Anda di halaman 1dari 9

Heat Transfer Mechanisms During Flow Boiling in Microchannels

The forces due to surface tension and momentum change during evaporation, in conjunction with the forces due to viscous shear and inertia, govern the two-phase ow patterns and the heat transfer characteristics during ow boiling in microchannels. These forces are analyzed in this paper, and two new nondimensional groups, K1 and K2 , relevant to ow boiling phenomenon are derived. These groups are able to represent some of the key ow boiling characteristics, including the CHF. In addition, a mechanistic description of the ow boiling phenomenon is presented. The small hydraulic dimensions of microchannel ow passages present a large frictional pressure drop in single-phase and two-phase ows. The small hydraulic diameter also leads to low Reynolds numbers, in the range 100 1000, or even lower for smaller diameter channels. Such low Reynolds numbers are rarely employed during ow boiling in conventional channels. In these low Reynolds number ows, nucleate boiling systematically emerges as the dominant mode of heat transfer. The high degree of wall superheat required to initiate nucleation in microchannels leads to rapid evaporation and ow instabilities, often resulting in ow reversal in multiple parallel channel conguration. Aided by strong evaporation rates, the bubbles nucleating on the wall grow rapidly and ll the entire channel. The contact line between the bubble base and the channel wall surface now becomes the entire perimeter at both ends of the vapor slug. Evaporation occurs at the moving contact line of the expanding vapor slug as well as over the channel wall covered with a thin evaporating lm surrounding the vapor core. The usual nucleate boiling heat transfer mechanisms, including liquid lm evaporation and transient heat conduction in the liquid adjacent to the contact line region, play an important role. The liquid lm under the large vapor slug evaporates completely at downstream locations thus presenting a dryout condition periodically with the passage of each large vapor slug. The experimental data and high speed visual observations conrm some of the key features presented in this paper. DOI: 10.1115/1.1643090 Keywords: Boiling, Bubble Growth, Heat Transfer, Interface, Microscale

Satish G. Kandlikar
e-mail: sgkeme@rit.edu Thermal Analysis and Microfluidics Laboratory, Department of Mechanical Engineering, Rochester Institute of Technology, Rochester, NY 14623

Introduction
It has been well recognized that the ow boiling heat transfer consists of a nucleate boiling component and a convective boiling component, e.g., Schrock and Grossman 1. The convective boiling component results from the convective effects, whereas the nucleate boiling component results from the nucleating bubbles and their subsequent growth and departure at the heated surface. The nucleate boiling and the convective boiling mechanisms cannot be distinguished precisely, as they are closely interrelated. The nucleation, growth and subsequent departure of bubbles from a heated wall constitute the major stages of a bubble ebullition cycle in pool boiling. The heat transfer mechanisms associated with pool boiling include: transient conduction, lm vaporization, microconvection, contact line heat transfer, and to a lesser degree, the vaporization condensation, and Marangoni convection. The bubble dynamics under ow boiling conditions are affected in such a way as to shift toward smaller nucleation cavity sizes and smaller departure bubble diameters. In this paper, the forces governing the liquid vapor interface in microchannels during ow boiling are analyzed and two new dimensionless groups are identied. The role of nucleate boiling is re-examined and the heat transfer mechanisms are discussed. The high-speed ow visualization conducted by the author and his co-workers and the experimental heat transfer data from literature are used in arriving at the mechanistic description of the ow boiling phenomena in microchannels. In this paper, the channel classication proposed by Kandlikar
Contributed by the Heat Transfer Division for publication in the JOURNAL OF HEAT TRANSFER. Manuscript received by the Heat Transfer Division February 20, 2003; revision received September 24, 2003. Associate Editor: M. K. Jensen.

and Grande 2 is used. This classication should be used as a mere guide indicating the size range, rather than rigid demarcations based on specic criteria. In reality, the effects will depend on uid properties and their variation with pressure changes. Undoubtedly, the actual ow conditions, such as single phase liquid or gas ow, or ow boiling or ow condensation, will present different classication criteria. The classication by Kandlikar and Grande is based on rarefaction effects for common gases around 1 atmospheric pressure. At smaller dimensions below about 10 m, molecular effects begin to be important, again subject to specic operating conditions. Molecular nanochannels are still very large compared to the molecular dimensions, but the intermolecular forces begin to play an important role especially in the near wall region. In two-phase applications, recently Kawaji and Chung 3 reported signicant departure for their 100 m microchannels from the linear relationship between volumetric quality versus void fraction proposed by Ali et al. 4 for narrow channels. Conventional channels: 3 mm Minichannels: 3 mm Dh 200 m Microchannels: 200 m Dh 10 m Transitional Channels: 10 m Dh 0.1 m 100 nanometer, nm, or 1000 Transitional Microchannels: 10 m Dh 1 m Transitional Nanochannels: 1 m Dh 0.1 m Molecular Nanochannels: 0.1 m Dh

Review of Available Non-Dimensional Groups Related to Two-Phase Flow and Interfacial Phenomena
Before deriving the relevant new non-dimensional groups in microchannels, the existing groups commonly applied in the twoTransactions of the ASME

8 Vol. 126, FEBRUARY 2004

Copyright 2004 by ASME

Table 1 Non-dimensional Groups relevant to two-phase ow studies in microchannels Non-dimensional number Martinelli parameter, X dp dp X2 dz F dz F Signicance Relevance to Microchannels

Convection Number, Co Co (1 x )/ x 0.9 G / L 0.5 Boiling number, Bo q Bo G hfg Bond number, Bo g L G D 2 Bo tvo s Number, Eo Eo g L G L 2 Eo Capillary number, Ca V Ca Ohnesorge Number, Z Z L 1/2 Weber Number, We L G2 We Jakob Number, Ja L c p,L T Ja hfg V

