lower W for a given havg below havg = 42 103 W m2 K1 and a higher W above this level. For a concentration level of 200 suns, the Martin model gives the optimal havg to be 27 103 W m2 K1 while the Huber model predicts it to be havg = 28 103 W m2 K1. For 500 suns the same models predict the optimal havg to be found at 38 and 37 103 W m2 K1, respectively. 4.4. Guidelines for design optimisation The correlations developed in this paper can be employed to nd an optimal device design. The design process is outlined in the following steps: (1) Determine the size of the cooling unit, described by the length of a square area Lheat.

(2) Determine the number of nozzles, N. This should be made as high as possible while still ensuring s/d > 4 and avoiding negative crossow eects. If more than 4 4 nozzles are needed the water should be drained in a distributed fashion (see Fig. 1c and d). (3) Find a suitable nozzle-to-plate to diameter ratio, z/d. The optimum separation is disputed in literature, but is generally found somewhere in the range 2 < z/d < 4. A smaller separation distance might be preferable as the unit becomes less bulky. (4) Find the optimal nozzle diameter, d, as a function of Lheat, N and z/d. If s/d is found to be below 4 for this conguration, the nozzle diameter or number of nozzles should be reduced. Reducing d would be preferable as it has a smaller impact on W, however a lower limit to d can be set from manufacturing constraints and the possibility of blocking of nozzles from small particles or build-up. (5) Determine the nozzle conguration (Royne and Dey, 2006) and possible surface modications (Webb and Ma, 1995), and (6) Characterise the resulting device to nd the heat transfer correlation constants C and m (Eq. (12)), and then use Eq. (14) along with a thermal and

T ( C)

80

T (oC)
o

P (W)

110

40

A. Royne, C.J. Dey / Solar Energy 81 (2007) 10141024

1023

electrical model for the PV cells (Royne et al., 2005) to nd the optimal pumping power and the expected electrical output for a given illumination. 5. Eects of nonuniform temperature on PV output The heat transfer coecient distribution of impinging jets is inherently nonuniform (see Fig. 4), which will result in temperature variations across the photovoltaic array. This may have both electrical and mechanical implications. The latter refers to stresses resulting in the material from dierent thermal expansions across the surface, which may cause adhesion problems or cracking. Despite being an important problem, the mechanical aspects of PV arrays are beyond the scope of this paper and will therefore not be further addressed. The question of how nonuniform temperatures inuence the output from single and interconnected PV cells is investigated below. 5.1. Single cells Even in cases where the incident solar ux is perfectly uniform, temperature nonuniformities are always present in concentrator cells, as a result of imperfections in the cell-to-substrate bond or as a result of the cell and heat sink geometry. Two prior studies could be found that investigate the eect of temperature nonuniformities under uniform illumination. Sanderson et al. (1980) addressed the problem of hot areas caused by voids in the bond between the cells and the heat sink. A theoretical model was made in which the unit cell was divided into element areas, each operating at a unique temperature. For a step dierence in temperature with DT = 50 K, the predicted cell eciency was found to decrease from 13.1% to 12.2% at 100 suns and from 12.6% to 11.6% at 10 suns. More specically, the nonuniform temperature was found to result in a decreased open-circuit voltage. Mathur et al. (1984) used a slightly
0.19 0.185 0.18 0.175 (%) 0.17
series

dierent approach, subdividing a circular cell into concentric rings with a temperature gradient from the centre to the edge of the cell, which is a model of a circular cell bonded onto a large, at heat sink. It was found that with increasing thermal nonuniformities, the short-circuit current would increase, whereas the open-circuit voltage and conversion eciency would both decrease. However, both of the above studies indicate that the eects of temperature nonuniformities across one cell are relatively small. 5.2. Interconnected cells To obtain an understanding of how a temperature dierence between cells connected in a module aect the electrical output, it is sucient to study the behaviour of two cells at dierent temperatures connected in series and parallel. The electrical characteristics of the cells are found using a semi-empirical model for concentrator silicon solar cells presented by Mbewe et al. (1985). The open-circuit voltage and short-circuit current for a cell of area Ac (cm2) at a given temperature Tc (K) and concentration level X (suns) are given by V oc 1:25 and I sc 0:034Ac X 1 3 104 T c 300 : 17 0:63 0:06 log X Tc 300 16

When Voc and Isc are known, the current I can be calculated at any voltage level V < Voc iteratively from the transcendent function  ! V Irs V oc I I sc 1 exp ; 18 kT c =q where rs denotes the series resistance, which is assumed to be a cell property independence of temperature. The values rs = 0.02 X cm2, Ac = 1 cm and X = 100 suns are all chosen for the subsequent calculations.

0.165 0.16 0.155 0.15 0 20

parallel

40 T (K)

60

80

100

Fig. 12. Eciency and voltages in two silicon cells connected in series with a constant average temperature of Tavg = 320 K.

