You are on page 1of 22

This article was downloaded by: [Universiti Teknologi Malaysia] On: 01 October 2011, At: 10:27 Publisher: Taylor

& Francis Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK

Microscale Thermophysical Engineering


Publication details, including instructions for authors and subscription information: http://www.tandfonline.com/loi/umte19

HEAT TRANSFER IN MICROCHANNELS


Bjrn Palm Available online: 29 Oct 2010

To cite this article: Bjrn Palm (2001): HEAT TRANSFER IN MICROCHANNELS, Microscale Thermophysical Engineering, 5:3, 155-175 To link to this article: http://dx.doi.org/10.1080/108939501753222850

PLEASE SCROLL DOWN FOR ARTICLE Full terms and conditions of use: http://www.tandfonline.com/page/ terms-and-conditions This article may be used for research, teaching, and private study purposes. Any substantial or systematic reproduction, redistribution, reselling, loan, sub-licensing, systematic supply, or distribution in any form to anyone is expressly forbidden. The publisher does not give any warranty express or implied or make any representation that the contents will be complete or accurate or up to date. The accuracy of any instructions, formulae, and drug doses should be independently verified with primary sources. The publisher shall not be liable for any loss, actions, claims, proceedings, demand, or costs or damages whatsoever or howsoever caused arising directly or indirectly in connection with or arising out of the use of this material.

Microscale Thermophysical Engineering, 5:155175, 2001 Copyright 2001 Taylor & Francis 1089-3954/01 $12.00 + .00

REVIEW

HEAT TRANSFER IN MICROCHANNELS Bjrn Palm


Downloaded by [Universiti Teknologi Malaysia] at 10:27 01 October 2011
Department of Energy Technology, Royal Institute of Technology, Stockholm, Sweden
In this article an attempt has been made to review the literature regarding heat transfer and pressure drop in one- and two-phase ow in microchannels. The emphasis has been on reports presented during the last few years. For single-phase ow, channels with hydraulic diameters less than 1 mm have been considered. For two-phase ow, very little information is available for such small channels. Also, for two-phase ow, deviations from largetube behavior start at diameters of a few millimeters. For these reasons a slightly larger diameter range has been considered in this case. As a conclusion, it can be stated that the understanding of ow in microchannels is increasing steadily, but that there are still many questions to be answered concerning the reasons for deviations from classical theory developed for larger channels.

During the last decade there has been growing interest in heat transfer in microchannels. Several reasons for this may be mentioned: with developments within the electronic industry, methods have been contrived for manufacturing complex geometries on a very small scale. These manufacturing techniques include etching, vapor depositioning, and bonding in silicon and other materials, but also precision machining and methods for the forming of polymers. In general, the materials technologies have taken a leap forward, making possible new manufacturing methods for microdesign, often at a low cost when used in large-scale production. However, the new manufacturing methods for microdesign would not have come to use in the area of heat exchange if there was not a market for miniature heat exchangers. Within the electronics industry, the need for micro heat exchangers is a result of the miniaturization of the electronics, which leads to denser packaging of components and thereby to higher heat uxes. As the cooling limit of air cooling is reached, the rst alternative is liquid cooling, with or without evaporation of the uid. So far, there has not been a broad introduction to the market of liquid-cooled electronics, with the exception of heat pipes used in laptop computers. However, intense activities are going on in this area, and it could be expected that within ve years the value of these types of cooling systems will increase 10-fold.

Received 13 September 2000; accepted 29 March 2001. Address correspondence to Prof. Bjrn Palm, Department of Energy Technology, Royal Institute of Technology, SE-100 44, Stockholm, Sweden. E-mail: bpalm@egi.kth.se 155

156

B. PALM

NOMENCLATURE
a Bo Br c d dh E F f G G eq thermal diffusivity, m2 /s boiling number [5 q (G l h fg ) ] Brinkman number [5 l u 2 (k D t) ] m speed of sound, m/s diameter, m hydraulic diameter, m enhancement factor suppression factor friction factor mass ux, kg/m2 s equivalent all-liquid mass ux 5 {G [(1 x m ) 1 x m (q l q v ) 0.5 ]}, kg/m2 s heat transfer coef cient, W/m2 K latent heat of vaporization, J/kg channel height, m thermal conductivity, W/m K Knudsen number (5 dh ) length, m molar mass number of parallel channels Nusselt number (5 h dh k) pressure, Pa reduced pressure Prandtl number (5 l cp k) heat ow, W heat ux, W/m2 radius of curvature, m Reynolds number (5 u dh t l 5 G dh l l ) equivalent all liquid Reynolds number (5 G eq dh l l ) l t q s Rp S t u V v9 v9 9 W Wc x D p D t surface roughness parameter, l m gap width, m temperature, C velocity, m/s volume ow, m3 /s speci c volume of liquid at saturation, m3 /kg speci c volume of vapor at saturation, m3 /kg channel width, m c/c distance between channels, m vapor fraction critical dimension for the microchannel, m pressure difference, Pa temperature difference, K mean free path, m dynamic viscosity, N s/m2 kinematic viscosity, m2 /s density, kg/m3 surface tension, N/m time, s

Downloaded by [Universiti Teknologi Malaysia] at 10:27 01 October 2011

h h fg H k Kn L M n Nu p pr Pr Q q r Re Reeq

Indices m mean nb nucleate boiling l liquid s saturation sup superheat v vapor

Microchannel heat exchangers have also found other applications. In the automotive industry, the need for low weight and small overall volume has pushed development toward smaller-diameter heat exchangers. Also, new manufacturing technologies have made possible more compact designs. One good example of this is multichannel tubes of aluminum, which are now manufactured by several companies. These tubes may have individual channels with diameters well below 1 mm. When discussing microchannel heat transfer, a de nition of the term is necessary. One de nition suggested is that microchannels are channels in which classical theory is no longer valid. However, since it is still not clear at what diameter this will occur, this de nition is dif cult to apply. An alternative, simple de nition is that a microchannel is any channel with a (hydraulic) diameter in the micrometer range, i.e., less than 1 mm. This de nition seems to be in accordance with the use of the word in the literature concerning single-phase ow. For two-phase ow, however, there is very little reported in the literature on ow in channels less than 1 mm. More important, deviations from classical theory may start at diameters of several millimeters. For these reasons it has

HEAT TRANSFER IN MICROCHANNELS

157

been considered appropriate for the purpose of this review to have a wider de nition of the term microchannel in the case of two-phase ow than that applied to single-phase ow. Almost all microchannels discussed in the literature have diameters larger than 1 l m. It should be noted that the diameter range of microchannels thus is quite large, and different phenomena may occur in different parts of this range. Looking through the literature, it is quite clear that the interest in the area of microchannel ow and heat transfer has increased substantially during the last decade. In general, there also seems to be a shift in the focus of published articles, from descriptions of the manufacturing technology to discussions of the physical mechanisms of ow and heat transfer.

