Вы находитесь на странице: 1из 8

Applied Thermal Engineering 30 (2010) 96103

Contents lists available at ScienceDirect

Applied Thermal Engineering


journal homepage: www.elsevier.com/locate/apthermeng

An experimental investigation of forced convective cooling performance of a microchannel heat sink with Al2O3/water nanouid
C.J. Ho *, L.C. Wei, Z.W. Li
Department of Mechanical Engineering, National Cheng Kung University, Tainan 70101, Taiwan, ROC

a r t i c l e

i n f o

a b s t r a c t
Experiments were conducted to investigate forced convective cooling performance of a copper microchannel heat sink with Al2O3/water nanouid as the coolant. The microchannel heat sink fabricated consists of 25 parallel rectangular microchannels of length 50 mm with a cross-sectional area of 283 lm in width by 800 lm in height for each microchannel. Hydraulic and thermal performances of the nanouidcooled microchannel heat sink have been assessed from the results obtained for the friction factor, the pumping power, the averaged heat transfer coefcient, the thermal resistance, and the maximum wall temperature, with the Reynolds number ranging from 226 to 1676. Results show that the nanouidcooled heat sink outperforms the water-cooled one, having signicantly higher average heat transfer coefcient and thereby markedly lower thermal resistance and wall temperature at high pumping power, in particular. Despite the marked increase in dynamic viscosity due to dispersing the alumina nanoparticles in water, the friction factor for the nanouid-cooled heat sink was found slightly increased only. 2009 Elsevier Ltd. All rights reserved.

Article history: Received 20 August 2008 Accepted 8 July 2009 Available online 14 July 2009 Keywords: Nanouid Microchannel Heat sink Forced convection Suspension

1. Introduction The major industries, such as aerospace, automotive, and electronic, are driving the development of compact and efcient thermal management technology for advanced electronic devices with increasingly high speed and high power density. For the two decades following the pioneering work of Tuckerman and Pease [1], microchannel heat sinks featuring a series of plate n structure on metallic or silicon substrates have been studied extensively, as indicated in the recent thorough reviews [24]. In particular, single-phase liquid-cooled microchannel heat sinks hold the great potential as the robust and effective cooling devices for removing high heat ux well beyond air-cooling limits, as demonstrated in the representative studies [57]. The conception of nanouid, formulated by dispersing metallic or non-metallic nanometer-size particles in base liquids such as water and ethylene glycol, was proposed rst by Choi [8], displaying an anomalously high thermal conductivity that classical theory of effective media fails to explain. Ever since there have been great research interests in exploring the effectiveness and feasibility of using nanouids as convective heat transfer uids. Several possible mechanisms for enhancement of heat transfer in nanouids have been postulated including thermal conductivity enhancement, Brownian motion, thermophoresis, diffusionphoresis, and so on. Excellent biographical reviews of the progress in exploring the heat
* Corresponding author. Tel.: +886 6 2757575x62146; fax: +886 6 2352973. E-mail address: cjho@mail.ncku.edu.tw (C.J. Ho). 1359-4311/$ - see front matter 2009 Elsevier Ltd. All rights reserved. doi:10.1016/j.applthermaleng.2009.07.003

transfer characteristics of various formulations of nanouid are available in Refs. [9,10]. To enhance the thermal conductivity of working liquids and thereby further improve performance of the liquid-cooled microchannel heat sinks, the use of nanouid has recently been considered and relatively little research effort in this aspect has been undertaken [1115]. For laminar fully developed ow of copper water nanouid in a silicon microchannel heat sink, Chein and Huang [11] presented a theoretical analysis based on empirical correlation, predicting signicant heat transfer enhancement due to the enhanced thermal conductivity and thermal dispersion effects. Koo and Kleinstreuer [12] numerically simulated the conjugate heat transfer characteristics of microchannel heat sinks using copper oxidewater or ethylene glycol nanouids. Effects of the viscous dissipation and the Brownian motion of nanoparticles in the base liquids were taken into account. Moreover, in a numerical study of laminar, hydrodynamically and thermally fully developed ows of copper and diamondwater nanouids [13], the cooling performance of the microchannel heat sink was found signicantly improved as reected by the marked reduction in the thermal resistance as well as the temperature difference between the heated microchannel wall and the nanouids. Experimentally, Lee and Mudawar [14] assessed cooling effectiveness of using small concentrations of aluminawater nanouid in a microchannel heat sink subject to uniform heat ux condition. Both singleand two-phase convective heat transfer results were obtained. The enhanced thermal conductivity associated with the presence of nanoparticles in water was found to promote the single-phase