Groups based on Empirical Considerations Because of its success in predicting pressure drop in Represents the Ratio of frictional pressure drops with compact evaporator geometries, it is expected to be a useful liquid and gas ow. It has been employed successfully parameter in microchannels as well. in two-phase pressure drop models in simple as well as complex geometries Co is a modied Martinelli parameter, introduced by Its direct usage beyond ow boiling correlation may be Shah 5 in correlating ow boiling data limited. Since it combines two important ow parameters, q and G , Heat ux is non-dimensionalized with mass ux and it will be used in empirical treatment of ow boiling. latent heat. It is not based on any fundamental considerations. Groups based on Fundamental Considerations Since the effect of gravitational force is expected to be Bo represents the ratio of buoyancy force to surface small, Bo is not expected to play an important role in tension force. Used in droplet atomization and spray microchannels. applications. tvo s number is similar to the Bond number, except Eo that the characteristic dimension L could be D h or any other physically relevant parameter. Ca represents the ratio of viscous to surface tension forces. It is useful in analyzing the bubble removal process. Z represents the ratio of the viscous force to the square root of the product of the inertia and the surface tension forces. It is used in analyzing liquid droplets and droplet atomization processes. We represents the ratio of the inertia to the surface tension forces. For ow in channels, D h is used in place of L . Ja represents the ratio of the sensible heat for a given volume of liquid to heat or cool through T in reaching its saturation temperature, to the latent heat required in evaporating the same volume of vapor. Similar to Bo, Eo is not expected to be important in microchannels except at very low ow velocities and vapor fractions. Ca is expected to play a critical role as both the surface tension and the viscous forces are important in microchannel ows. The combination of the three forces into one masks the individual effects of each force. Although Z may come out as an important variable, it may not be suitable in the early stages of microchannel research. We I useful in studying the relative effects of surface tension and inertia forces on ow patterns in microchannels. Ja may be used as an important parameter in studying the effects of liquid superheat prior to initiation of nucleation in microchannels. It may also be useful in studying the subcooled boiling conditions.

phase and ow boiling applications are reviewed in this section. Non-dimensional groups are useful in arriving at key basic relations among system variables that are valid for different uids under different operating conditions. Some of these groups have been derived empirically, often on the basis of experimental data. Fundamental considerations of the governing forces and their mutual interactions lead to non-dimensional groups that provide important insight into the physical phenomena. A listing of both types of non-dimensional groups, based on empirical considerations and on theoretical analysis of forces, is given in Table 1. A brief description of their signicance along with their relevance to ow boiling in microchannels is also included in the table. The non-dimensional groups based on empirical considerations have been proposed in literature on the basis of extensive data analysis. Their extension to other systems requires rigorous validation, often requiring modications of constants or exponents. Convection number and boiling number fall under this category. Although the Martinelli parameter is derived from fundamental considerations of the gas and the liquid phase friction pressure gradients, it is used extensively as an empirical group in correlating experimental results on pressure drop, void fraction, and is some cases, heat transfer as well. It can be seen from Table 1 that among the groups derived from fundamental considerations, Capillary number, Ca, and Weber number, We, are important in microchannel two-phase ows. They represent the ratios of viscous and surface tension forces, and inertia and surface tension forces respectively. The ratio of inertia to viscous forces, represented by Reynolds number, Re, is extensively used in single phase ows; it is also expected to have an important role in two-phase ows, as the single-phase heat transfer and pressure drop characteristics are often found to be useful in predicting the two-phase ow behavior. The Bond number represents the ratio of the gravitational to surface tension forces on a droplet or a bubble. Departure bubble Journal of Heat Transfer

diameter is sometimes used as the characteristic dimension in the denition of the Bond number in pool boiling. Since the gravitational forces are expected to be small compared to the surface tension and viscous forces, usefulness of Bond number in microchannel application is limited. Additionally, there are a number of non-dimensional groups used in literature for correlating different boiling phenomena. For example, Stephan and Abdelsalam 6 utilized eight dimensionless groups in developing a comprehensive correlation for saturated pool boiling heat transfer. Similarly, Kutateladze 7 combined the critical heat ux with other parameters through dimensional analysis to obtain a non-dimensional group. However, only those non-dimensional groups that are derived from some basic considerations are reviewed here. Forces Acting on a Liquid-Vapor Interface Due to Evaporation Momentum Change, Inertia and Surface Tension. The non-dimensional groups employed so far in literature for analyzing the ow patterns during ow boiling have been similar to those employed in the adiabatic two-phase ow applications. The Jakob number is used in analyzing the bubble growth phenomenon. The Weber number has been recently introduced in ow pattern maps for microchannels and minichannels. The effect of heat ux appears only in the Boiling number, which is used in heat transfer correlations along with the density ratio; the Boiling number has not been used in ow pattern maps. Heat ux has a signicant effect on the two-phase structure during ow boiling. A higher heat ux results in a more rapid bubble growth leading to quickly lling the channel with a vapor slug. The forces generated due to rapid evaporation become signicant for microchannels, as the growing bubbles interact with the channel walls. Further discussion on this aspect is presented in the following sections. As a nucleating bubble grows, the difference in the densities of FEBRUARY 2004, Vol. 126 9

Similarly, the force per unit length due to the surface tension force is given by

cos FS

(3)

where is the appropriate contact angle, advancing or receding, depending on the direction of the motion of the liquid into or away from the interface respectively. The net gravitational force due to buoyancy may be represented in a similar manner as

L G gD 2 Fg

(4)

Fig. 1 Forces acting on a vapor bubble growing on a heater surface in a liquid pool, only half of a symmetric bubble shown