Downloaded by [Universiti Teknologi Malaysia] at 10:27 01 October 2011

A MOTIVATING EXERCISE For single-phase ow, it is easy to show the advantages of going toward smaller channel diameter using the classical correlations derived for macroscopic channels. In fully developed laminar ow, the Nusselt number is constant, which means that the heat transfer coef cient is inversely proportional to the diameter: Nu 5 hd 5 k const ha 1 d (1)

To transfer a certain amount of heat at a given logarithmic mean temperature difference (LMTD), the product of the surface area and the heat transfer coef cient must be constant. Assuming circular tubes, and allowing any number n of parallel channels, this leads to the conclusion that the product of the tube length L and the number of channels should be constant: Q5 hA D t 5 h ( d L n) D t L n5 const (2)

The friction factor in fully developed laminar ow is inversely proportional to the Reynolds number, which means that the pressure drop is inversely proportional to the diameter to the power of 4: D p 5 C L q u2 5 Re d 1 d4 Ct q u ud V d2 4 L d (3)

D p a

This also shows that, to keep the pressure drop constant, at a given mass ow, while changing the tube diameter and the number of parallel tubes, the ratio L (n d 4 ) must be kept constant D p 5 C 9 t q ( V tot n) L 4 d L 5 n d4 const (4)

However, as L n also should be constant to maintain the mean temperature difference, the product (n d 2 ) should be constant.

158

B. PALM

Downloaded by [Universiti Teknologi Malaysia] at 10:27 01 October 2011

In conclusion, to maintain the pressure drop and the mean temperature difference at a given mass ow and heat ow, while changing the tube diameter and the number of channels, the product n d 2 should be kept constant. The tube length is then given by the condition (L n 5 const) above. Applying these simple relations gives the results shown in Figure 1. As an example, decreasing the diameter by half would double the heat transfer coef cient and reduce the necessary surface area to one-half of the original. The tube length would be reduced to one-fourth of the original, and the number of parallel channels would increase by a factor of 4. In addition, the internal volume (n d 2 4 L ), would decrease by a factor of 4. As an additional bene t, the thickness of the tube wall necessary to withstand a certain pressure would decrease. In conclusion, a decrease of the tube diameter in the laminar- ow region leads to much more compact designs. A similar analysis could be made also for the turbulent- ow region. The results are similar, although the advantage of reducing the tube diameter is not as large in this case.

Figure 1. Dependence of tube length, number of parallel tubes, heat transfer area, and heat transfer coef cient on tube diameter for equal mass ow, heat ow, and LMTD. Laminar ow.

HEAT TRANSFER IN MICROCHANNELS

159

Downloaded by [Universiti Teknologi Malaysia] at 10:27 01 October 2011

Figure 2. Example of dependence of tube length, heat transfer area, heat transfer coef cient, internal volume, ow area, and number of parallel tubes on tube diameter for equal mass ow, heat ow, and LMTD. Laminar and turbulet ow.

To get the complete picture, the turbulent and the laminar regions may be combined, looking at a speci c case. The result of such a comparison is shown in Figure 2. The reference case is a single 10-m tube of 16-mm inner diameter, with water owing at a rate of 6 kg/min. The heat exchange rate is 20 kW. As the tube diameter is decreased to below 3 mm the ow will turn laminar, leading to a sudden change in the parameters. The range of diameters between 3 and 1 mm would obviously not be the best choice, as the total surface area is even larger than that of the original single tube. However, the diameter does not have to be decreased much further to give a considerable advantage in compactness. As an example, this analysis shows that the original tube (16 mm diameter, 10 m length) could be exchanged for a bundle of 0.5-mm tubes, 10 cm in length, having a total internal cross-sectional area less than twice that of the single tube. In ow boiling and condensation a similar analysis would be more complex, as the heat transfer coef cients depend on the ow regime. Also, there are no generally accepted correlations for calculation of heat transfer and pressure drop even if the ow regimes are known. For larger-diameter tubes (> 5 mm) it is generally recognized that the ow boiling heat transfer coef cients increase with decreasing diameter. However, for smallerdiameter tubes new phenomena become important which prohibit the extrapolation of this trend. ONE-PHASE FLOW Introduction For the practical application of microchannel heat exchangers in one-phase ow, the two parameters of prime interest are the friction factor and the heat transfer coef cient. Most of the work described in the literature is concerned with these two parameters, their possible deviations from classical theory, and the reasons for these deviations. Un-

160

B. PALM

Downloaded by [Universiti Teknologi Malaysia] at 10:27 01 October 2011

fortunately, the literature is not conclusive concerning the dependence of any of these parameters on the channel diameter. Reviews of heat transfer in microchannels have been presented by a number of authors [15]. Duncan and Peterson [2] provided a wide review of both one-phase and two-phase microscale convective heat transfer as well as micro-heat pipes, microscale conduction, and radiation. Peng and Wang [3, 4] gave a short review of one- and twophase ow, but concentrated on presenting their own extensive research. Bailey et al. [5] gave a thorough treatment of one-phase forced convection and concluded that the literature is inconclusive as to the effect of miniaturization on heat transfer and pressure drop. They end, however, by stating that there seem to be indications that friction factors in laminar ow are lower, and turbulent ow heat transfer coef cients are higher than expected for larger-diameter channels. Previous Work in One-Phase Flow One of the rst reports concerning heat transfer and pressure drop in microchannels was presented by Tuckerman and Pease [6]. They designed and tested a multichannel heat exchanger for cooling of electronic components. The channel width was about 50 l m and the depth about 300 l m. Deionized water was used as test uid. The ow was found to obey classical laminar- ow theory, and the total thermal resistance was independent of the mass ux, as would be expected for fully developed laminar ow. Another early investigation of microchannel pressure drop was done by Wu and Little [7], who designed a microminiature Joule-Thompson refrigerator. The hydraulic diameters of their rectangular test species ranged from 50 to 80 l m. The measured friction factors were found to be much higher than expected from classical theory (Moody diagram), even though the Reynolds number dependence had about the same shape as expected. The same authors also investigated the heat transfer [8]. In this case, they used nitrogen gas passing through rectangular channels with hydraulic diameters close to 150 l m. The authors de ned the ranges of the three ow regimes as laminar (Re < 1,000), transition (1,000 < Re < 3,000), and turbulent (Re > 3,000). This indicates transition from laminar ow at lower Reynolds number than expected for larger tubes. For most Reynolds numbers, the heat transfer coef cients were found to be higher than expected from classical theory. For the turbulent regime, they proposed the following correlation: Nu 5 0.00222 Re1.09 Pr 0.4 (5)

In the laminar regime, the Nusselt number was found to be a function of the Reynolds number with an exponent slightly higher than that in the turbulent regime. At the lowest Re the results were lower than expected. The relative roughness of the channel walls was fairly high and unsymmetric, which might have in uenced the results and could explain, at least partly, why both pressure drop and heat transfer were higher than expected. Peng and Peterson published several reports on single-phase as well as two-phase ow in microchannels. In one article [9] they investigated heat transfer and pressure drop of water owing through arrays of rectangular microchannels having different aspect ratios and hydraulic diameters in the range 0.150.34 mm. The range of Reynolds num-

HEAT TRANSFER IN MICROCHANNELS

161

bers was from 50 to 4,000. For both the laminar and the turbulent regimes, the Nusselt number was found to be a function not only of the Reynolds and Prandtl numbers but also of different geometric parameters. For laminar ow they suggested the following correlation: Nu 5 0.1165 dh Wc
0.81

