international journal of heat and mass transfer - me.sc.edu flow boiling in... · enhanced flow...

10
Enhanced flow boiling in microchannels using auxiliary channels and multiple micronozzles (I): Characterizations of flow boiling heat transfer Wenming Li a , Fanghao Yang b , Tamanna Alam a , Xiaopeng Qu a , Benli Peng a , Jamil Khan a , Chen Li a,a Department of Mechanical Engineering, University of South Carolina, Columbia, SC 29208, United States b Princeton Plasma Physics Laboratory, Princeton, NJ 08540, United States article info Article history: Received 15 January 2017 Received in revised form 7 August 2017 Accepted 2 September 2017 Available online 17 October 2017 Keywords: Flow boiling Microchannels Heat transfer coefficient (HTC) Two-phase flow instability Two-phase oscillation Two-phase mixing Electronic cooling abstract Flow boiling in parallel microchannels can be dramatically enhanced through inducing self-excited and self-sustained high frequency two-phase oscillations as demonstrated in our previous studies using a two-nozzle microchannel configuration. Two-phase mixing induced by the rapid bubble collapse is shown to be the major enhancement mechanism. However, in the two-nozzle configuration microchan- nels, the mixing effect is limited to the downstream of the microchannels, meaning that only half length of the entire microchannel is functionalized as designated. In this study, a four-nozzle microchannel con- figuration is developed with an aim at extending the highly desirable mixing effect to the entire channel. Flow boiling in the four-nozzle configuration microchannels is experimentally studied with deionized water and the mass flux ranging from 120 kg/m 2 s to 600 kg/m 2 s. The onset of nucleate boiling temper- ature is considerably reduced by 14% because of more nucleation sites created by the multiple nozzles. Equally important, the improved microchannel configuration successfully extends the mixing to the entire channel as validated by the enhanced heat transfer rate and visualization study. Compared to the previous two-nozzle configuration, the overall heat transfer coefficient (HTC) is significantly improved primarily owing to the enhanced nucleate boiling. For example, the peak overall HTC of 262 kW/m 2 K is achieved at a mass flux of 150 kg/m 2 s, accounting for 83.7% enhancement. Additionally, the peak effective HTC reaches 97.6 kW/m 2 K, accounting for 67% enhancement. Ó 2017 Elsevier Ltd. All rights reserved. 1. Introduction As high power electronic devices miniaturize down to nanome- ter size, high cooling capability is demanded. Among all cooling techniques, flow boiling in microchannels is one of the most effec- tive and efficient approaches to cool high power electronic devices [1,2]. However, it is challenging to improve the boiling heat trans- fer in microchannels without compromise because of the limita- tions primarily imposed by the confined vapor bubbles [3], which usually lead to flow choking and premature dryout [4,5]. The per- sistent bubble confinement inherently retards primary and effi- cient heat transfer modes including nucleation, convection, and evaporation. Extensive studies have been conducted to enhance flow boiling in microchannels in last decade [4,6–10]. Numerous techniques have been explored to promote flow boiling perfor- mance through structure and surface modifications [5,8,11] or change of working fluids [12–14]. These include enhancing nucle- ate boiling by increasing number of nucleation sites through inte- grating micro/nano-structures [8,11,15] as well as promoting thin film evaporation by creating micro-structure on the bottom surface of channels [5,16]. However, these approaches have not shown strong effectiveness on removing or managing the bubble confine- ment. The existence of confined bubbles/vapor slugs in microchan- nels severely deteriorates the heat transfer performances. For example, confined bubbles in microchannels during boiling pro- cess can induce severe two-phase flow instabilities and hence, result in liquid flow crisis, eventually retarding highly efficient thin film evaporation [2,3]. Therefore, rapidly and passively removing these confined bubbles can be an effective method to substantially improve microchannel flow boiling performances. Technically, bubble collapse/removal can be achieved by direct interface condensation [17–20]. But a slow bubble collapse/ removal cannot meet the need of high working heat loads, where generation of vigorous vapor would result in chaotic and violent two-phase flows. Many techniques have been developed to remove or manage confined bubbles. For example, inlet orifices or restric- tors were designed to overcome the bubble flow reversal [3,10]; taper geometry was explored to enhance bubble removal by changing cross-section area [21]. However, they either induce http://dx.doi.org/10.1016/j.ijheatmasstransfer.2017.09.009 0017-9310/Ó 2017 Elsevier Ltd. All rights reserved. Corresponding author. E-mail address: [email protected] (C. Li). International Journal of Heat and Mass Transfer 116 (2018) 208–217 Contents lists available at ScienceDirect International Journal of Heat and Mass Transfer journal homepage: www.elsevier.com/locate/ijhmt