The gravitational force is generally quite small compared to the forces due to inertia and the momentum change during ow boiling in microchannels. The viscous forces are not considered here, although they are important in determining the ow stability in microchannels. Their effect is represented through the Capillary number and the Reynolds number. New Non-Dimensional Groups Relevant to Flow Boiling in Microchannels. The relative magnitudes of the forces due to evaporation momentum, inertia, and surface tension, introduced in Eqs. 13, inuence the two-phase structure during ow boiling. The ratios of these forces, taken two at a time, yield the relevant non-dimensional groups in ow boiling. New Non-Dimensional Group K1. This group represents the ratio of the evaporation momentum force, Eq. 1, and the inertia force, Eq. 2. It is given by q 2D h fg G q K1 G 2D Gh f g L

the two phases causes the vapor phase to leave at a much higher velocity than the corresponding liquid velocity toward the receding and evaporating interface. The resulting change in momentum due to evaporation introduces a force on the interface. The magnitude of this force is highest near the heater surface due to a higher evaporation rate in the contact line region between the bubble base and the heater surface as illustrated in Fig. 1. For the case of saturated ow boiling, the heat ux at the wall results in an evaporation mass ux of q / h f g . The resulting momentum change introduces a force on the liquid interface that can be expressed as described below per unit characteristic dimension, either based on the bubble diameter as shown in Fig. 1, or based on the channel diameter as shown in Fig. 2. This force was rst introduced by Kandlikar 8 in developing a CHF model based on the force balance at the contact line at the base of a bubble on the heater surface. Thus force per unit length due to momentum change caused by evaporation at the interface is given by F M q qD q 1 h fg h fg G hfg

L G

(5)

D G

(1)

where D is the characteristic dimension. The other forces acting on the interface are due to the inertia of the ow and the surface tension in the contact line region. The forces per unit length are given by the respective equations below. The stress resulting from the inertia forces is given by V 2 . The resulting force per unit length due to the inertia is then given by

LV 2D FI

G 2D L

(2)

The non-dimensional group K1 includes the Boiling number and the liquid to vapor density ratio. The Boiling number alone does not represent the true effect of the evaporation momentum, and its coupling with the density ratio is important in representing the evaporation momentum force. A higher value of K1 indicates that the evaporation momentum forces are dominant and are likely to alter the interface movement. As an example, consider the bubble growth shown in Fig. 3 inside a 200 m square microchannel, Steinke and Kandlikar, 9. Each frame in Fig. 3 is 8 ms apart. The bubble grows rapidly aided by the evaporation at the liquid-vapor interface. The high evaporation rate causes the interface to move rapidly downstream. The upstream interface also experiences the force due to evaporation momentum and the increased pressure inside the bubble, causing it to move against the ow direction in one of the channels as shown. The ow reversal occurs at high heat uxes and has been observed in 1 mm square minichannels as well by Kandlikar et al. 10 and Kandlikar 11. Further discussion on the actual values of K1 employed in microchannels and minichannels is presented in later sections. New Non-Dimensional Group K2. This new group represents the ratio of the evaporation momentum force, Eq. 1, and the surface tension force, Eq. 3. It is given by q 2D h fg G q K2 hfg

D G

(6)

Fig. 2 Forces acting on a liquid-vapor interface that covers the entire channel cross section

The contact angle is not included in the above ratio. The actual force balance in a given situation may involve more complex dependence on contact angles and surface orientation. However, it should be recognized that the contact angles play an important role in bubble dynamics and contact line movement and need to be included in a comprehensive analysis. The non-dimensional group K2 governs the movement of the interface at the contact Transactions of the ASME

10 Vol. 126, FEBRUARY 2004

Fig. 4 Variation of K2,CHF with diameter for ow boiling CHF data of Vandervort et al. 12 for different values of G in kgm2 s

Fig. 3 Rapid bubble growth causing reversed ow in a parallel microchannel. Water ow is from left to right, single channel shown, t a 0 ms, b 8 ms, c 16 ms, d 24 ms, e 32 ms, f 40 ms, g 48 ms, and h 56 ms, Steinke and Kandlikar 9.

line. The high evaporation momentum force causes the interface to overcome the retaining surface tension force. This group was effectively used by Kandlikar 8 in developing a model for CHF in pool boiling. The characteristic dimension D was replaced with the departure bubble diameter. The resulting expression for the CHF in pool boiling from a smooth planar surface oriented at an angle of with the horizontal was obtained as
1/2 q CHF h f g G

diameter for three different mass uxes. The result is shown in Fig. 4. The value of K2,CHF for each mass ux set is seen to be quite constant, considering the scatter present in the original data. This is quite remarkable, and further, as expected, a dependence of the respective K2,CHF values on the mass ux is seen since K2,CHF does not include the inertia related forces. Further CHF modeling needs to include the mass ux effect. Incidentally, K2,CHF is seen to increase with mass ux as one would expect. 0.75 As a rst attempt, applying a mass ux correction factor K1 ,CHF to K2,CHF nearly collapses the data of Vandervort et al. 12 for different mass uxes on a single horizontal line. Figure 5 shows the resulting plot. Considering the scatter in the CHF data due to measurement uncertainties, the agreement is very encouraging. However, further experimental data need to be included in the comparison, and additional property effects including contact angles need to be considered while developing a CHF model for ow boiling in microchannels. Importance of Weber Number and Capillary Number in Microchannel Flow Boiling. The Weber number, We, and the Capillary number, Ca, should prove to be useful in determining the transition boundaries between different ow patterns. We and Ca are believed to inuence the CHF phenomenon as well. These groups are commonly employed in analyzing the gas-liquid adiabatic ows. They may prove to be useful in rening the mass ux 0.75 and viscous effects on CHF in the K2,CHF K1 ,CHF group shown in Fig. 5. Further work on this can be undertaken after reliable experimental data become available for CHF in microchannels with a wide variety of uids. Ranges of Non-Dimensionless Parameters Employed in Microchannel and Minichannel Flow Boiling Experiments in Literature. A large body of experimental data is available for ow boiling in conventional sized channels. There are only a handful

1 cos 16

2 1 cos cos 4

1/2

g L G 1/4

(7)

Alternatively, the value of K2 at CHF for the case of pool boiling may be written as K2,CHF


q CHF hfg
2

D 1 cos G 16
1/2

2 1 cos cos 4

g L G

(8)