H W

0.79

Re0.62 Pr 1

(6)

where W c is the center-to-center distance between channels. For turbulent ow the suggested correlation is similar to the Dittus-Boelter correlation but with the constant being dependent on the geometry. The friction factors corresponding to the measured pressure drops deviated considerably from the classical values and were more strongly dependent on the Reynolds number than expected. Both higher and lower friction factors were measured. They also noted that transition from laminar to transition and turbulent ow started at much lower Reynolds number than for larger-diameter tubes. Similar results were reported earlier by Peng and Wang [10] and Wang and Peng [11]. Cuta et al. [12] tested a heat exchanger consisting of 54 parallel channels with rectangular cross section (1.0 mm 0.27 mm), corresponding to a hydraulic diameter of 0.425 mm. The length of the channels was 20.52 mm, and R-124 was used as test uid. The range of Reynolds number was from 100 to 570. The authors report that the friction factors were considerably smaller than expected from classical theory. The Nusselt number, on the other hand, was found to be substantially larger than expected from laminar- ow theory. In addition, the Nusselt number increased with increasing Reynolds number, but at a lower rate than would be expected for turbulent ow (Nu a Re0.6 ). Harms et al. [13] investigated the ow of deionized water through 68 parallel rectangular microchannels, each being 0.251 mm wide and 1.000 mm deep, having a length of approximately 25 mm. The Reynolds number range covered was from 173 to 12,900. The channels were etched from a silicon wafer and covered by a glass plate. The friction factor was found to be reasonably well predicted by classical theory, both in the laminar and in the turbulent regimes. However, the critical Reynolds number was found to be lower, about 1,500, but the authors attributed this to the severity of the inlet condition. The experimentally determined Nusselt numbers were compared to the theoretical values for developing laminar ow at low Reynolds number and for fully developed turbulent ow at the higher Reynolds number. The agreement was reasonably good at the higher Reynolds numbers, but at low Re the results were signi cantly lower than expected. This deviation was thought to be due to ow bypass in the manifold, and the authors conclude that the classical relations for the local Nusselt number should also be valid for microchannels. Mala and Li [14] studied the ow of water through microtubes with diameters ranging from 50 to 254 l m. Tubes were manufactured from two different materials, stainless steel (SS) and fused silica (FS). Friction factors were generally higher than expected from classical theory. For Reynolds numbers below 1,000 and for the larger diameters tested, the friction factors were approximately in agreement with classical

Downloaded by [Universiti Teknologi Malaysia] at 10:27 01 October 2011

162

B. PALM

theory. The deviations increased with decreasing diameter and with increasing Reynolds number. The authors conclude that in microtubes, for Re 1,500, the Blasius equation is f Blasius 5 0.3164 Re
0.25

(7)

Downloaded by [Universiti Teknologi Malaysia] at 10:27 01 October 2011

This is an indication of an earlier transition from laminar to turbulent ow in microtubes as compared to larger-diameter tubes. A diameter dependence of this transition is also discernible from the experimental data. However, the authors suggest that the early transition may also be a result of surface roughness effects. Flockhart and Dhariwal [15], tested the ow of water in trapezoidal channels having hydraulic diameters in the range 50120 l m. The Reynolds numbers were kept below 600 to give laminar ow. The experimentally determined friction factors were in good agreement with results of a numerical calculation based on classical theory. Qu et al. [16] recently presented an investigation of water ow through trapezoidal silicon microchannels with hydraulic diameters ranging from 51 to 169 l m. The investigated Re numbers were below 1,500. They found the friction factors to be 10% to 40% higher than expected from classical theory. Possible Explanations for Deviations from Classical Theory As seen above, the data reported in the literature are inconclusive. Some investigators have reported increased heat transfer and/or pressure drop, while others reported the opposite. To some degree these deviations may be caused by the dif culties in measuring the parameters necessary for the theoretical calculation. As noted above, the pressure drop in laminar ow is inversely proportional to the diameter to the power of 4, meaning that a very precise value of the diameter is necessary for determining the friction factor. Also, it could be dif cult to measure the tube wall temperatures, which are necessary for determining the logarithmic mean temperature difference (LMTD) correctly. It should be noted that the Nusselt number in laminar ow should only be expected to be constant for Long tubes, at isothermal conditions, and if the heat transfer coef cient is referred to the LMTD. A second cause for deviations could be entrance effects, i.e., that the experimental results are compared to solutions for fully developed ow, which is not always a good approximation. Pressure drop at tube entrances and exits may also have been neglected in some cases. A third possible source of deviations, mentioned by some authors, is the relative surface roughness of the tube surfaces. Even though some of the deviations may be explained by the rather trivial causes mentioned above, there are also new phenomena which may be of importance due to the small scale in microtubes. Papautsky et al. [17] used a numerical model including micropolar uid theory to calculate the normalized friction coef cient for ow-in microchannels. The model took into account microrotational effects of the molecules and the variation in the apparent viscosity of the uid close to the wall. It was then compared to experimental results using water as the test uid in the Reynolds number range 120. The tested microchannels had rectangular cross sections with widths ranging from 50 to 600 l m and heights from 20 to 30 l m. The friction coef cients were found to be around 12% higher than expected from

HEAT TRANSFER IN MICROCHANNELS

163

classical macroscopic theory. However, when the results were compared to the numerical model the deviations were much smaller. It was also shown that experimental data from other authors [18, 19] were well correlated by the numerical model. Tso and Mahulikar have presented several articles advocating the inclusion of the Brinkman number in the heat transfer correlation for laminar ow in microchannels [2023]. l u2 m Br 5 (8) kD t This number re ects the relative importance of viscous heating to uid conduction. As viscous heating is not important in normal ow through conventional size channels, the Brinkman number is not included in the classical heat transfer correlations. However, the authors point to the fact that the velocity gradients in laminar ow in microchannels are extremely high, and that the length-to-diameter ratios may be large in these cases. Through a conventional dimensional analysis [20] they arrive at a correlation of the form Nu 5 A Rea Pr b
c

Downloaded by [Universiti Teknologi Malaysia] at 10:27 01 October 2011

dh

Brd

(9)

where is a critical dimension for the microchannel. The exponent d is positive for heating of the uid and negative for cooling. In a later article [23] they report on new experimental data supporting their earlier proposition. In both studies, they use the exponents of Re and Pr given by Peng and Peterson [9]. Another phenomenon, which has been proposed as responsible for deviations from macroscopic behavior in liquid ow, is the electric double-layer (EDL) effect. Yang et al. [24] investigated the effects of the electric double layer and the in uence of this effect on pressure drop and heat transfer. Based on their model they estimated the friction factor and the Nusselt number for an aqueous solution of low ionic concentration and a wall surface of high zeta potential. For rectangular channels with the shortest side around 2040 l m it was found that the electrokinetic effects could have signi cant in uence on the friction factor and the Nusselt number. However, it was also noted that for the conditions used in the evaluation of the model, EDL effects should not be important for pressure drop or heat transfer in channels larger than 40 l m. When considering gas ow in microchannels it must be recognized that the continuum assumption is invalid when the mean free path of the molecules is of the same order of magnitude as the hydraulic diameter of the channel. This relation could be expressed by the Knudsen number, Kn 5 dh (10)

At high Knudsen numbers, the assumption of zero velocity at the wall will not be valid, as there is little interaction between the molecules close to the surface. Slip ow will then lead to a lower friction factor than expected for larger channels. Beskok and Karniadakis [25] suggested a range of the Knudsen number between 10 3 and 0.1 in which slip ow could be expected, while at lower Knudsen number the uid could be treated as a continuum and the zero-wall velocity assumption thus should apply. Other effects may also be present in gas ow in microchannels. A theoretical analysis of the thermal conductivity of the gas in the wall-adjacent layer was presented

164

B. PALM

recently by Li et al. [26]. The analysis is based on a simpli ed kinetic gas theory assuming rigid-sphere molecules. The result of this analysis is that the thermal conductivity in a layer close to the wall is considerably lower than in the bulk region. Within this region the ratio k k bulk varies from 0.5 at the surface to 1 at a distance of three to ve times the mean free path of the molecules. A similar analysis was presented by the same authors in an earlier article concerning the viscosity close to the wall [27]. From theses analyses the Nusselt number in fully developed laminar ow was deduced. This showed that for circular tubes a slight reduction in the Nusselt number should be expected at Kn 0.005. Also, the friction factor should be smaller than expected according to classical theory.