Upload: truongdieu

Post on 24-Mar-2018

218 views

Category:

Documents


1 download

TRANSCRIPT

Page 1: International Journal of Heat and Mass Transfer - me.sc.edu Flow Boiling in... · Enhanced flow boiling in microchannels using auxiliary channels and multiple micronozzles (I): Characterizations

International Journal of Heat and Mass Transfer 116 (2018) 208–217

Contents lists available at ScienceDirect

International Journal of Heat and Mass Transfer

journal homepage: www.elsevier .com/locate / i jhmt

Enhanced flow boiling in microchannels using auxiliary channels andmultiple micronozzles (I): Characterizations of flow boiling heat transfer

http://dx.doi.org/10.1016/j.ijheatmasstransfer.2017.09.0090017-9310/� 2017 Elsevier Ltd. All rights reserved.

⇑ Corresponding author.E-mail address: [email protected] (C. Li).

Wenming Li a, Fanghao Yang b, Tamanna Alam a, Xiaopeng Qu a, Benli Peng a, Jamil Khan a, Chen Li a,⇑aDepartment of Mechanical Engineering, University of South Carolina, Columbia, SC 29208, United Statesb Princeton Plasma Physics Laboratory, Princeton, NJ 08540, United States

a r t i c l e i n f o a b s t r a c t

Article history:Received 15 January 2017Received in revised form 7 August 2017Accepted 2 September 2017Available online 17 October 2017

Keywords:Flow boilingMicrochannelsHeat transfer coefficient (HTC)Two-phase flow instabilityTwo-phase oscillationTwo-phase mixingElectronic cooling

Flow boiling in parallel microchannels can be dramatically enhanced through inducing self-excited andself-sustained high frequency two-phase oscillations as demonstrated in our previous studies using atwo-nozzle microchannel configuration. Two-phase mixing induced by the rapid bubble collapse isshown to be the major enhancement mechanism. However, in the two-nozzle configuration microchan-nels, the mixing effect is limited to the downstream of the microchannels, meaning that only half lengthof the entire microchannel is functionalized as designated. In this study, a four-nozzle microchannel con-figuration is developed with an aim at extending the highly desirable mixing effect to the entire channel.Flow boiling in the four-nozzle configuration microchannels is experimentally studied with deionizedwater and the mass flux ranging from 120 kg/m2 s to 600 kg/m2 s. The onset of nucleate boiling temper-ature is considerably reduced by �14% because of more nucleation sites created by the multiple nozzles.Equally important, the improved microchannel configuration successfully extends the mixing to theentire channel as validated by the enhanced heat transfer rate and visualization study. Compared tothe previous two-nozzle configuration, the overall heat transfer coefficient (HTC) is significantlyimproved primarily owing to the enhanced nucleate boiling. For example, the peak overall HTC of262 kW/m2 K is achieved at a mass flux of 150 kg/m2 s, accounting for �83.7% enhancement.Additionally, the peak effective HTC reaches 97.6 kW/m2 K, accounting for �67% enhancement.

� 2017 Elsevier Ltd. All rights reserved.