C.J. Ho et al. / Applied Thermal Engineering 30 (2010) 96103

97

Nomenclature cp Dh Hch  h k Lch m _ m N Nu Dp _ Q qf Pe Re Ritd Rlm T DT um specic heat, kJ/kg K hydraulic diameter, m, 2WchLch/(Wch + Lch) height of microchannel, lm average heat transfer coefcient, W/m2 K thermal conductivity, W/m K length of microchannel, mm n parameter mass ow rate, g/min number of microchannel  =k average Nusselt number, hD h nf pressure drop, kPa volumetric ow rate, cm3/min heat transfer rate removed by uid, W Peclet number, qnfcp,nfumDh/knf Reynolds number, qnfumDh/lnf inlet temperature difference thermal resistance, K/W log-mean temperature difference thermal resistance, K/ W temperature, K temperature difference, K mean velocity, m/s Wch width of a unit microchannel, mm

Greek symbols temperature control effectiveness n efciency dynamic viscosity, N s/m2 density, g/cm3 / volumetric fraction of nanoparticles

eT w gn l q

Subscripts bf base uid c cover ch microchannel n n in inlet lm log-mean quantity m arithmetic-mean quantity nf nanouid out outlet tc thermocouple w base wall

heat transfer coefcient for laminar ow regime, mostly in the entrance region of the microchannels; while the accompanied decrease in specic heat results in larger axial temperature rise along the microchannel. Overall merit of using the nanouids in microchannel heat sink was concluded practically questionable for the two-phase cooling, in particular. Meanwhile, in an experiment [15] using copper oxidewater nanouid in a silicon trapezoidal microchannel heat sink, signicantly improved cooling performance was found only when the ow rate was kept lower than 15 ml/min. Among the foregoing predictions and experiments, there apparently exists disparity concerning the efcacy of using nanouids for better cooling performance of a uniformly heated microchannel heat sink. In this study, forced convection experiments for a microchannel heat sink were performed to provide supplementary heat transfer data concerning the efcacy of using aluminawater nanouid as the coolant against that of using the pure water.

of 283 lm in width (Wch) and 800 lm in height (Hch). The inlet and outlet pendulums were fabricated at two ends of the microchannels to provide relative uniform ow distribution, where two T-type thermocouples and pressure taps were positioned to measure the temperature rise and pressure drop across the microchannel heat sink, respectively. Six small holes were drilled along the centerline of the heat sink base, where T-type thermocouples were installed to measure temperature at a distance of 6 mm beneath the base surfaces of the microchannels. Heat input was provided by means of a plate heater powered by a DC power supply. The power supplied was determined using the measured voltage and current supplied to the heater. To minimize the heat loss, a guard heater is installed parallel to the rear surface of the main heater to ensure negligible temperature gradient between the two heaters. All the measured quantities were logged by a data acquisition system.

3. Preparation and properties of nanouids 2. Experimental setup A schematic of the main components of the close loop experimental facility constructed in the present study is shown in Fig. 1. The working uid enters the loop from a reservoir through a lter and is continuously circulated by a gear pump. A constant temperature bath installed upstream of the test module controls the inlet ow temperature. Exiting from the test module, the uid passes through another constant temperature bath to restore its temperature before returning to the reservoir. Volumetric ow rate inside the loop was monitored by a ow meter. The conguration of the test module fabricated is illustrated in Fig. 2. The test module consists mainly of a microchannel heat sink, housing, a cover plate, and two plate heaters. The geometric structure of the heat sink fabricated is depicted schematically in Fig. 3 with the detailed dimensions summarized in Table 1. Twenty-four parallel rectangular microchannels were machined into an oxygenfree copper block to form the microchannel heat sink. The microchannels are equidistantly spaced with a n width of 300 lm and each has a length (Lch) of 50 mm with a cross-sectional area The nanometer-sized particles of alumina (Al2O3) (Nanotech, Kanto Chemical Co. Inc., Japan) with an averaged particle size about 33 nm and 99.95% purity were dispersed in ultra-pure Milli-Q water (the base uid) to form the aluminawater nanouid. Nanouid of the desired volume fraction of alumina was formulated by mixing appropriate quantities of nanoparticles with the base uid in a ask and then dispersing in an ultrasonic vibration bath for at least 2 h. From a preliminary sedimentation test for the nanouids formulated under various pH conditions, it was found that the nanoparticles dispersed well in the base uids at pH = 3. Two volumetric fractions of the aluminawater nanouid, / = 1 and 2 vol.%, were prepared for the experiment. The volumemean diameters of the alumina particles in the nanouids formulated with 1 and 2 vol.% of alumina were measured by means of laser diffraction technique to be 116.3 mm and 119.3 mm, respectively. The thermophysical properties of the aluminawater nanouid formulated pertinent to the present experiment include the density qnf, the specic heat cp,nf, the thermal conductivity knf, and the dynamic viscosity lnf, for which the following formulae in

98

C.J. Ho et al. / Applied Thermal Engineering 30 (2010) 96103

Fig. 1. Schematic of the experimental test facility.