The characteristic dimension D is simply half of the critical wavelength, alternatively, it may be taken as the departing bubble diameter, given by D2

g L G

1/2

(9)

Thus, K2,CHF for pool boiling is given by K2,CHF 2

1 cos 16

2 1 cos cos 4

(10)

In other words, the non-dimensional number K2 has a unique value for CHF in pool boiling and depends on the contact angle and the orientation angle of the heater with respect to the horizontal. This group is expected to be important in CHF analysis during ow boiling in microchannels and minichannels as well. The experimental CHF data of Vandervort et al. 12 during ow boiling is plotted with K2,CHF as a function of the hydraulic Journal of Heat Transfer

0.75 Fig. 5 Variation of K2,CHF K1,CHF with diameter for ow boiling CHF data of Vandervort et al. 12 for different values of G in kgm2 s

FEBRUARY 2004, Vol. 126 11

Fig. 6 Ranges of the new non-dimensional parameter K1 employed in various experimental investigations

Fig. 10 Ranges of Reynolds Number, Re, employed in various experimental investigations

of reliable experimental data sets available for minichannels and microchannels, e.g., 9 and 1321, although more are emerging in literature as the interest continues to grow in this eld. It is interesting to look at the ranges of dimensionless parameters employed in these experiments as a function of hydraulic diameter. Figures 6 10 show the ranges of non-dimensional parameters employed in several investigations. The ranges of K1 over the channel hydraulic diameters employed in literature are shown in Fig. 6. The 6 and 15-mm diameter data for R-11 by Chawla 13 covers a wide range of K1 . Looking into the experimental data sets, it is seen that the higher end values of K1 correspond to the nucleate boiling dominant region, while the lower end values of K1 indicate the convective boiling dominant region. The large diameter data sets cover a wide range, whereas it is seen that most

Fig. 7 Ranges of the new non-dimensional parameter K2 employed in various experimental investigations

of the minichannel and microchannel data sets fall toward the higher end of K1 values. Especially the data set of Yen et al. 14 is seen to fall at the very high end. Higher values of K1 are indicative of a relatively higher heat ux for a given mass ux value, and also of the increased importance of the evaporation momentum force. Figure 7 shows a similar plot for K2 plotted against D h . The maximum value of K2 not shown in the plot as no data are available is indicative of the CHF location. There are very few studies available on CHF in Minichannels, and practically none for the microchannels. This is an area where further research is warranted. Figure 8 shows the range of We in the data sets analyzed here. No appreciable differences are noted, and the data sets do not show any shift in the We range indicative of any change in the experimental conditions for smaller hydraulic diameter channels. Although We does not appear to be a signicant parameter in this plot, it may be representative of the inertia effects that are important in modeling the CHF in ow boiling. The Capillary number is expected to be important in the microchannel ows as both the viscous and the surface tension forces are expected to be important. Figure 9 shows the variation of Ca range with hydraulic diameter for representative data sets. It can be seen that there is a systematic shift toward higher values of Ca for smaller channel diameters. This indicates that the viscous forces become more important than the surface tension forces under ow boiling conditions in microchannels. This is somewhat of an unexpected result, as one would believe the dominance of the surface tension forces at smaller channel dimensions. Finally, Fig. 10 shows the range of all liquid ow Reynolds numbers employed in the data sets. As the channel diameter is decreased, a clear shift toward lower Reynolds number is seen. This points to the increased importance of the viscous forces in microchannels. Combining this observation with the discussion on the Capillary number in the above paragraph, it can be concluded that the viscous forces are extremely important in microchannel ows. The effect of lower Re on ow boiling heat transfer modeling will be discussed in the section on heat transfer coefcient prediction.

Fig. 8 Ranges of Weber Number, We, employed in various experimental investigations

Inuence of Channel Hydraulic Diameter on Flow Boiling Heat Transfer Mechanisms


Flow Patterns. Flow boiling in a channel is studied from the subcooled liquid entry at the inlet to a liquid-vapor mixture ow at the channel outlet. As the liquid ows through a microchannel, nucleation occurs over cavities that fall within a certain size range under a given set of ow conditions. Assuming that cavities of all sizes are present on the heater surface, the wall superheat necessary for nucleation may be expressed from the equations developed by Hsu and Graham 22 and Sato and Matsumura 23:

Fig. 9 Ranges of Capillary Number, Ca, employed in various experimental investigations

T sat,ONB

4 T satv f g h 1 kh f g

kh f g T sub 2 T satv f g h

(11)

12 Vol. 126, FEBRUARY 2004

Transactions of the ASME

Fig. 11 Wall superheat required to initiate nucleation in a channel with hydraulic diameter D h for saturated liquid conditions

For a conservative estimate, T sub 0 is taken in estimating the local superheat at the onset of nucleate boiling. Since the single phase liquid ow is in the laminar region, the Nusselt number is given by Nu hD h C k (12)

where C is a constant dependent on the channel geometry and the wall thermal boundary condition. Combining Eqs. 11 and 12 with zero subcooling, we get, T sat,ONB 8 T satv f g C D hh f g (13)
Fig. 12 High speed ow visualization of water boiling in a 197 m 1054 m channel showing rapid dryout and wetting phenomena with periodic passage of liquid and vapor slugs, each successive frame is 5ms apart, Kandlikar and Balasubramanian 25