Downloaded by [Universiti Teknologi Malaysia] at 10:27 01 October 2011

Conclusions, One-Phase Flow In conclusion, there seem to be no general agreement as to what diameters classical theory can be applied for determining friction factors and heat transfer coef cients in microchannels. Several authors have concluded that transition from laminar to transition and turbulent ow starts at lower Reynolds numbers than expected for larger hydraulic diameters and that the critical Reynolds number decreases with decreasing hydraulic diameter. The reported friction factors are both above and below predictions of classical laminar theory. The friction coef cient has also been reported to be dependent on the Reynolds number. It seems, though, that the deviations are in most cases smaller than 6 30%, although deviations of more than 100% have been reported. For the heat transfer there are several reports indicating slightly higher Nusselt numbers in turbulent ow than expected. In laminar ow both higher and lower Nusselt numbers have been reported. Suggested reasons for these deviations are surface roughness effects, entrance effects, electric double-layer effects, nonconstant uid properties, two- and threedimensional transport effects, and slip ow (for gases). Finally, the deviations in the experimental results may be caused by the dif culties in accurately determining hydraulic diameters and uid and surface temperatures in microchannels. Considering these dif culties, the scatter in the results is not surprising. TWO-PHASE FLOW Introduction Heat transfer in two-phase ow in microchannels has not been studied as extensively as single-phase ow. Especially, the size range below 1 mm has been investigated by only a few researchers. Most of what has been presented concerns evaporation, and only four reports are known to the author about ow condensation inside tubes of diameters less than 3 mm. As noted above, even at these relatively large diameters deviations from large diameter behavior should be expected. The main reasons for this is that surface tension forces are more dominant and gravity forces less dominant in small-diameter tubes. Evaporation Flow boiling heat transfer is often assumed to be the result of two different mechanisms, nucleate boiling and convective boiling. In general, nucleate boiling is dominant

HEAT TRANSFER IN MICROCHANNELS

165

at high heat ux and low vapor quality, while convective boiling is important at high mass ux and high vapor quality, where nucleate boiling is suppressed. The local heat transfer coef cient is then calculated as a sum of the two contributions: hn 5 tp (E h l ) n 1 (F h nb ) n (11)

where h l and h nb are the heat transfer coef cients for one-phase liquid ow and for pool boiling respectively. E and F are enhancement and suppression factors. h l in the above equation is calculated by, e.g., the Dittus-Boelter equation, while h nb is calculated by a pool boiling correlation, e.g., the Cooper [28] correlation.

Downloaded by [Universiti Teknologi Malaysia] at 10:27 01 October 2011

h nb 5

55 p r

0.12 0.2log 10 R p

log10 p r )

0.55

0.5

q 0.67

(12)

In the nucleate boiling regime the heat transfer coef cient is thus dependent on the heat ux and the saturation pressure. In the convective boiling regime heat transfer is mainly in uenced by the mass ux and the vapor quality. The in uence of channel diameter on the heat transfer coef cient is not obvious from the above correlation. The Pierre correlation for complete evaporation in horizontal tubes [29], which has proved reasonably accurate for conditions common in refrigeration applications, indicate that, at constant mass ux, the heat transfer coef cient is inversely proportional to the diameter to the power of 0.2, i.e., a slight increase with decreasing diameter. However, this correlation has not been tested against microchannel data. When choosing working uid for two-phase cooling systems it should be recognized from the Cooper correlation that uids with low molar mass and high (reduced) pressure are expected to give the highest heat transfer coef cients. Such uids could also be expected to have high critical heat ux [30]. The uids in uence on the global environment must also be considered. One of the rst reports on boiling in narrow tubes was presented by Lazarek and Black [31]. In their tests, R-113 was evaporated in vertical tubes with a diameter of 3.1 mm. Pressure drop, critical heat ux, and heat transfer coef cients were measured. It was found that heat transfer was relatively independent of vapor fraction but highly dependent on heat ux, indicating a dominant in uence of nucleate boiling. They suggested the following correlation for the heat transfer coef cient: Nu l 5 30 Re0.857 Bo0.714 l (13)

It should be noted that as Re is proportional, and Bo inversely proportional to the mass ux, this means that the heat transfer coef cient is only a weak function of this parameter and mainly dependent on the heat ux. Another early report on boiling in narrow channels in between one heated plate and an adiabatic glass plate was given by Fujita et al. [32]. The gap sizes tested were 5, 2, 0.6, and 0.15 mm. The channel widths were 30 mm and the heights 30 and 120 mm. The test uid was water at atmospheric pressure. Tests were run with the sides of the channel either closed or open. With the sides open it was found that the data for the three widest channels were well predicted by the equation Nu 5 16 (Re1 2 ) 2
3

(14)

166

B. PALM

where Nu is calculated using the gap width as characteristic length and the liquid thermal conductivity, and Re is de ned by Re 5 q D hf g q
t

S t

q q0

(15)

Downloaded by [Universiti Teknologi Malaysia] at 10:27 01 October 2011

and q0 5 1.6 106 W/m2 . With closed sides this correlation often underpredicted the experimental results. For the 0.15-mm channel, most of the surface area was found to be dry and only the edges were wetted by liquid. Aligoodarz and Kenning [33] investigated the behavior of water vapor bubbles during evaporation in single channels having a cross section of 2 1 mm. Three sides of the channels were heated, while the fourth side was a glass window through which the bubbles could be monitored. The surface temperature at the bottom wall was measured from the color (hue) of a thin layer of thermochromic liquid crystals. Three types of ow regimes were noted: saturated nucleating bubbles, sliding bubbles, and slightly subcooled bubbles. They stated that part of the bubble growth had similarities to the process in pool boiling when a thin liquid lm evaporates underneath a growing bubble. Cornwell and Kew [34] used high-speed video to investigate forced convective boiling of R-141b inside single tubes of diameters from 1.39 to 3.69 mm. Relating to previous publications, they stated that there were very few reports in the literature of ow boiling in channels where the con nement number, de ned as Co 5 { [g (q
l

1 2 g ) ]}

dh

(16)

is in the range 0.510. For this range, they show experimentally that four different ow regimes may occur: isolated bubble (IB), con ned bubble (CB), annular slug ow (ASF), and partial dryout (PD). According to the authors, the heat transfer coef cient calculated by a traditional pool boiling correlation such as Coopers [28] gives the lower limit for all these regions except for partial dryout. They suggest that the heat transfer coef cients in the CB and ASF regimes may be calculated by a fairly simple model, where all thermal resistance is concentrated to the liquid lm in between the bubble and the channel wall. They also suggest methods of predicting this lm thickness. Finally, they show that their model is in reasonable agreement with experimental data. A clear difference is seen in the test results between the larger- and the smallerdiameter tubes. In the larger tubes, the heat transfer coef cient increases with increasing vapor fraction, whereas with the smaller-diameter tubes, the heat transfer coef cient decreases with increasing vapor fraction. This last behavior is thought to be caused by partial dryout in the smaller tubes. Tran et al. [35] published a very interesting report on boiling of R-12 and R-113 in circular and rectangular channels with hydraulic diameters of 2.4, 2.46, and 2.92 mm. Even though this is at the upper limit of what could be called microtubes, it was found that the heat transfer mechanisms deviate substantially from those in larger-diameter tubes. The authors found that at all but the lowest heat uxes, the heat transfer coef cient was independent of vapor quality and of mass ux. On the other hand, it was heavily dependent on heat ux and also on the pressure level. This showed clearly that nucleate