Fig. 2. Conguration for test module.

terms of the corresponding properties of the base uid and the nanoparticle were adopted except the dynamic viscosity: Density:

Thermal conductivity:

qnf 1 /qbf /qnp

knf 2 knp =kbf 2/knp =kbf 1 kbf 2 knp =kbf /knp =kbf 1

C.J. Ho et al. / Applied Thermal Engineering 30 (2010) 96103

99

x z Lch Wfin Wch

PMMA cover

Dc Hch

eld) was conducted in the present study at various temperatures from 20 C to 40 C and the results are listed in Table 3. Comparing with the results for the pure water (/ = 0 vol.%), the dispersion of alumina nanoparticles in water gives rise to a nonlinear increase in the dynamic viscosity with the volume fraction /. For instance, at temperature of 30 C, an increase of around 26% in the dynamic viscosity was found for the nanouid of 2 vol.%, while only 4.8% for the that of 1 vol.%. 4. Data reduction In the present study, forced convection heat transfer experiments have been undertaken for the microchannel heat sink using the pure water (/ = 0 vol.%) or the nanouid formulated (/ = 1 and 2 vol.%) as the working uid under the following operating condi_ 1151026 cm3 =min, the Reynolds tions: the volume ow rate Q number Re = 2261676, and the inlet temperature Tin = 31.1 32.6 C. During the experiments for the nanouid containing 2 vol.% of alumina, the nanoparticles driven by the intensied inlet contraction ow as the ow rate was increased tends to clog the inlet passages into some of the parallel microchannels so that the experimental results obtained were limited for the ow rate up to 663 cm3/min, indicative of occurrence of agglomeration/clustering of the suspended nanoparticles observed in the Ref. [14]. Experimental results for the temperature and pressure drop at a given ow rate generally reached steady-state after approximately 3060 min. The steady-state quantities measured in the experi_ ), the pressure ment primarily include the volumetric ow rate (Q drop across the test module (Dpmeasured), the temperatures of thermocouples embedded in the base block of the heat sink (Ttc), as well as the inlet and outlet uid temperatures (Tin and Tout). The measured pressure drop Dpmeasured consists of the net pressure drop across the microchannels, Dp, and the pressure drops at the inlet contraction and outlet expansion. The net pressure drop over the microchannels is thus calculated as

Dbase

q"
Fig. 3. Geometric conguration of copper microchannel heat sink.

Table 1 Dimensions of the microchannel heat sink. Dc (lm) 10 Dbase (mm) 24 Hch (lm) 800 Lch (mm) 50 Wch (lm) 283 Wn (lm) 300

Specic heat:

cp;nf 1 /cp;bf /cp;np 1 1 /qbf cp;bf /qnp cp;np cp;nf

3a 3b

qnf

Here for the specic heat of nanouid, two formulas Eqs. (3a) and (3b), which were, respectively, used in the Refs. [1315], were adopted to elucidate effect due to uncertainty in the relative change in the specic heat of nanouid on its heat transfer efcacy as illustrated in a recent analysis [16]. The foregoing thermophysical properties for the nanoparticle (alumina), base uid (water), and the aluminawater nanouid of two volume fractions formulated are listed in Table 2. For the aluminawater nanouid of two volume fractions formulated, both density and thermal conductivity appear signicantly increased compared to the pure water. The nanouid of 2 vol.% alumina, for instance, has a relative increase of 5.2% and 5.4% in the density and thermal conductivity, respectively, compared with water. Another fact worthy of attention is that the two formulas Eqs. (3a) and (3b) adopted for the specic heat of nanouid give signicantly different results but all markedly lower than that of the water. That is, for the aluminawater nanouid formulated, the specic heat evaluated using Eq. (3a) appears increasingly higher than that using (3b) as the volume fraction of nanoparticles increases. For the nanouid of 2 vol.% alumina, for instance, the former predicts a relative decrease of 1.4% while the latter predicts a relative decrease of 5.6% in the specic heat with respect to the pure water. To further exploit the possible inuence due to the disparity in the specic heat, both formulas (Eqs. (3a) and (3b)) were considered for the heat transfer data reduction in the present experiment and the corresponding results obtained are, respectively, identied as A and B hereafter. As for the dynamic viscosity of the nanouid formulated, direct measurement using a viscometer (Model: LVDV-II + Pro, BrookTable 2 Thermophysical properties of nanoparticle, base uid, and nanouid at 30 C. Properties Nanoparticle (alumina) 3600 0.765 36.0 Base uid (water) 995.1 4.178 0.620 Nanouid (aluminawater) / = 1 vol.% 1021.1 4.144 (Eq. (3a)) 4.058 (Eq. (3b)) 0.635 2 vol.% 1047.2 4.110 (Eq. (3a)) 3.943 (Eq. (3b)) 0.654