Taking the case of a constant wall temperature boundary condition, representative of microchannels in a large copper block, the variation of T sat with D h is plotted in Fig. 11 for water at a saturation temperature of 100C and for R-134a at 30C. For channels larger than 1 mm, the wall superheat is quite small, but as the channel size becomes smaller, larger superheat values are required to initiate nucleation. For water in a 200 m channel, a wall superheat of 2C is required before nucleation can begin. The actual nucleation wall superheat may increase due to the absence of cavities of all sizes, while it may reduce due to a the presence of dissolved gases in the liquid, Kandlikar et al., 24, b corners in the channel cross section, and c pressure uctuations. From the above discussion, for channels smaller than 50 m, the wall superheat may become quite large, in excess of 10C with water, and above 2 3C for refrigerants. Flow boiling in channels smaller than 10 m will pose signicant challenges. The buildup of wall superheat presents another issue that is discussed in the following paragraph. Once the nucleation begins, the large wall superheat causes a sudden release of energy into the vapor bubble, which grows rapidly and occupies the entire channel. A number of investigators have considered this vapor slug as an elongated bubble. The rapid bubble growth pushes the liquid-vapor interface on both caps of the vapor slug at the upstream and the downstream ends, and leads to a reversed ow in the parallel channel case as the liquid on the upstream side is pushed back, and the other parallel channels carry the resulting excess ow. Kandlikar et al. 10 observed this reversed ow in 1-mm square parallel channels, while Steinke and Kandlikar 9 report a rapid ow reversal inside individual channels in a set of six parallel microchannels, D h 200 m. Reversed ow was also observed by Peles et al. 26. Kandlikar and Balasubramanian 25 observed only a small reverse displacement of the interface in a 197 m x 1054 m single microchannel. Due to the relatively high inlet stiffness of the water supply loop, the single microchannel does not show as dramatic reversed ow behavior as the parallel microchannels. In case of parallel channels, the other channels act as the upstream sections for any given channel. Figure 3 by Steinke and Kandlikar 9 shows a sequence of bubble formation and reversed ow from right to left in a channel. The images are 8 ms apart and show a bubble from its nucleation stage, as it occupies the entire channel, and eventually Journal of Heat Transfer

becomes a vapor slug that expands in both directions. The ow patterns are seen as periodic, with small nucleating bubbles growing into large bubbles that completely ll the tube and lead to rapid movement of the liquid-vapor interface along the channel. Figure 12 shows the periodic dryout and rewetting through an image sequence obtained with a ow of water in a single channel, 197 m 1054 m in cross-section, by Kandlikar and Balasubramanian 25. Here the bubble growth is extremely rapid, and the moving interface could not be precisely imaged. However, the ow goes from all liquid ow in frames a and b to a thin lm in frames c-d-e. In frames f-l, the liquid lm has been completely evaporated, and the channel walls are under a dryout condition. The liquid front is seen to pass through the channel in frame m, followed by all liquid ow in the last frame. The periodic wetting and rewetting phenomena were also observed by Hetsroni et al. 27 for parallel microchannels of 103129 m hydraulic diameters. Other ow patterns reported by Steinke and Kandlikar 9 include slug ow, annular ow, churn ow and dryout condition. A number of investigators, including Hetsroni et al. 27 and Zhang et al. 28 have also reported these ow patterns. The long vapor bubble occurring in a microchannel is similar to annular ow with intermittent slugs of liquid between two long vapor trains. The periodic phenomena described above is seen to be very similar to the nucleate boiling phenomena, with the exception that the entire channel acts like the area beneath a growing bubble, going through periodic drying and rewetting phases. The transient heat conduction under the approaching rewetting lm and the heat transfer in the evaporating meniscus region of the liquid-vapor interfaces in the contact line region exhibit the same characteristics as the nucleate boiling phenomena. Dynamic Advancing and Receding Contact Angles. The contact angles are believed to play an important role during dryout and rewetting phenomena during ow boiling in a microchannel. These angles are important in the events leading to the CHF as FEBRUARY 2004, Vol. 126 13

Fig. 14 Comparison of the Kandlikar 1,2,35 correlation and Eqs. 15 and 16 for the Nucleate Boiling Dominant Region with Yen et al. 14 data in the laminar region, plotted with heat transfer coefcient as a function of local vapor quality x , R-134a, D h 190 m, ReLO 73, G 171 kgm2 s, q 2 6.91 kWm

Fig. 13 Changes in contact angle at the liquid front during ow reversal sequence, water owing in a parallel microchannel with D h 200 m, Steinke and Kandlikar 9

well. The pool boiling CHF model by Kandlikar 8 incorporates these angles as seen from Eq. 7. During normal bubble growth, the dynamic receding contact angle comes into play. During the rewetting process, the liquid front is advancing with its dynamic advancing contact angle. As the CHF is approached, or as the ow reversal takes place due to ow instabilities, the liquid front changes its shape from the dynamic advancing angle to the dynamic receding angle. This event was captured by Steinke and Kandlikar 9 using high-speed imaging techniques and is displayed in Fig. 13 for the ow of water in a set of six parallel microchannels, each with a 200 m square cross section. The events in Fig. 13 show the changes in the contact angles. The liquid front is moving forward from left to right in frames a-d. In frame d, it stops moving forward. Note the dynamic advancing contact angle in this frame, which is greater than 90 deg. During frames d-i, it experiences rapid evaporation at the right side of the front interface and changes to the receding contact angle value, which is less than 90 deg. The liquid slug subsequently becomes smaller, making way to the vapor core with its residual liquid lm in frames i-m. Eventually, the liquid lm dries out completely as seen in frames n-p. Additional studies on the contact angle changes at evaporating interfaces are reported by Kandlikar 29. More detailed experiments are needed to further understand the precise effect of contact angles on CHF.

the function f 2 (FrLO ) is taken as 1.0. The single-phase, all-liquid ow heat transfer coefcient h LO is given by appropriate correlations given by Petukhov and Popov 32, and Gnielinski 33, depending on the Reynolds number in the turbulent liquid ow. The values of the uid-dependent parameter F Fl can be found in 34 or 35. For microchannels with the all-liquid ow in the laminar region, the corresponding all-liquid ow Nusselt number is given by NuLO h LO D h C, kL (16)

where the constant C is dependent on the channel geometry and the wall thermal boundary condition. Kandlikar and Steinke 34 showed that the use of Eq. 16 for h LO is appropriate in Eq. 15 when the all-liquid ow is in the laminar region. The use of turbulent ow correlation by Gnieliski was found to be appropriate for Re3000, while for Re1600, the use of laminar ow correlation yielded good agreement. In the transition region, a linear interpolation was recommended between the limiting ReLO values of 1600 and 3000. With further reduction in the Reynolds number, Re300, Kandlikar and Balasubramanian 35 showed that the ow boiling mechanism is nucleate boiling dominant and the use of only the nucleate boiling dominant expression in Eq. 15 for h T P ,NBD is appropriate in the entire range of vapor quality. Thus, Kandlikar and Balasubramanian 35 recommend the following scheme: h T P h T P ,NBD for ReLO 300 (17)