1. Introduction

As high power electronic devices miniaturize down to nanome-ter size, high cooling capability is demanded. Among all coolingtechniques, flow boiling in microchannels is one of the most effec-tive and efficient approaches to cool high power electronic devices[1,2]. However, it is challenging to improve the boiling heat trans-fer in microchannels without compromise because of the limita-tions primarily imposed by the confined vapor bubbles [3], whichusually lead to flow choking and premature dryout [4,5]. The per-sistent bubble confinement inherently retards primary and effi-cient heat transfer modes including nucleation, convection, andevaporation. Extensive studies have been conducted to enhanceflow boiling in microchannels in last decade [4,6–10]. Numeroustechniques have been explored to promote flow boiling perfor-mance through structure and surface modifications [5,8,11] orchange of working fluids [12–14]. These include enhancing nucle-ate boiling by increasing number of nucleation sites through inte-

grating micro/nano-structures [8,11,15] as well as promoting thinfilm evaporation by creating micro-structure on the bottom surfaceof channels [5,16]. However, these approaches have not shownstrong effectiveness on removing or managing the bubble confine-ment. The existence of confined bubbles/vapor slugs in microchan-nels severely deteriorates the heat transfer performances. Forexample, confined bubbles in microchannels during boiling pro-cess can induce severe two-phase flow instabilities and hence,result in liquid flow crisis, eventually retarding highly efficient thinfilm evaporation [2,3]. Therefore, rapidly and passively removingthese confined bubbles can be an effective method to substantiallyimprove microchannel flow boiling performances.

Technically, bubble collapse/removal can be achieved by directinterface condensation [17–20]. But a slow bubble collapse/removal cannot meet the need of high working heat loads, wheregeneration of vigorous vapor would result in chaotic and violenttwo-phase flows. Many techniques have been developed to removeor manage confined bubbles. For example, inlet orifices or restric-tors were designed to overcome the bubble flow reversal [3,10];taper geometry was explored to enhance bubble removal bychanging cross-section area [21]. However, they either induce

Page 2: International Journal of Heat and Mass Transfer - me.sc.edu Flow Boiling in... · Enhanced flow boiling in microchannels using auxiliary channels and multiple micronozzles (I): Characterizations

Nomenclature

Ab base area, m2

Af fin surface, m2

As surface area, m2

At total heat transfer area, m2

Cp heat capacity at constant pressure, J/kg�KG mass flux, kg/m2�s�h overall heat transfer coefficient, W/m2�Khfg latent heat of vaporization, kJ/kgH channel height, mI electrical current, Ak thermal conductivity, W/m�KL length, mm parameter for fin efficiency_m mass flow rate, kg/sN number of microchannelsp pressure, kPaP power, Wq00 heat flux, W/cm2

R electrical resistance, XK slope of linear function, X/Kt thickness between heater and microchannel base, mT temperature, �C�T overall temperature, �C

V electrical voltage, VW microchannel width, m

Greek symbolsg efficiencyv vapor quality

Subscriptsa ambientb baseCHF critical heat fluxe exiteff effectivef finHTC heat transfer coefficienti inletl linears surfacesat saturatedsp single-phasetp two-phase

W. Li et al. / International Journal of Heat and Mass Transfer 116 (2018) 208–217 209

additional pressure drop or require high flow rate. Most recently,without using external controllers and forces, a microfluidic tran-sistor was devised to timely collapsing confined bubbles and gen-erate highly desirable two-phase mixing by inducing highfrequency self-sustained two-phase oscillations (TPOs) in our pre-vious studies [4,22]. However, with one pair of nozzles located inthe middle of each main channel, persistent confined vapor slugsin the upstream led to a limited mixing range. Additionally, lessnucleation sites are unfavorable to enhance flow boiling heattransfer rate.

To enhance nucleate boiling and extend the mixing range, inthis study, an improved microchannel configuration is proposed.For better compactness, an auxiliary channel is shared by twoneighboring main channels. The new configuration is developedbased on the experimental and theoretical study of two-nozzlemicrochannel configuration [4,9]. On the improved configuration,nucleate boiling can be further enhanced and the range of mixingcan be extended to the entire channel.