Dp Dpmeasured K c K e

qu2 m
2

where Kc and Ke are the loss coefcients for the inlet contraction and outlet expansion, respectively, and were estimated following the method described in [14,17]. The hydraulic performance of the heat sink can be evaluated by means of the friction factor dened as

DpDh 2qnf u2 m Lch

where Dh is the hydraulic diameter of a single microchannel and has a value of 418 lm for the heat sink fabricated. The mean uid veloc_ based on the ity um was calculated from the volume ow rate Q _ cross-sectional area of a single microchannel as Q =NW ch Hch : Thereby, the corresponding Reynolds number is dened as

Re

qnf um Dh lnf

Here the properties of the base uid involved in evaluating the properties of nanouid were evaluated based a mean uid temperature of Tm(=(Tin + Tout)/2).

Table 3 Measured dynamic viscosity of aluminawater nanouid. / (vol.%)

lnf 103 (N s/m2)


T = 20 C 25 C 0.8550 0.8955 1.0750 30 C 0.7690 0.8062 0.9689 35 C 0.6950 0.7483 0.8750 40 C 0.6310 0.6507 0.7947 0.9590 1.0040 1.0930

lnf T 20 lnf T 40

C C

q (kg/m )
cp (kJ/kg K) k (W/m K)

0 1 2

1.5198 1.5430 1.3754

100

C.J. Ho et al. / Applied Thermal Engineering 30 (2010) 96103

The measured temperatures at the six locations underneath the base wall of the microchannels were extrapolated assuming onedimensional heat conduction between the plane of the thermocouple inside the base block and the plane containing the microchannel base wall as

T w T tc

qf Htc Nks Lch W ch W fin

where Htc denotes the distance of the thermocouple embedded beneath the base surface of the microchannel and ks is the thermal conductivity of the base block. Furthermore, the average wall temperature T w was calculated by averaging these local wall temperatures extrapolated using Eq. (7), among which rather little spatial variation (typically less than 2 C) was detected except at the locations near the inlet and outlet of the microchannels. As a result, the microchannel heat sink in the present experiment was treated as that cooled under the uniform wall temperature condition, in contrast to that cooled under the uniform heat ux condition in the Refs. [14,15]. Moreover, the steady-state heat transfer rate removed by the nanouid owing through the heat sink, qf, can be determined from an energy balance as

Uncertainties in the measured quantities for the present study were estimated to be 0.3 C in the temperature, 2% in the ow rate, 0.5% in the pressure drop. Following the uncertainty propagation analysis, the uncertainties for the deducted experimental results were estimated as follows: 3.85.8% for the friction factor; 3.84.0% for the Reynolds number; 5.140.8% for the average heat transfer coefcient; 4.741.0% for the average Nusselt number; 4.722.9% and 5.141.0% for the inlet and log-mean temperature difference thermal resistances, respectively. 5. Correlation analysis for heat transfer efcacy of nanouids In what follows, how the heat transfer efcacy of a nanouid in a microchannel heat sink relates with its physical properties is deducted by a correlation analysis. Consider simultaneously developing laminar forced convection in a single channel with constant wall temperature. The average Nusselt number for the combined developing regime can be evaluated by a correlation due to Seider and Tate [18] as

 1=3  0:14 RePr l Nu 1:86 Lch =Dh ls

15

_ cp;nf T out T in qf qnf Q

The difference between the measured heat input power and the heat removed by the uid is denoted as the heat loss due to conduction losses through the housing and insulation, which was found to be within 13% of the heat input for the experiments conducted. Moreover, the repeatability of the experiments were checked by performing duplicate tests for certain operating conditions; and the measured pressure drop and temperatures were found virtually identical within the measurement uncertainties. The thermal performance of the heat sink is characterized by the thermal resistance, Ritd, based on difference between the average base wall temperature and the inlet uid temperature as

A relation of the average heat transfer coefcient with the relevant thermophysical properties of the uid as well as the characteristic lengths of the channel can then be expressed as
=3 2=3 1 1=3 $m _ 1=3 c 1 Dh Lch h p k

 0:14

l ls

16

Ritd

T w T in qf

Alternatively, another thermal resistance Rlm can be evaluated based on the log-mean temperature difference DTlm as