Heat Transfer Coefcients during Flow Boiling in Microchannels


As the channel size becomes smaller, the Reynolds number also becomes smaller for a given mass velocity, and the ow patterns exhibit characteristics similar to a bubble ebullition cycle as discussed in the previous sections. It is therefore expected that the ow boiling correlation for the nucleate boiling dominant region should work well here. Consider the ow boiling correlation by Kandlikar 30,31. h T P larger of

h T P ,NBD h T P ,CBD

(14)

h T P ,NBD 0.6683Co 0.2 1 x 0.8 f 2 FrLO h LO 1058.0Bo 0.7 1 x 0.8F Fl h LO h T P ,CBD 1.136Co 0.9 1 x 0.8 f 2 FrLO h LO 667.2Bo 0.7 1 x 0.8F Fl h LO where FrLO is the Froude number with all ow as liquid. Since the effect of Froude number is expected to be small in microchannels, 14 Vol. 126, FEBRUARY 2004 (15)

Using the above scheme, a recent data set published by Yen et al. 14 was tested. Figure 15 shows a comparison between the experimental data and the above correlation scheme. Here ReLO 73, and only the nucleate boiling dominant part of the correlation in Eq. 15 is used. At such low Reynolds numbers, the convective boiling does not become the dominant mechanism in the entire vapor quality range. In this plot, the rst few points show a very high heat transfer coefcient. This is due to the sudden release of the high liquid superheat following nucleation that causes rapid bubble growth in this region as seen in Fig. 14 as well. Steinke and Kandlikar 36 observed a similar behavior with their water data as shown in Fig. 15. The ow boiling mechanism is quite complex in microchannels as seen from the experimental data and ow visualization studies. Although simple models available in literature, e.g., 37, present a good analysis under certain idealized conditions, phenomena such as rapid bubble growth, ow reversal, and periodic dryout introduce signicantly greater complexities. Effect of parallel channels on heat transfer is another area where little experimental work is published. Transactions of the ASME

Conclusions
Flow boiling in microchannels is addressed from the perspective of interfacial phenomena. The following major conclusions are derived from this study: 1. After reviewing the non-dimensional groups employed in the two-phase applications, the forces acting on the liquid-vapor interface during ow boiling are investigated, and two new nondimensional groups K1 and K2 are derived. These groups represent the surface tension forces around the contact line region, the momentum change forces due to evaporation at the interface, and the inertia forces. 2. In a simple rst-cut approach, the combination of the nondimensional groups, K2K0.75 is seen to be promising in represent1 ing the ow boiling CHF data by Vandervort et al. 12. Further analysis and testing with a larger experimental data bank with a wide variety of uids is recommended in developing a more comprehensive model for CHF. In additon, the inuence of We, and Ca on the CHF needs to be studied further. 3. The nucleation criterion in ow boiling indicates that the wall superheat required for nucleation in microchannels becomes signicantly large ( 2 10C) as the channel hydraulic diameter becomes smaller, about 50100 m depending on the uid. The presence of dissolved gases in the liquid, sharp corners in the cross section geometry, and ow oscillations will reduce the required wall superheat, while the absence of cavities of appropriate sizes may increase the nucleation wall superheat. 4. Considerable similarities are observed between the ow boiling in microchannels and the nucleate boiling phenomena with the bubbles nucleating and occupying the entire channel. The process is seen to be very similar to the bubble ebullition cyle in pool boiling. 5. Heat transfer during ow boiling in microchannels is seen to be nucleate boiling dominant. 6. A description of the heat transfer mechanisms during ow boiling in microchannels is provided based on the above observations. Some of the key features, such s rapid bubble growth, ow reversal, nucleate boiling dominant heat transfer have been conrmed with high speed video pictures and experimental data. 7. Further experimental research is needed to generate more data for developing more accurate models and predictive techniques for ow boiling heat transfer and CHF in microchannels.

Fig. 15 Comparison of the Kandlikar 1,2,35 correlation and Eqs. 15 and 16 for the Nucleate Boiling Dominant Region with Steinke and Kandlikar 36 data in the laminar region, plotted with heat transfer coefcient as a function of local vapor quality x , water, D h 207 m, ReLO 291, G 157 kgm2 s, q 473 kWm2

Heat Transfer Mechanism during Flow Boiling in Microchannels


Based on the above discussion, the heat transfer mechanisms can be summarized as follows: 1. When the wall superheat exceeds the temperature required to nucleate cavities present on the channel walls, nucleate boiling is initiated in a microchannel. Absence of nucleation sites of appropriate sizes may delay nucleation. Other factors such as sharp corners, uid oscillations, and dissolved gases affect the nucleation behavior. The necessary wall superheat is estimated to be 2 10C for channels smaller than 50100 m hydraulic diameter with R-134a and water respectively at atmospheric pressure conditions. 2. Immediately following nucleation, a sudden increase in heat transfer coefcient is experienced due to release of the accumulated superheat in the liquid prior to nucleation. This leads to rapid evaporation, sometimes accompanied by ow reversal in parallel channels. 3. The role of the convective boiling mechanism is diminished in microchannels. The nucleate boiling plays a major role as the periodic ow of liquid and vapor slugs in rapid succession continues. The wall dryout may occur during the vapor passage. The wetting and rewetting phenomena have marked similarities with the nucleate boiling ebullition cycle. The dryout period becomes longer at higher qualities and/or at higher heat uxes. The nucleate boiling heat transfer suffers further as the liquid lm completely evaporates dryout during extended periods of vapor passage. 4. As the heat ux increases, the dry patches become larger and stay longer, eventually reaching the CHF condition. Closing Remarks. Although it is desirable to have a good mathematical description of the heat transfer mechanism during ow boiling in microchannels, the rapid changes of ow patterns, complex nature of the liquid lm evaporation, transient conduction, and nucleation make it quite challenging to develop a comprehensive analytical model. Further, the conduction in the substrate, parallel channel instability issues, header-channel interactions, and ow maldistribution in parallel channels pose additional challenges. The usage of the new non-dimensional groups K1 and K2 in conjunction with the Weber number and the Capillary number is expected to provide a better tool for analyzing the experimental data and developing more representative models. The scale change in hydraulic diameter introduces a whole new set of issues in microchannels that need to be understood rst from a phenomenological standpoint. Journal of Heat Transfer