1024

A. Royne, C.J. Dey / Solar Energy 81 (2007) 10141024

Iteration was used to nd the eciencies of the interconnected cells at various DT. The average temperature was held constant. The results are shown in Fig. 12. The series connection remains unaected because the gain in power for one cell equals the loss in power from the other cell (Royne, 2005). On the contrary, the total eciency does decrease with temperature dierence for two cells connected in parallel, with the eciency being a relative 12% lower when DT = 100 K than when the cell temperatures are equal. This is due to a necessary decrease in voltage in order to maintain the same current through the two cells (Royne, 2005). However, as the change is reasonably small and a temperature dierential of 100 K is highly unlikely, it seems that a temperature dierential across a parallel connection should not be a major cause for concern in concentrating PV modules. In any case, the eect of temperature nonuniformity is much lower than that of average temperature. 6. Conclusion This paper explores the viability of jet impingement cooling devices for arrays of densely packed PV cells. A prototype jet device was designed where the liquid is drained in a direction normal to the heated surface around the edges of the central array. The local and average heat transfer coecient under arrays of four and nine jets with side drainage was measured, and the results were found to agree with results in the literature. Based on correlations from the literature and experimental results, a model was developed which predicts the pumping power required for an average heat transfer coecient for dierent device congurations. The model predicts that a higher number of nozzles per unit area will perform better than a lower number of nozzles. It also predicts an optimal nozzle diameter for a given area and number of nozzles. Combined with a model for cell performance at dierent temperatures, it was found that there exists a broad optimal operating region for any system of photovoltaic cells and cooling device at a given illumination level. A cooling system design procedure was presented. The nonuniform heat transfer coecient distribution, which is an inherent property of jet impingement devices, was shown to have only a weak eect on the electrical output of the photovoltaic system compared with the eect of changing the average temperature of the cells. Therefore, the cooling system should be optimised to yield a high average heat transfer coecient rather than a high level of uniformity. Acknowledgements The authors would like to thank Terry Pfeier, Phil Denniss and John Piggott for valuable technical assistance. Financial assistance from the Universal Surface and Solar Science Pty. Ltd. is also gratefully acknowledged.

References
Brevet, P., Dejeu, C., Dorignac, E., Jolly, M., Vullierme, J.J., 2002. Heat transfer to a row of impinging jets in consideration of optimization. International Journal of Heat and Mass Transfer 45 (20), 41914200. Garimella, S.V., Nenaydykh, B., 1996. Nozzle-geometry eects in liquid jet impingement heat transfer. International Journal of Heat and Mass Transfer 39 (14), 29152923. Garimella, S.V., Rice, R.A., 1995. Conned and submerged liquid jet impingement heat transfer. Journal of Heat Transfer 117, 871877. Garimella, S.V., Schroeder, V.P., 2001. Local heat transfer distributions in conned multiple air jet impingement. Journal of Electric Packaging 123, 165172. Garrett, K., Webb, B.W., 1999. The eect of drainage conguration on heat transfer under an impinging liquid jet array. Journal of Heat Transfer 121, 803810. Hallcrest Inc., 2004. Product information research and testing applications. Available from: <www.hallcrest.com>. Huber, A.M., Viskanta, R., 1994. Eect of jet-to-jet spacing on convective heat transfer on conned, impinging arrays of axisymmetric jets. International Journal of Heat and Mass Transfer 37 (18), 28592869. Jambunathan, K., Lai, E., Moss, M.A., Button, B.L., 1992. A review of heat transfer data for single circular jet impingement. International Journal of Heat and Fluid Flow 13 (2), 106115. Li, C.-Y., Garimella, S.V., 2001. Prandtl-number eects and generalized correlations for conned and submerged jet impingement. International Journal of Heat and Mass Transfer 44 (18), 34713480. Martin, H., 1977. Heat and mass transfer between impinging gas jets and solid surfaces. Advances in Heat Transfer 13, 160. Mathur, R.K., Mehrotra, D.R., Mittal, S., Dhariwal, S.R., 1984. Thermal non-uniformities in concentrator solar cells. Solar Cells 11, 175 188. Mbewe, D.J., Card, H.C., Card, D.C., 1985. A model of silicon solar cells for concentrator photovoltaic and photovoltaic thermal system design. Solar Energy 35 (3), 247258. Pan, Y., Webb, B.W., 1994. Visualization of local heat transfer under arrays of free-surface liquid jets. In: Proceedings of The 10th International Heat Transfer Conference, New York, pp. 7782. Reader-Harris, M.J., Sattary, J.A., Spearman, E.P., 1995. The orice plate discharge coecient equation further work. Flow Measurement and Instrumentation 6 (2), 101114. Royne, A., 2005. Cooling devices for densely packed, high concentration PV arrays. M.Sc. thesis, School of Physics, University of Sydney, Sydney, Australia. Royne, A., Dey, C., 2006. Eect of nozzle geometry on pressure drop and heat transfer in submerged jet arrays. International Journal of Heat and Mass Transfer 49, 800804. Royne, A., Dey, C.J., Mills, D.R., 2005. Cooling of photovoltaic cells under concentrated illumination: a critical review. Solar Energy Materials and Solar Cells 86 (4), 451483. San, J.-Y., Lai, M.-D., 2001. Optimum jet-to-jet spacing of heat transfer for staggered arrays of impinging air jets. International Journal of Heat and Mass Transfer 44 (21), 39974007. Sanderson, R.W., ODonnell, D.T., Backus, C.E., 1980. The eects of nonuniform illumination and temperature proles on silicon solar cells under concentrated sunlight. In: Proceedings of 14th IEEE Photovoltaic Specialists Conference, San Diego, pp. 431436. Walshaw, A.C., Jobson, D.A., 1972. Mechanics of Fluids, second ed. Longman Group Ltd., London. Webb, B.W., Ma, C.-F., 1995. Single-phase liquid jet impingement heat transfer. Advances in Heat Transfer 26, 105217. Womac, D.J., Ramadhyani, S., Incropera, F.P., 1993. Correlating equations for impingement cooling of small heat sources with single circular liquid jets. Journal of Heat Transfer 115, 106115. Womac, D.J., Incropera, F.P., Ramadhyani, S., 1994. Correlating equations for impingement cooling of small heat sources with multiple circular liquid jets. Journal of Heat Transfer 116, 482486.

You might also like