Rlm

DT lm qf

10

The DTlm takes the form of

DT lm

T w T in T w T out lnT w T in =T w T out

11

On the other hand, another quantity of practical interest for thermal characterization of a microchannel heat sink is the average heat transfer coefcient which can be determined from conventional n analysis giving:

 h

qf NDT lm W ch 2gfin Hch Lch

12

where gn denotes the n efciency and takes the form of

gfin

q =ks W . Based on Here the n parameter m is given by m 2h fin the foregoing average heat transfer coefcient, the corresponding average Nusselt number for the heat sink is dened as

tanhmHch mHch

13

_ denotes the mass ow rate. In effect, as indicated in _ qQ Here m Eq. (16), the average heat transfer coefcient appears inversely proportional to the hydraulic diameter of the channel, which is coherent with the very impetus for the commonly adopted heat transfer enhancement strategy of reducing the channel size to tens of micrometers. Moreover, among the pertinent thermophysical properties appeared in Eq. (16), the relative increase in the thermal conductivity of the nanouid with respect to its base uid can be seen to act predominantly as a benecial factor, leading to an enhancement in the average heat transfer coefcient. Meanwhile, it can be inferred that the decrease in the specic heat of the nanouid relative to its base uid as indicated in Table 2 might play a counteracting role to the benecial inuence due to the enhanced thermal conductivity on cooling effectiveness of the heat sink. Such negative role that the specic heat could play can also be inferred readily from the energy balance equation Eq. (8), from which under a condition of xed mass ow rate and temperature difference between outlet and inlet uid, the heat transfer rate carried by the uid ow is proportional to the specic heat. Moreover, the ratio of dynamic viscosity (l/ls) in Eq. (16) is to account for the variation of viscosity with a large wall-to-uid temperature difference, which can be expected in particular relevant in high-power heat sinks. Based on the data for the ratio of the dynamic viscosity at 20 C to that at 40 C tabulated in Table 3, the relatively lesser variation of dynamic viscosity with temperature for the nanouid of 2 vol.%, in particular, than that for the base uid might further contribute negatively to heat transfer efcacy of using nanouid. In contrast, for the nanouid of 1 vol.%, the viscosity variation with temperature appears very close to that for the base uid, thus exerting negligible inuence on the average heat transfer coefcient. 6. Experimental results First of all, experimental results of the friction factor obtained for the microchannel heat sink using the pure water (/ = 0 vol.%) as well as the nanouid (/ = 1 and 2 vol.%) as the coolant are presented as illustrated in Fig. 4 along with the theoretical result for

Nu

 hD h knf

14

C.J. Ho et al. / Applied Thermal Engineering 30 (2010) 96103

101

100

(%) 0 1 2

Present work

Lee & Mudawar [14]

25 20

(Vol.%) 0 1 2

Symbols A B

15

10-1

|
16/Re

Nu
10 5 200

10

-2

Re

500

1000

15002000

500

Re

1000

1500

2000

Fig. 4. Variation of the friction factor for the nanouid with the Reynolds number.

Fig. 5. Variation of the average Nusselt number with the Reynolds number. (Symbols A and B, respectively, based on specic heat from Eqs. (3a) and (3b).)

the fully-developed friction factor f = 16/Re. Also included in Fig. 4 are the experimental data reported by Lee and Mudawar [14]. As a benchmark for those obtained using the nanouid, the data obtained for the pure water can be seen to agree favorably with the corresponding results in the Ref. [14]. Specically, the increasing departure of the measured friction factor from the fully-developed value with the increase of the Reynolds number reects clearly the increasing inuence of hydrodynamically developing ow through the microchannels. For the nanouid of / = 1 or 2 vol.%, the calculated friction factor, similar to that of [14], displays a monotonic drop with the Reynolds number similar to that of base uid (water). Moreover, despite its markedly enhanced dynamic viscosity due to the presence of the nanoparticles in water, the nanouid of 1 or 2 vol.% alumina owing through the heat sink appears to give rise to only slight increases in the friction factor. Similar nding was reported in the correlation analysis [11] and the previous experiments [14,15] for the microchannel heat sink cooled by various nanouids. Next, the thermal performance of using nanouid in the microchannel heat sink is examined by plotting results of the average Nusselt number at various Reynolds numbers as shown in Fig. 5, in which two sets of data for the nanouid based on the different specic heat evaluated using Eqs. (3a) and (3b) are shown and designated by A and B, respectively. Qualitative variations of the average Nusselt number for the nanouids of two volume fractions with the Reynolds number can be seen somewhat unaffected by the difference in the specic heat used for the data reduction. From a closer scrutiny of these two sets of data in Fig. 5, one can nevertheless notice that the average Nusselt numbers based on the higher specic heat (designated by A) are appreciably higher than those based on lower specic heat (designated by B), further reecting their dependence on the specic heat as inferred in Eq. (15). Above all, comparison with the data for the pure water (/ = 0 vol.%) also shown in Fig. 5 clearly reveals signicant enhancement in the Nusselt number based on either specic heat that can be obtained by dispersing alumina nanoparticles in water. Specically, based on the higher specic heat, the average Nusselt number for the nanouid containing 1 vol.% alumina increases about 40% and 53% at Re = 332 and 1641, respectively, compared with that for the pure water. For the nanouid containing 2 vol.% alumina, further increase in the average Nusselt number beyond that for the nanouid of 1 vol.% alumina can only be detected for the Reynolds number