Nomenclature
A Bo Bo C Ca Co cp d d P / dz Dh Eo F Fl Fr F f G g h hfg Ja K1 K2 k area, (m2 ) Bond number, Table 1 Boiling number, Table 1 a constant capillary number, Table 1 convection number, ( G / L ) 0.5((1 x )/ x ) 0.8 specic heat, J/kg K diameter, m pressure gradient, N/m3 hydraulic diameter, m tvo s number, Table 1 Eo Fluid dependent parameter in Kandlikar 1,2 correlation, 1 for water, 1.60 for R-134a 2 gD) Froude number, G2 /( L force per unit length, N/m friction factor mass ux, (kg/m2 s) gravitational acceleration, (m/s2 ) heat transfer coefcient, (W/m2 K) latent heat of vaporization, J/kg Jakob number, Table 1 new non-dimensional constant, eq. 5 new non-dimensional constant, eq. 6 thermal conductivity, W/m K FEBRUARY 2004, Vol. 126 15

L,l m Nu P q Re T T V v We w X x Z Greek

length m mass ow rate kg/s Nusselt number ( hDh /k) pressure Pa heat ux (W/m2 ) Reynolds number ( GDh / ) temperature (C) temperature difference K velocity m/s specic volume (m3 /kg) Weber number, Table 1 channel width m Martinelli parameter, Table 1 vapor mass fraction Ohnesorge number, Table 1

orientation with respect to horizontal () dynamic viscosity (Ns/m2 ) contact angle () density (kg/m3 ) surface tension N/m Subscripts CHF critical heat ux F friction CBD convection dominated region G gas or vapor g due to gravity I inertia L liquid LO all ow as liquid f g latent quantity, difference between vapor and liquid properties M due to momentum change NBD nucleate boiling dominant region ONB onset of nucleate boiling S surface tension sat saturation sub subcooled T P two-phase

References
1 Schrock, V. E., and Grossman, L. M., 1959, Forced Convection Boiling Studies, Forced Convection Vaporization ProjectIssue No. 2, Series No. 73308UCX 2182, Nov. 1, 1959, University of California, Berkeley, CA. 2 Kandlikar, S. G., and Grande, W. J., 2002, Evolution of Microchannel Flow PassagesThermohydraulic Performance and Fabrication Technology, Heat Transfer Eng., 251, pp. 317. 3 Kawaji, M., and Chung, P. M.-Y., 2003, Unique Characteristics of Adiabatic Gas-Liquid Flows in Microchannels: Diameter and Shape Effects on Flow Patterns, Void Fraction and Pressure Drop, Paper No. ICMM2003-1013, Proceedings of the First International Conference on Microchannels and Minichannels, ASME, NY, pp. 115127. 4 Ali, M. I., Sadatomi, M., and Kawaji, M., 1993, Two-Phase Flow in Narrow Channels between Two Flat Plates, Can. J. Chem. Eng., 715, pp. 449 456. 5 Shah, M. M., 1982, Chart Correlation for Saturated Boiling Heat Transfer: Equations and Further Study, ASHRAE Trans., 88, Part I, pp. 185196. 6 Stephan, K., and Abdelsalam, M., 1980, Heat Transfer Correlation for Natural Convection Boiling, Int. J. Heat Mass Transfer, 23, pp. 73 87. 7 Kutateladze, 1948, On the Transition to Film Boiling under Natural Convection, Kotloturbostroenie, No. 3, pp. 1012. 8 Kandlikar, S. G., 2001, A Theoretical Model To Predict Pool Boiling CHF Incorporating Effects Of Contact Angle And Orientation, ASME Journal of Heat Transfer, 123, pp. 10711079. 9 Steinke, M. E., and Kandlikar, S. G., 2003, Flow Boiling and Pressure Drop in Parallel Microchannels, Paper No. ICMM2003-1070, First International Conference on Microchannels and Minichannels, ASME, NY, pp. 567579. 10 Kandlikar, S. G., Steinke, M., Tian, S., and Campbell, L., 2001, High-Speed Photographic Observation of Flow Boiling of Water in Parallel MiniChannels, Paper # NHTC01-11262, ASME National Heat Transfer Conference, ASME, New York. 11 Kandlikar, S. G., 2002, Fundamental Issues Related To Flow Boiling In Min-