HEAT TRANSFER IN MICROCHANNELS

167

boiling was the dominant heat transfer mechanism. At the lowest heat uxes (< 68 kW/m2 ), corresponding to temperature differences less than 2.75 C, there was a clear dependence on the mass ux but not on heat ux. The authors concluded that below this temperature difference forced convection was dominant. The transition between the two regions was more obvious than expected for larger-diameter tubes and, more important, appeared at lower temperature differences than for larger tubes. Comparing the circular tube (diameter 2.46 mm) and the rectangular channel (1.7 mm 4.06 mm, dh 5 2.4 mm), it was found that the differences in heat transfer coef cients were very small. A correlation for the average heat transfer coef cient (in W/(m 2 K) in the nucleate boiling regime (D t > 2.75 C) was also presented:

Downloaded by [Universiti Teknologi Malaysia] at 10:27 01 October 2011

h5

8.4

10

(Bo2 Wel ) 0.3 q

l v

0.4

(17)

Compared to correlations for larger-diameter tubes, the heat transfer coef cients were considerably higher. Xia et al. [36] measured the heat transfer as R-113 boiled at atmospheric pressure between two plates forming a vertical narrow channel. The heat ux, the gap width, and the channel height were varied and the in uence on heat transfer and critical heat ux were studied. It was found that there exists an optimum gap size resulting in the highest heat transfer coef cients. However, this optimum also depends on the height of the channel. It was also found that the critical heat ux is reduced with decreasing channel size, and that the smaller the gap size, the lower the incipient heat ux. Bonjour and Lallemand [37] reported on tests of boiling in between one heated and one adiabatic surface. The channel height was 120 mm and the gap size ranged from 0.5 to 2 mm. Hot-wire anemometry was used to detect the phases. Three ow regimes were detected: nucleate boiling with isolated deformed bubbles, nucleate boiling with coalesced bubbles, and partial dryout. A new ow regime map was also suggested based on the Bond number and the ratio between heat ux and critical heat ux. Lin and Kew [38] used a 1-mm vertical tube for tests with R-141b as refrigerant. They found that, for the lower heat uxes (1840 kW/m2 ), the heat transfer coef cients increased with increasing vapor fraction, while for the higher heat uxes (> 60 kW/m2 ), the trend was opposite. At vapor fractions between 40% and 80% the heat transfer coef cients were more or less independent of vapor fraction at all but the highest heat ux (72 kW/m2 ). In general, the levels of h were in the range 2,0006,000 W/(m 2 K). Yan and Lin [39] measured heat transfer and pressure drop during evaporation of R-134a in circular tubes of inside diameter 2 mm. They reported that heat transfer was generally higher (3080%) than expected for an 8-mm-diameter tube (as calculated by correlations from the literature). At low vapor qualities, the heat transfer coef cient increased with increasing heat ux, while at higher vapor qualities it seemed to be more dependent of mass ux. The friction factor was found to be fairly constant, varying in the range 0.030.08, decreasing only slightly with increasing Re. Hapke et al. [40] used infrared thermography to determine the surface temperatures of a vertical heated tube with 1.5-mm inside diameter in which water was boiling at atmospheric pressure. From these measurements, they were able to calculate the local surface temperatures on the inside of the tube. These temperatures decreased at the

168

B. PALM

Downloaded by [Universiti Teknologi Malaysia] at 10:27 01 October 2011

location of bubble nucleation. They also suggested a correlation for determining the location of and the temperature difference at the onset of boiling. Bao et al. [41] tested R-11 and R-123 boiling in a smooth copper tube with an internal diameter of 1.95 mm. Heat uxes were varied from 5 to 200 kW/m2 and mass uxes from 50 to 1,800 kg/(m2 s). The in uence of mass ux, vapor quality, heat ux, and system pressure on the heat transfer coef cient was investigated for the two uids. It was found that the in uence of mass ux and vapor quality was very small, while that of heat ux and system pressure was large. These results suggest that the dominant heat transfer mechanism was nucleate boiling. The experimental results were compared to several ow boiling correlations, but none of them predicted the experimental data for all conditions. The authors conclude by stating that Coopers [28] correlation for pool boiling is in quite good agreement with the experimental data. Flow patterns, pressure drop, and void fractions in adiabatic two-phase ow of water and air have been studied by Triplett et al. [42, 43]. The tubular test sections had diameters of 1.09 and 1.49 mm. New ow pattern maps for microtubes are suggested, and recommendations for the calculation of void fraction and pressure drop are given. However, it should be noted that the results from these adiabatic tests may not be applicable to the highly dynamic and unstable situation in ow boiling. Zhao et al. [44], investigated the ow boiling of CO2 and R-134a in microchannels (diameter not stated). The heat transfer coef cient was found to be independent of both the mass ux and the heat ux. This was interpreted as indicating a dominant in uence of nucleate boiling, but, simultaneously, a suppression of nucleation at increased vapor fraction. Compared to R-134a, CO2 was found to give twice as high heat transfer coef cients at less than half the pressure drop. These differences were attributed to the lower surface tension and liquid viscosity of CO2 . It should be noted that the use of CO2 as a boiling coolant is restricted to low temperatures, as the critical temperature of this uid is 1 31 C. Phenomena in Small-Diameter Channels Peng, with different co-workers, has made extensive studies of one- and two-phase ow phenomena in very small microchannels, reported in numerous publications. These studies have revealed some extremely interesting phenomena concerning boiling in microchannels. Peng and Wang [10, 45] and Peng et al. [46] used rectangular ducts with sides 0.6 0.7, 0.4 0.7, and 0.2 0.7 mm2 , which were heated from three sides while the fourth had a Pyrex glass cover. Deionized water and methanol were used as test uids. An interesting phenomenon was that no bubbles were seen inside the microchannels even at high heat uxes. However, in the manifold, streams of bubbles were seen by the outlet from the microchannels. Although no bubbles were formed inside the channels, the authors still found the temperatureheat ux relation (boiling curve) indicating that the process was in the nucleate boiling regime. From these experiments they concluded that for bubble nucleation there is a critical minimum space, which they called evaporating space. The boiling without visible bubble nucleation was called ctitious boiling. The heat transfer coef cients in this boiling regime seemed to be independent of the ow velocity and of the subcooling. Peng et al. [47] reported on continued studies of subcooled boiling in 12 different rectangular cannel con gurations with hydraulic diameters ranging from 0.15 to