higher than 775. For the lower Reynolds numbers, the somewhat insensitiveness of the heat transfer enhancement to further increasing the particle fraction from 1 vol.% to 2 vol.%, as displayed in Fig. 5, might be caused by the agglomeration/clustering of the nanoparticles mentioned earlier in the Section 4, which may in turn give rise to particle deposition onto the channel wall, particularly at the lower ow rates through microchannels as that reported in the Ref. [15]. In effect, the actual particle fraction suspended in the nanouid owing through the heat sink can be expected to be lower than that of 2 vol.% considered. Effect of nanoparticle deposition on the heat transfer surface has been examined in the Refs. [19,20] having signicant bearing on wettability and/or nucleation site density, and thus the pool boiling behavior of nanouids; while the exact mechanisms remain largely unclear. In contrast, on the single-phase forced convective heat transfer of nanouids, the particle deposition effect has not been reported in the previous experimental studies [2124] except the Ref. [15]. One possible reason is that the ow channels/tubes considered in these works were mostly in mini-meter sizes, relatively much larger than that considered in [15] and the present study. Nevertheless, future effort should be put into gain basic understanding of the mechanisms for the nanoparticle deposition as well as its effects on convective heat transfer of nanouids. Furthermore, Fig. 6 illustrates heat transfer efcacy of replacing the base uid with the aluminawater nanouid formulated, in which the ratio of average heat transfer coefcient of the nanouid  , is plotted against the Reynolds  =h to that of the base uid, h nf bf number based on the properties of the base uid, Re(lnf/lbf)(qbf/ qnf). In conformity with the results of the average Nusselt number shown in Fig. 5, there exists marked dependence of the heat transfer coefcient ratio on the specic heat used. The results for the heat transfer coefcient ratio based on either specic heat display a somewhat non-monotonic variation over the range of Reynolds number. Especially, for the nanouid containing 1 vol.% alumina, local maximum and minimum enhancements of nearly 70% and 30% in the average heat transfer coefcient based on the higher specic heat can be discerned at Re(lnf/lbf)(qbf/qnf) = 1544 and 500, respectively, compared with that for the pure water. Next, the cooling performance of the heat sink using the nanouid is assessed by exploring the results for the thermal resistances Ritd and Rlm based on inlet and log-mean temperature difference (dened in Eqs. (8) and (9)), respectively, versus the

102

C.J. Ho et al. / Applied Thermal Engineering 30 (2010) 96103

0.029 K/W at the pumping power of 0.758 W. In contrast, the lowest value of Ritd for the water-cooled heat sink obtained is 0.038 K/ W at the pumping power of 0.660 W. On the other hand, a reduction of more than 28% in the log-mean temperature difference thermal resistance can be discerned for the nanouid containing 1 vol.% alumina at a pumping power of 0.660 W, compared with that for the pure water. Alternatively, the efcacy of using nanouid for the microchannel heat sink can be gauged in its effectiveness of minimizing the wall temperature, which is of particular interest in the aspect of temperature control/thermal management. Following that proposed in [16], a temperature control effectiveness eT w for the nanouid-cooled heat sink is dened based on the maximum wall temperature as

hnf / hbf

eT w 1

T w;max T in nf T w;max T in bf

17

Re( bf / nf )( nf / bf )
Fig. 6. Heat transfer coefcient enhancement of using aluminawater nanouid. (Symbols A and B, respectively, based on specic heat from Eqs. (3a) and (3b).)

pumping power as illustrated in Fig. 7. The pumping power was determined from the measured volume ow rate and pressure _ Dp. In contrast to that found for the heat transfer drop as P Q coefcient, Fig. 7 illustrates that the results of the thermal resistances are rather insensitive to the difference in the specic heat evaluated for the nanouid. As expected, both the inlet and logmean temperature difference thermal resistances display generally a decaying trend with increasing pumping power for the heat sink cooled by the pure water or nanouid. Most importantly, the two volume fractions of the nanouid appear to have signicantly lower value for either thermal resistance at a given pumping power than the pure water. The nanouid-cooled heat sink at higher pumping power, in particular, outperforms the water-cooled heat sink. For instance, with the increasing pumping power, the inlet temperature difference thermal resistance Ritd for the nanouid containing 1 vol.% of alumina levels off reaching a value of