ichannels and Microchannels, Exp. Therm. Fluid Sci., 262 4, pp. 389 407. 12 Vandervort, C. L., Bergles, A. E., and Jensen, M. K., 1994, An Experimental Study of Critical Heat Flux in very High Heat Flux Subcooled Boiling, Int. J. Heat Mass Transfer, 37, Suppl 1, pp. 161173. rmeu bergang und Druckabfall in waagrechten Ro13 Chawla, J. M., 1967, Wa mung von verdampfenden Ka ltemitteln, VDIhren bei der Stro Forschungsheft, No. 523. 14 Yen, T.-H., Kasagi, N., and Suzuki, Y., 2002, Forced Convective Boiling Heat Transfer at Low Mass and Heat Fluxes, Proceedings of the International Symposium on Compact Heat Exchangers, Grenoble, Edizioni ETS, pp. 190. 15 Kim, J., and K. Bang, 2001, Evaporation Heat Transfer of Refrigerant R-22 in Small Hydraulic-Diameter Tubes, Proceedings of 4th International Conference on Multiphase Flow. 16 Ravigururajan, T. S., Cuta, J., McDonald, C. E., and Drost, M. K., 1996, Effects of Heat Flux on Two-Phase Flow Characteristics of Refrigerant Flows in a Micro-Channel Heat Exchanger, Proceedings of National Heat Transfer Conference, HTD-329, pp. 167178. 17 Lin, S., P. A. Kew, and K. Cornwell, 2001, Flow Boiling of Refrigerant R141B in Small Tubes, Transactions of IchemE, 79, Part A, pp. 417 424. 18 Wambsganss, M. W., France, D. M., Jendrzejczyk, J. A., and Tran, T. N., 1993, Boiling Heat Transfer in a Horizontal Small-Diameter Tube, ASME Journal of Heat Transfer, 115, pp. 963972. 19 Yan, Y., and Lin, T., 1998, Evaporation Heat Transfer and Pressure Drop of Refrigerant R-134a in a Small Pipe, Int. J. Heat Mass Transfer, 41, pp. 4183 4194. 20 Bao, Z. Y., Fletcher, D. F., Hayes, B. S., 2000, Flow Boiling Heat Transfer of Freon R11 and HCFC123 in Narrow Passages, Int. J. Heat Mass Transfer, 43, pp. 33473358. 21 Kamidis, D. E., and T. S. Ravigururajan, 1999, Single and Two-Phase Refrigerant Flow in Mini-Channels, Proceedings of 33rd National Heat Transfer Conference. 22 Hsu, Y. Y., and Graham, R. W., 1961, An Analytical and Experimental Study of the Thermal Boundary Layer and Ebullition Cycle in Nucleate Boiling, NASA TN-D-594. 23 Sato, T., and Matsumura, H., 1964, On the Conditions of Incipient Subcooled Boiling with Forced Convection, Bull. JSME, 726, pp. 392398. 24 Kandlikar, S. G., Steinke, M. E., and Balasubramanian, P., 2002, SinglePhase Flow Characteristics and Effect of Dissolved Gases on Heat Transfer Near Saturation Conditions in Microchannels, Paper No. IMECE2002-39392, International Mechanical Engineering Conference and Exposition 2002, ASME, New York. 25 Kandlikar, S. G., and Balasubramanian, P., 2003, High Speed Photographic Observation of Flow Patterns During Flow Boiling in a Single Rectangular Minichannel, Paper No. HT2003-47175, Proceedings of the ASME Summer Heat Transfer Conference. 26 Pelles, Y. P., Yarin, L. P., and Hetsroni, G., 2001, Steady and Unsteady Flow in a Heated Capillary, Int. J. Multiphase Flow, 274, pp. 577598. 27 Hetsroni, G., Segal, Z., and Mosyak, A., 2000, Nonuniform Temperature Distribution in Electronic Devices Cooled by Flow in Parallel Microchannels, Packaging of Electronic and Photonic Devices, EEP- Vol. 28, ASME, New York, pp. 19. 28 Zhang, L., Koo, J., Jiang, L., Asheghi, M., Goodson, K. E., Santiago, J. G., and Kenney, T. W., 2002, Measurements and Modeling of Two-Phase Flow in Microchannels with Nearly Constant Heat Flux Boundary Conditions, J. Microelectromech. Syst., 111, pp. 1219. 29 Kandlikar, S. G., 2001, A Photographic Insight into the Motion of LiquidVapor Interface of a Liquid Droplet on a Heated Surface, Paper #IMECE01, paper presented at the ASME IMECE 2001, ASME, New York. 30 Kandlikar, S. G., 1990, A General Correlation for Two-phase Flow Boiling Heat Transfer Coefcient Inside Horizontal and Vertical Tubes, ASME Journal of Heat Transfer, 112, pp. 219228. 31 Kandlikar, S. G., 1991, Development of a Flow Boiling Map for Subcooled and Saturated Flow Boiling of Different Fluids in Circular Tubes, ASME Journal of Heat Transfer, 113, pp. 190200. 32 Petukhov, B. S., and Popov, V. N., 1963, Theoretical Calculations of Heat Exchange in Turbulent Flow in Tubes of an Incompressible Fluid with Variable Physical Properties, High Temp., 11, pp. 69 83. 33 Glienski, V., 1976, New Equations for Heat and Mass Transfer in Turbulent Pipe and Channel Flow, Int. Chem. Eng., 16, pp. 359368. 34 Kandlikar, S. G., and Steinke, M. E., 2003, Predicting Heat Transfer During Flow Boiling in Minichannels and Microchannels, ASHRAE Trans., 109, Part 1. 35 Kandlikar, S. G., and Balasubramanian, P., 2003, Extending the Applicability of the Flow Boiling Correlation at Low Reynolds Number Flows in Microchannels, Paper ICMM2003-1075, presented at the First International Conference on Microchannels and Minichannels, ASME, New York, pp. 603 608. 36 Kandlikar, S. G., and Steinke, M. E., 2003, Flow Boiling and Pressure Drop in Parallel Flow Microchannels, First International Conference on Microchannels and Minichannels, April 24 25, 2003, Rochester, NY, USA, Paper No. ICMM 2003-1070, S. G. Kandlikar, ed., ASME, pp. 567579. 37 Jacobi, A. M., and Thome, J. R., 2002, Heat Transfer Model for Evaporation of Elongated Bubble Flows in Microchannels, ASMR Journal of Heat Transfer, 124, pp. 11311136.

16 Vol. 126, FEBRUARY 2004

Transactions of the ASME

Anda mungkin juga menyukai