Fig. 1. The design and major dimensions of the present four-nozzle microchannelconfiguration. The two-nozzle microchannel configuration [4,22] is included forcomparisons. (a) The improved four-nozzle microchannel configuration. (b) SEMimage of the micronozzle distribution with a pitch of 2 mm. (c) SEM image of themicronozzles. (d) The two-nozzle configuration developed in our previous study [4].(e) A close look of the two micronozzle locations [4]. (f–g) SEM images of a pair ofmicronozzles and a close-look of the micronozzle, respectively [4].

2. Experimental apparatus and procedures

2.1. Design of the device architecture

As shown in Fig. 1(a–c), a four-nozzle microchannel configura-tion is developed with an aim at extending the mixing to the entirechannel length, which can potentially double the mixing rangecompared to our previous two-nozzle microchannel configuration(Fig. 1(d–g)) [4]. Herein, an array of five parallel microchannels(W = 200 mm, H = 250 mm, L = 10 mm), where each main channelis connected to two auxiliary channels (H = 250 lm, W = 60 lm,L = 8 mm), is developed to experimentally characterize flow boilingheat transfer. Each 60 lm wide auxiliary channel has four 20 lmwide nozzles connected to the main channel. An inlet restrictor(H = 250 lm; W = 20 lm; L = 400 lm) is integrated in each mainchannel to trap elongated bubbles. A resistor, which serves as botha micro heater to generate heat flux and a thermistor to measurethe wall temperature, is deposited onto the back side of the siliconchip. The heating area (10 mm � 2 mm) is identical to the total

base area of microchannel arrays. A comparison of main featuresbetween the present design and our previous microchannel config-uration is summarized in Table 1. More descriptions of device andexperimental details can be found in our previous studies [4,22].

Page 3: International Journal of Heat and Mass Transfer - me.sc.edu Flow Boiling in... · Enhanced flow boiling in microchannels using auxiliary channels and multiple micronozzles (I): Characterizations

210 W. Li et al. / International Journal of Heat and Mass Transfer 116 (2018) 208–217

2.2. Microfabrication of the device

The microdevice was made from a silicon wafer bonded to aPyrex wafer by a standard microfabrication process detailed inour previous study [4]. This process started with a double-side-polished n-type h1 0 0i silicon wafer. First, before fabricating amicroheater, a 1 ± 0.01 lm thick thermal oxide layer was grownon both sides of the silicon wafer. The silicon oxide film provideselectrical insulation for the micro heater and acts as a mask fordeep reactive ion etching (DRIE) in the subsequent microfabrica-tion steps. Next, a thin film heater was fabricated through a lift-off process on the backside of the wafer. In the lift-off process, apattern mold of a thin film heater was prepared with a negativephotoresist by photolithography first. Then, a 1.5 ± 0.05 lm thicklayer of aluminum was deposited followed by a 63 Å thick of tita-nium layer by DC sputtering. Once the thin films were successfullydeposited, a thin film heater was formed by a lift-off process. A1 ± 0.05 lm thick plasma-enhanced chemical vapor deposition(PECVD) oxide layer was then deposited to protect the thin filmheater in the subsequent fabrication processes.

After the heater was formed on the backside, a patterned maskof microchannels on the top side of the wafer were etched in thesilicon oxide through photolithography and reactive ion etching(RIE). The area under the oxide mask was protected and theremaining areas were etched out to create 250 ± 3 lm deeptrenches by DRIE.

A Pyrex glass wafer was anodically bonded to the silicon sub-strate to seal the device. The transparent glass cover also servesas a visualization window. RIE was used to remove oxide coatingson the backside to expose the contact pads after anodic bonding.The individual microchannel test chips (length 30 ± 0.005 mm;width 10 ± 0.005 mm; thickness 1 ± 0.005 mm) were cut from thewafer by a dice saw.