HEAT TRANSFER IN MICROCHANNELS

169

Downloaded by [Universiti Teknologi Malaysia] at 10:27 01 October 2011

0.343 mm (0.1 0.3 mm2 to 0.4 0.3 mm2 ) using water, methanol, and mixtures of these as working uids. A direct comparison of the in uence of the channel diameter was dif cult to make as the authors used the area of the base plates containing the channels when de ning the heat ux. The pure uids were found to give the highest heat transfer coef cients. Peng et al. [48] presented an analytical model explaining the phenomenon of ctitious boiling reported in previous articles. They started by stating that if the channel size is smaller than the evaporating space, then ctitious boiling may be induced. The physical explanation to this may be that internal evaporation and bubble growth have not yet been realized or there may exist countless microbubbles within the liquid that cannot be visualized by ordinary means. The authors then presented their analysis, based on thermodynamic phase stability theory, and arrived at a dimensionless parameter, Nmb , which gives a condition for nucleation in microchannels: Nmb 5 h f g av (v 9 9 v 9 ) q dh (18)

(c is an empirical constant) If Nmb 1 nucleation should be expected, while at larger values of Nmb ctitious boiling will occur. The uid is then in a highly non-equilibrium state with an exceptional capability to absorb, transfer and transport thermal energy. A correlation for the superheat necessary for nucleation is also presented: D tsup 4 A ts (v 9 9 v9 ) h f g dh (19)

[A ( 280) is an empirical constant] Fictitious boiling has also been explained in terms of an interphase propagation and superposition model by Hu et al. [49]. Recently, Peng et al. [50] reported on an investigation of the in uence of pressure perturbations on the development of high-energy clusters in a superheated liquid. They proposed that the pressure wave from the initial growth of a cluster could be re ected in the walls of a microchannel and thereby suppress the growth of the emerging bubble. A criterion for the development of ctitious boiling would then be L < 0.5 c s
c

(20)

where L is the scale of the microchannel, c is the speed of sound, and s c is the time required for the initial development phase of the bubble embryo ( 10 l s). Conclusions, Evaporation From the information above, it may be concluded that ow boiling is governed mainly by nucleate boiling mechanisms in the diameter range below ~ 4 mm. A pool boiling correlation such as Coopers could be expected to give reasonable but conservative values for the heat transfer coef cient as long as the critical heat ux is not reached.

170

B. PALM

At diameters below ~ 1 mm the data of Peng and co-workers show that a new phenomenon, ctitious boiling, could be expected. In this boiling mode no bubbles are detected, but the heat exchange and temperature differences are as in nucleate boiling. CONDENSATION Very little information is found in the literature on condensation in microtubes. A reason for the lower interest for condensation as compared to evaporation is that the main application for two-phase micro heat exchangers is for cooling of electronics. For this application the evaporator is placed in contact with the electronic components and is therefore the more critical part. Condensation heat transfer is governed by the thickness of the liquid lm at the cooled wall. In two-phase ow condensation there are two mechanisms acting to decrease the lm thickness: gravity and shear stress. In general, gravity should be expected to have less in uence as the channel diameter is reduced. Shear stress could be expected to increase with decreasing diameter. Also, surface tension forces could be of importance. This is clearly seen from the Laplace equation for a cylindrical interphase, pv pl 5 r (21)

Downloaded by [Universiti Teknologi Malaysia] at 10:27 01 October 2011

If the tube radius is small there will be a noticeable pressure difference between the vapor and the liquid phases, which will in uence the saturation temperature and also might tend to increase the lm thickness. Yang and Webb [51, 52] investigated heat transfer and pressure drop as R-12 condensed in two different at extruded aluminum tubes. These tubes had external dimensions 16 3 mm and a wall thickness of 0.5 mm. At the inside the tubes were divided into four channels by thin longitudinal walls. The difference between the two tubes was that one had smooth internal surfaces while the other had longitudinal micro ns, 0.2 mm in height. The tubes were rst tested in one-phase ow and the results were compared to predictions of the Petukhov equation using the total surface area as a base for the heat transfer coef cient. The agreement was very good, especially for the plain tube. In condensation it was found that the heat transfer coef cient increased with increasing vapor quality, increasing mass ux, and with increasing heat ux. The results were compared to predictions by the following correlation by Akers et al. [53]: hd 5 kl 0.0265 Re0.8 Pr 1.3 eq l Reeq > 50,000 (22)

The agreement was good for the plain tube at low mass ux (400 kg/m2 s) but 1020% lower than predicted at higher mass ux (1,000 kg/m2 s). There was also a slight dependence on heat ux not accounted for in the correlation. For the micro n tube, the enhancement ratio h micro n h plain was found to be higher than the surface area ratio A micro n A plain . The additional enhancement was thought to be due to surface tension acting to pull the liquid into the valleys in between the ns, thereby reducing lm thickness on the n tips.

HEAT TRANSFER IN MICROCHANNELS

171

The friction factors in one-phase ow were compared to the Blasius equation and found to be 14% higher (plain tube) and 36% higher (micro n tube) than predicted. For two-phase ow, the pressure drops were compared to the Martinelli two-phase multiplier model and to a correlation proposed by Akers et al. [53]. The Martinelli method did not correlate the data well, while the comparison with Akers et al. correlation was very good. Akers et al. de ned the friction factor by f 5 where G eq 5 D p G 2 2q eq dh 5 4L D p Re2 l eq
2 l

2q

3 dh 4L

(23)

Downloaded by [Universiti Teknologi Malaysia] at 10:27 01 October 2011

" G (1 xm) 1 xm q q

l v

0.5

# (24)

The two-phase friction factor is then calculated from the single-phase friction factor by the relation f 5 0.435 Re0.12 (25) eq fl Yan and Lin [58] measured heat transfer and pressure drop of R-134a condensing in an array of 28 tubes with 2-mm inner diameter connected in parallel. The results were compared to correlations from the literature and new correlations for the heat transfer coef cient and the friction factor were proposed. The heat transfer coef cients were compared to average values for a long, 8-mm tube according to Eckels and Pate [54] and were found to be lower than expected at vapor fractions below about 0.4 and higher than expected above this value. When averaged over the complete range of vapor fractions, the values were about 10% higher than expected from the correlation. The following equation was proposed based on the authors results: hd Pr l kl
0.33

Bo0.3 Re 5

6.48 Re1.04 eq

(26)

The friction factor, de ned as by Akers [Eq. (23)], was correlated by the equation f tp 5 498.3 Reeq1.074 (27)

Wang and Du [55] recently proposed an analytical model of condensation in tubes. They compared the results of the model with measurements conducted with four copper tubes with inside diameters 1.94, 2.8, 3.95, and 4.98 mm at different inclination angles using water as test uid. It was found that the inclination angle had very little in uence on heat transfer in the smallest diameter tube, but that this in uence became more obvious in the larger tubes. The Reynolds numbers had opposing effects on the Nusselt number, depending on the size of the tube. In the two smallest tubes Nu increased with Re, while the lowest Re gave the highest Nu in the two larger tubes. This shows that gravity is more important in large-diameter tubes, where strati cation may lead to increased heat transfer.