Fig. 8 shows the temperature control effectiveness of the heat sink using the nanouid relative to the base uid as a function of the dimensionless parameter based on the properties of the base uid, (Lch/Dh)/[Pe(qbf/qnf)(cp,bf/cp,nf)(knf/kbf)]. Similar to that observed for results of the thermal resistance illustrated in Fig. 7, the temperature control effectiveness of the nanouid-cooled the heat sink shows little dependence on the effective specic heat formula adopted for nanouid. As can be clearly discerned in Fig. 8, positive effectiveness always results from replacing the nanouids for water in the heat sink. The maximum wall temperature of the nanouid-cooled heat sink can be increasingly suppressed with the ow rate (or the Peclet number) compared with that of the water-cooled one. More specically, a reduction of about 25% in the maximum wall temperature arises for the largest ow rate of 1026 cm3/min through the nanouid-cooled heat sink. The forgoing results for the thermal resistance and the temperature control effectiveness strongly reect that the nanouidcooled microchannel heat sink outperforms the water-cooled one. However, it can be noticed in Figs. 7 and 8 that further increase of the volume fraction of alumina to 2% induces virtually no effect on variations of thermal resistances and/or temperature control effectiveness of the nanouid-cooled heat sink. To further clarify its effect on the cooling effectiveness, the volume fraction

0.1 0.08

Ritd (K/W)

0.06 0.04 0.02 0

(Vol.%) 0 1 2

Symbols A B

0.4 0.35 0.3 0.25

Symbols (vol.%) A B 1 2

T
Rlm (K/W)
0.04 0.02 0 0 0.2 0.4 0.6 0.8

0.2

0.15 0.1 0.05 0

0.01

P(W)
Fig. 7. Relation of thermal resistance with pumping power. (Symbols A and B, respectively, based on specic heat from Eqs. (3a) and (3b).)

(Lch /Dh)/[Pe( bf /nf )(cp,bf /cp,nf)(knf /kbf)]

0.02

0.03

0.04

0.05

0.06

0.07

0.08

0.09

Fig. 8. Temperature control effectiveness of using aluminawater nanouid. (Symbols A and B, respectively, based on specic heat from Eqs. (3a) and (3b).)

C.J. Ho et al. / Applied Thermal Engineering 30 (2010) 96103

103

of nanoparticles in the nanouid shall be considered in a wider range with ner variation in the future experiments. 7. Concluding remarks In this article, the experimental results concerning hydraulic and thermal performances of a copper microchannel heat sink cooled by aluminawater nanouid of 1 and 2 vol.% have been presented. The present experiment conrms the ndings in the previous studies that the friction factor in the microchannel heat sink cooled by the nanouid containing small fraction of nanoparticles tends to increase minutely relative to the pure water. Through a correlation analysis, the relative decease in the specic heat of the nanouid with respect to its base uid might play a counteracting role to the benecial inuence due to its enhanced thermal conductivity on the average heat transfer coefcient in a microchannel. Accordingly, for the nanouid formulated, the disparity in the specic heat evaluated from the two formulas adopted was found to have a marked bearing on the experimental results for the average heat transfer coefcient. Moreover, compared with that of the base uid, the relatively lesser variation of dynamic viscosity with temperature of the nanouid of 2 vol.% might further serve as a deteriorating factor in heat transfer enhancement. By comparing with the heat transfer results obtained using the pure water, signicant enhancement in the average heat transfer coefcient and thereby marked reductions in the thermal resistance as well as in the maximum wall temperature were found for the heat sink cooled by the nanouid formulated. In all, the microchannel heat sink cooled by the aluminawater nanouid formulated was found to outperform that cooled by the pure water at high pumping power particularly. For the largest ow rate tested for the nanouid of 1 vol.%, the average heat transfer coefcient increases by about 70% compared with that with water; while the thermal resistance based on inlet temperature difference and the maximum wall temperature of the heat sink can be reduced, respectively, to as low as 0.029 K/W and by about 25%. The experimental results obtained are certainly far from complete particularly in the aspects of ranges and variations of relevant parameters such as the nanoparticle volume fraction, the geometric parameters of the heat sink, and so on. More efforts are denitely needed to resolve the disparity among the existing literature concerning the thermophysical properties of the nanouid, the cooling effectiveness of using nanouid as well as to address other technical problems, such as agglomeration/clustering/deposition of the nanoparticles, the passage clogging entering the microchannel as well as the long-term suspension stability of the nanoparticles in the base uid, which are all vital to practical implementation of the nanouid-cooled microchannel heat sink. Acknowledgements The authors greatly appreciate the support of the National Science Council of ROC in Taiwan for this study through Projects of NSC94-2212-E006-101, NSC95-2212-E006-233, and NSC96-2212-