172

B. PALM

Downloaded by [Universiti Teknologi Malaysia] at 10:27 01 October 2011

In general, the Nusselt numbers were approximately the same at equal vapor quality and equal Reynolds number in all tubes, ranging from about 150 at the tube intlet to 50 at the outlet. That surface tension forces may have a large in uence on condensation in microchannels was also shown by Tengblad and Palm [56, 57]. In this test, vertical, square, 2 2 mm channels, connected at the top and the bottom, without forced convection, were used for condensing different refrigerants in a closed-loop thermosyphon. The heat transfer coef cients were compared to the Nusselt correlation for vertical plates. It was found that the measured heat transfer coef cients were at least three times higher than expected from the Nusselt theory. The increase was thought to be the result of surface tension forces attracting the liquid to the corners of the channel, thereby decreasing the lm thickness on the main part of the walls and also facilitating the drainage of liquid through the streams forming in the corners. Conclusions, Condensation As a conclusion it may be stated that very little data are available in the literature concerning condensation inside microchannels. For forced-convection condensation the correlations of Akers et al. for heat transfer and pressure drop seem to be in good agreement with experimental data for 2-mm channels. No information is available for smaller channels at this time. Surface tension effects should be expected to be more dominant, while the in uence of gravity is expected to be smaller than for larger tubes. CONCLUDING REMARK The literature concerning heat transfer and pressure drop in one- and two-phase ow in microchannels has been reviewed. It can be concluded that there are still many open questions to be answered before reliable design tools are available in the form of correlating equations for heat transfer and pressure drop. More research is therefore needed in this relatively new and exciting eld. REFERENCES
1. J. Goodling, Microchannel Heat Exchangers, SPIE Vol 1997 High Heat Flux Enginering II, pp. 6682, 1993. 2. A. B. Duncan and G. P. Peterson, Review of Microscale Heat Transfer, ASME Appl. Mech. Rev., vol. 47, no. 9, pp. 397428, 1994. 3. X. F. Peng and B. X. Wang, Liquid Flow and Heat Transfer in Microchannels with/without Phase Change, Proc. 10th Int. Heat Transfer Conf., Brighton, England, 1418 August 1994, in Heat Transfer 1994, Taylor & Francis, Washington, DC, vol. 1 (Sk-11), pp. 159177, 1994. 4. X. F. Peng and B. X. Wang, Forced-Convection and Boiling Characteristics in Microchannels, Proc. 11th Int. Heat Transfer Conf., Kyongyu, Korea, vol. 1, pp. 371390, 1998. 5. D. K. Bailey, T. A. Ammel, R. O. Warrington, and T. I. Savoie, Single Phase Forced Convection Heat Transfer in MicrogeometriesA Review, IECEC Conference, ES-396, Orlando, FL, ASME, 1995. 6. D. B. Tuckerman and R. F. W. Pease, High Performance Heat Sinking for VLSI, IEEE Electron Dev. Lett. EDL-2, pp. 126129, 1981.

HEAT TRANSFER IN MICROCHANNELS

173

7. P. Wu and W. A. Little, Measurement of Friction Factors for the Flow of Gases in Very Fine Channels Used for Microminiature Joule-Thompson Refrigerators, Cryogenics, vol. 23, no. 5, pp. 273277, 1983. 8. P. Wu and W. A. Little, Measurement of the Heat Transfer Characteristics of Gas Flow in Fine Channel Heat Exchangers Used for Microminiature Refrigerators, Cryogenics, vol. 24, pp. 415420, 1984. 9. X. F. Peng and G. P. Peterson, Forced Convection Heat Transfer of Single-Phase Binary Mixtures through Microchannels, Exp. Thermal Fluid Sci., vol. 12, pp. 98104, 1996. 10. X. F. Peng and B. X. Wang, Forced Convection and Flow Boiling Heat Transfer for Liquid Flowing through Microchannels, Int. J. Heat Mass Transfer, vol. 36, no. 14, pp. 34213427, 1993. 11. B. X. Wang and X. F. Peng, Experimental Investigation of Liquid Forced Convection Heat Transfer through Microchannels, Int. J. Heat Mass Transfer, vol. 37 (suppl. 1), pp. 7382, 1994. 12. J. M. Cuta, C. E. McDonald, and A. Shekarriz, Forced Convection Heat Transfer in Parallel Channel Array Microchannel Heat Exchanger, ASME PID-Vol. 2/HTD-Vol. 338, Advances in Energy Ef ciency, Heat/Mass Transfer Enhancement, pp. 1723, 1996. 13. T. M. Harms, M. J. Kazmierczak, and F. M. Gerner, Developing Convective Heat Transfer in Deep Rectangular Microchannels, Int. J. Heat Fluid Flow, vol. 20, pp. 149157, 1999. 14. G. M. Mala and D. Li, Flow Characteristics of Water in Microtubes, Int. J. Heat Fluid Flow, vol. 20, pp. 142148, 1999. 15. S. M. Flockhart and R. S. Dhariwal, Experimental and Numerical Investigation into the Flow Characteristics of Channels Etched in (100) Silicon, Trans. ASME: J. Fluids Eng., vol. 120, pp. 291295, 1998. 16. W. Qu, G. M. Mala, and D. Li, Pressure-Driven Water Flows in Trapezoidal Silicon Microchannels, Int. J. Heat Mass Transfer, vol. 43, pp. 353364, 2000. 17. I. Papautsky, J. Brazzle, T. Ammel, and A. B. Frazier, Laminar Fluid Behavior in Microchannels Using Micropolar Fluid Theory, Sensors and Actuators, vol. 73, pp. 101108, 1999. 18. X. N. Jiang, Z. Y. Zhou, J. Yao, Y. Li, and X. Y. Ye, Micro- uid Flow in Microchannel, Proc. Transducers 95, Stockholm, Sweden, pp. 317320, 2529 June 1995. 19. P. Wilding, M. A. Shoffner, and L. J. Kircka, Manipulation and Flow of Biological Fluids in Straight Channels Micromachined in Silicon, Clin. Chem., vol. 40, pp. 4347, 1994. 20. C. P. Tso and S. P. Mahulikar, The Use of the Brinkman Number for Single Phase Forced Convective Heat Transfer in Microchannels, Int. J. Heat Mass Transfer, vol. 41, no. 12, pp. 1759 1769, 1998. 21. C. P. Tso and S. P. Mahulikar, Proc. 2nd IEEE Electronics Packaging Technology Conf., Singapore, pp. 126132, 1998. 22. C. P. Tso and S. P. Mahulikar, The Role of the Brinkman Number in Analysing Flow Transitions in Microchannels, Int. J. Heat Mass Transfer, vol. 42, pp. 18131833, 1999. 23. C. P. Tso and S. P. Mahulikar, Experimental Veri cation of the Role of Brinkman Number in Microchannels Using Local Parameters, Int. J. Heat Mass Transfer, vol. 43, pp. 18371849, 2000. 24. C. Yang, D. Li, and J. H. Hasliyah, Modeling Forced Liquid Convection in Rectangular Microchannels with Electrokinetic Effects, Int. J. Heat Mass Transfer, vol. 41, pp. 42294249, 1998. 25. A. Beskok and G. E. Karniadakis, Simulation of Slip-Flows in Complex Micro-geometries, ASME Proc. DSC, vol. 40, pp. 355370, 1992. 26. J.-M. Li, B.-X. Wang, and X.-F. Peng, Wall-Adjacent Layer Analysis for Developed-Flow Laminar Heat Transfer of Gases in Microchannels, Int. J. Heat Mass Transfer, vol. 43, pp. 839 847, 2000.