E006-173. The constructive comments from the reviewers are sincerely appreciated. References
[1] D.B. Tuckerman, R.F.W. Pease, High-performance heat sinking for VLSI, IEEE Electron. Dev. Lett., EDL 2 (5) (1981) 126129. [2] C.B. Sobhanm, S.V. Garimella, A comparative analysis of studies on heat transfer and uid ow in microchannels, Microscale Thermophys. Eng. 15 (2001) 293311. [3] S.G. Kandlikar, W.J. Grande, Evolution of microchannel ow passages thermohydraulic performance and fabrication technology, Heat Transfer Eng. 24 (2003) 317. [4] S.G. Kandlikar, W.J. Grande, Evaluation of single phase ow in microchannels for high heat ux chip cooling thermohydraulic performance enhancement and fabrication technology, Heat Transfer Eng. 25 (2004) 516. [5] G. Hestroni, A. Mosyak, Z. Segal, G. Ziskind, A uniform temperature heat sink for cooling of electronic devices, Int. J. Heat Mass Transfer 45 (2002) 3275 3286. [6] P.S. Lee, S.V. Garimella, D. Liu, Investigation of heat transfer in rectangular microchannels, Int. J. Heat Mass Transfer 48 (2005) 16881704. [7] M.E. Steinke, S.G. Kandlikar, J.H. Magerlein, E.G. Colgan, A.D. Raisanen, Development of an experimental facility for investigating single-phase liquid ow in microchannels, Heat Transfer Eng. 24 (2006) 4152. [8] S.U.S. Choi, Enhancing thermal conductivity of uids with nanoparticles, Developments and Applications of Non-Newtonian Flows, FED-vol. 231/MDvol. 66, 1995, 99105. [9] W. Daungthongsuk, S. Wongwises, A critical review of convective heat transfer of nanouids, Renew. Sust. Energy Rev. 11 (2007) 797817. [10] X.Q. Wang, A.S. Mujumdar, Heat transfer characteristics of nanouids: a review, Int. J. Thermal Sci. 46 (2007) 119. [11] R. Chein, G. Huang, Analysis of microchannel heat sink performance using nanouids, Appl. Thermal Eng. 25 (2005) 31043114. [12] J. Koo, C. Kleinstreuer, Laminar nanouid ow in microheat-sinks, Int. J. Heat Mass Transfer 48 (2005) 26522661. [13] S.P. Jang, S.U.S. Choi, Cooling performance of a microchannel heat sink with nanouids, Appl. Thermal Eng. 26 (2006) 24572463. [14] J. Lee, I. Mudawar, Assessment of the effectiveness of nanouids for singlephase and two-phase heat transfer in micro-channels, Int. J. Heat Mass Transfer 50 (2007) 452463. [15] R. Chein, J. Chuang, Experimental microchannel heat sink performance studies using nanouids, Int. J. Thermal Sci. 46 (2007) 5766. [16] T.L. Bergman, Effect of reduced specic heats of nanouids on single phase, laminar internal forced convection, Int. J. Heat Mass Transfer 52 (2009) 1240 1244. [17] W.M. Kays, A.L. London, Compact Heat Exchangers, McGraw-Hill, New York, 1984. [18] F. Incropera, D. Dewitt, Fundamentals of Heat and Mass Transfer, fth ed., Wiley, New York, 2002. [19] S.J. Kim, I.C. Bang, J. Buongiorno, L.W. Hu, Effects of nanoparticle deposition on surface wettability inuencing boiling heat transfer in nanouids, Appl. Phys. Lett. 89 (2006) 153107. [20] S.J. Kim, I.C. Bang, J. Buongiorno, L.W. Hu, Surface wettability change during pool boiling of nanouids and its effect on critical heat ux, Int. J. Heat Mass Transfer 50 (2007) 41054116. [21] B.C. Pak, Y.I. Cho, Hydrodynamic and heat transfer study of dispersed uids with submicron metallic oxide particles, Exp. Heat Transfer 11 (1998) 150 170. [22] D.S. Wen, Y.S. Ding, Experiment investigation into convective heat transfer of nanouids at the entrance region under laminar ow conditions, Int. J. Heat Mass Transfer 47 (2004) 51815188. [23] S. Zeinali Heris, M. Nasr Esfahany, S.Gh. Etemad, Experimental investigation of convective heat transfer of Al2O3/water nanouid in circular tube, Int. J. Heat Fluid Flow 28 (2007) 203210. [24] U. Rea, T. McKrell, L.W. Hu, J. Buongiorno, Experimental study of laminar convective heat transfer and viscous pressure loss of aluminawater nanouid, in: Proceedings of ASME MNHT Conference, MNHT-2008-52263, 2008.

Вам также может понравиться