Downloaded by [Universiti Teknologi Malaysia] at 10:27 01 October 2011

174

B. PALM

27. J.-M. Li, B.-X. Wang, and X.-F. Peng, The Wall Effect for Laminar Flow through Microtubes, Proc. Int. Centre for Heat and Mass Transfer (ICHMT) Symp. on Molecular and Microscale Heat Transfer in Material Processing and Other Applications, Yokohama, Japan, pp. 5565, 15 Dec. 1996. 28. M. G. Cooper, Saturation Nucleate BoilingA Simple Model, Proc. 1st UK Natl. Conf. on Heat Transfer, vol. 2, pp. 785793, 1984. 29. B. Pierre, Vrmevergngen vid kokande kldmedier i horisontella rr, Kylteknisk tidskrift, no. 3, June 1957 (in Swedish). See also ASHRAE Handbook, Fundamentals, p. 4.7, 1997. 30. B. Palm and R. Khodabandeh, Optimum Design of Two Phase Thermosyphon Systems for Cooling of Electronics, Proc. InterPack99, Hawaii, June 1999. In ASME Advances in Electronic Packaging 1999, EEP-Vol. 26-2, pp. 491498, 1999. 31. G. M. Lazarek and S. H. Black, Evaporative Heat Transfer, Pressure Drop and Critical Heat Flux in a Small Vertical Tube, Int. J. Heat Mass Transfer, vol. 25, no. 7, pp. 945960, 1982. 32. Y. Fujita, H. Ohta, S. Uchida, and K. Nishikawa, Nucleate Boiling Heat Transfer and Critical Heat Flux in Narrow Space between Rectangular Surfaces, Int. J. Heat Transfer, vol. 31, no. 2, pp. 229239, 1988. 33. M. R. Aligoodarz and D. B. R. Kenning, Vapour Bubble Behaviour in a Single Narrow Channel, IMechE, C510/076/95, pp. 273276, 1995. 34. K. Cornwell and P. A. Kew, Evaporation in Microchannel Heat Exchangers, IMechE, C510/ 117/95, pp. 289293, 1995. 35. T. N. Tran, M. W. Wambsganss, and D. M. France, Small Circular- and Rectangular-Channel Boiling with Two Refrigerants, Int. J. Multiphase Flow, vol. 22, no. 3, pp. 485498, 1996. 36. C. Xia, W. Hu, and Z. Guo, Natural Convection Boiling in Vertical Rectangular Narrow Channels, Exp. Thermal Fluid Sci., vol. 12, pp. 313324, 1998. 37. J. Bonjour and M. Lallemand, Flow Patterns during Boiling in a Narrow Space between Two Vertical Surfaces, Int. J. Multiphase Flow, vol. 24, pp. 947960, 1998. 38. S. Lin and P. A. Kew, Characteristics of Two-Phase Flow and Heat Transfer in Small Tubes, IMechE, C565/040/99, 1999. 39. Y.-Y. Yan and T.-F. Lin, Evaporation Heat Transfer and Pressure Drop of Refrigerant R134a in a Small Pipe, Int. J. Heat Mass Transfer, vol. 41, pp. 41834194, 1998. 40. I. Hapke, H. Boye, and J. Schmidt, Onset of Nucleate Boiling in Minichannels, Int. J. Thermal Sci., vol. 39, pp. 505513, 2000. 41. Z. Y. Bao, D. F. Fletcher, and B. S. Haynes, Flow Boiling Heat Transfer of Freon R11 and HCFC123 in Narrow Passages, Int. J. Heat Mass Transfer, vol. 43, pp. 33473358, 2000. 42. K. A. Triplett, S. M. Ghiaasiaan, S. I. Abdel-Khalik, and D. L. Sadowski, Gas-Liquid TwoPhase Flow in Microchannels, Part I: Two-Phase Flow Patterns, Int. J. Multiphase Flow, vol. 25, pp. 377394, 1999. 43. K. A. Triplett, S. M. Ghiaasiaan, S. I. Abdel-Khalik, and D. L. Sadowski, Gas-Liquid TwoPhase Flow in Microchannels, Part II: Void Fraction and Pressure Drop, Int. J. Multiphase Flow, vol. 25, pp. 395410, 1999. 44. Y. Zhao, M. Molki, M. M. Ohadi, and S. V. Dessiatoun, Flow Boiling of CO2 in Microchannels, ASHRAE Trans., vol. 106, pt. 1, DA-00-2-1, 2000. 45. X. F. Peng and B. X. Wang, Cooling Characteristics with Microchanneled Structures, J. Enhanced Heat Transfer, vol. 1, no. 4, pp. 315326, 1994. 46. X. F. Peng, B. X. Wang, G. P. Peterson, and H. B. Ma, Experimental Investigation of Heat Transfer in Flat Plates with Rectangular Microchannels, Int. J. Heat Mass Transfer, vol. 38, no. 1, pp. 127137, 1995. 47. X. F. Peng, G. P. Peterson, and B. X. Wang, Flow Boiling of Binary Mixtures in Microchanneled Plates, Int. J. Heat Mass Transfer, vol. 39, no. 6, pp. 12571264, 1996. 48. X. F. Peng, H. Y. Hu, and B. X. Wang, Boiling Nucleation during Liquid Flow in Microchannels, Int. J. Heat Mass Transfer, vol. 41, no. 1, pp. 101106, 1998.

Downloaded by [Universiti Teknologi Malaysia] at 10:27 01 October 2011

HEAT TRANSFER IN MICROCHANNELS

175

Downloaded by [Universiti Teknologi Malaysia] at 10:27 01 October 2011

49. H. Y. Hu, G. B. Peterson, X. F. Peng, and B. X. Wang, Interphase Fluctuation Propagation and Superposition Model for Boiling Nucleation, Int. J. Heat Mass Transfer, vol. 41, pp. 3483 3489, 1998. 50. X. F. Peng, D. Liu, D. J. Lee, Y. Yan, and B. X. Wang, Cluster Dynamics and Fictitious Boiling in Microchannels, Int. J. Heat Mass Transfer, vol. 43, pp. 42594265, 2000. 51. C.-Y. Yang and R. L. Webb, Condensation of R-12 in Small Hydraulic Diameter Extruded Aluminum Tubes with and without Micro- ns, Int. J. Heat Mass Transfer, vol. 39, no. 4, pp. 791800, 1996a. 52. C.-Y. Yang and R. L. Webb, Friction Pressure Drop of R-12 in Small Hydraulic Diameter Extruded Aluminum Tubes with and without Micro- ns, Int. J. Heat Mass Transfer, vol. 39, no. 4, pp. 801809, 1996. 53. W. W. Akers, H. A. Deans, and O. K. Crosser, Condensation Heat Transfer within Horizontal Tubes, Chem. Eng. Prog. Symp. Ser., vol. 55, no. 29, pp. 171176, 1959. 54. S. J. Eckels and M. B. Pate, An Experimental Comparison of Evaporation and Condensation Heat Transfer Coef cients for HFC-134a and CFC-12, Int. J. Refrig., vol. 14, pp. 7077, 1991. 55. B.-X.Wang and X.-Z. Du, Study on Laminar Film-wise Condensation for Vapor Flow in an Inclined Small/Mini-Diameter Tube, Int. J. Heat Mass Transfer, vol. 43, pp. 18591868, 2000. 56. N. Tengblad and B. Palm, Flow Boiling and Film Condensation Heat Transfer in Narrow Channels of Thermosiphons for Cooling of Electronic Components, Proc. Eurotherm, Sem. No. 45, Leuven, Belgium, September 1995. 57. N. Tengblad and B. Palm, External Two-Phase Thermosiphons for Cooling of Electronic Components, Int. J. Microcircuits Electronic Packaging, vol. 19, no. 1, pp. 2229, 1996. 58. Y.-Y. Yan and T.-F. Lin, Condensation Heat Transfer and Pressure Drop of Refrigerant R-134a in a Small Pipe, Int. J. Heat Mass Transfer, vol. 42, pp. 697708, 1999.

You might also like