2.3. Experimental setup

The experimental setup and test procedures detailed in our pre-vious study [4] were used and applied to the present study. Majorcomponents of the experimental setup include an optical imagingsystem, a data acquisition unit, and an open coolant loop. A pres-surized water tank was used to supply deionized (DI) water, whichwas degassed prior to tests and pumped by compressed nitrogen(N2). Flow rates were measured by an Omega flow meter (FLV-4604A) with a 1 ml/min resolution. Electrical power was suppliedby a high precision digital programmable power supply (BK-PRECISION XLN10014). The voltage on the micro heater was mea-sured by an Agilent digital multimeter (34972A). Two Omega Ktype thermocouples were used to measure the inlet and outletfluid temperatures. Flow rates, pressures, inlet and outlet temper-atures, and voltage and current were recorded by a customizeddata acquisition system developed from NI LabVIEW�. A visualiza-tion system comprised of a high-speed camera (Phantom V 7.3)with 256 � 256 pixels at approximate 40,000 frames per secondand an Olympus microscope (BX-51) with 400� amplificationswas used to study the bubble dynamics and two-phase flowstructures.

Table 1Comparisons of primary features between two microchannel configurations.

Geometryconfiguration

Nozzle number in eachchannel

Length of auxiliarychannel

Alignmnozzles

The presentconfiguration

4 8 mm At a pit2 mm

Two-nozzleconfiguration

2 5 mm At 5 mm

2.4. Experimental procedures

The microheater served as a thermistor was calibrated in anisothermal oven prior to tests. The relationship between micro-heater temperature and its electric resistance was generated usinga linear curve fitting. The confidence of the correlation coefficientwas estimated to be higher than 0.9999. The heat loss as a functionof temperature difference between micro device and ambient tem-perature was experimentally evaluated. Without fluid flows, thedevice was gradually heated up into steady state. The device tem-perature change is proportional to the heat input. The curveobtained by linear fitting was used to estimate heat loss with ahigh accuracy [4].

After assembling the microchannel device on the test package,the flow rate was kept constant at a set value ranging from120 kg/m2 s to 600 kg/m2 s. Uniform heat fluxes at a constant stepof around 3–20W/cm2 were applied by a digital power supplythrough the microheater until approaching CHF conditions. Foreach data point, the data acquisition system recorded 90 sets ofsteady state experimental data including voltages, currents, localpressures and temperatures at inlet and outlet at 4 min intervals.

3. Data reduction

3.1. Flow boiling heat transfer

In this study, the data of voltage, current and pressure drop areobtained to deduce wall temperatures and HTCs. The electricalinput power (P) and resistance of the heater (R) is calculated as,respectively,

P ¼ V � I ð1Þand

R ¼ V=I ð2ÞThen heat loss (Ploss) between the environment and the testing

chip is deducted from P to calculate the effective power:

Peff ¼ P � Ploss ð3ÞThe overall temperature of the thin film heater is calculated as,

�Theater ¼ KðR� R0Þ þ T0 ð4Þwhere R0 is the resistance of the microheater at room temperatureand K is the slope of the heater electrical resistance-temperaturecalibration curve. The surface temperature (Twall) at the base areaof the microchannels is then calculated using the overall surfacetemperature of the thin film heater as

�Twall ¼ �Theater �q00eff t

ksð5Þ

where q00eff ¼ Peff =Ab.

To accurately derive the two-phase heat transfer rate, themicrochannels length (L) is divided into two sections: single-phase region (Lsp) and two-phase region (Ltp) [4]. The length ofthe two sections varies significantly with vapor quality as observed

ent of Number of mainchannel

Opening ofnozzle

How nozzles connected

ch of 5 25 lm 4 nozzles share one auxiliarychannel

4 20 lm 1 nozzle in each auxiliarychannel

Page 4: International Journal of Heat and Mass Transfer - me.sc.edu Flow Boiling in... · Enhanced flow boiling in microchannels using auxiliary channels and multiple micronozzles (I): Characterizations

Fig. 2. Overall HTC based on heating area versus (a) effective heat flux, and (b) exitvapor quality with mass flux ranging from 120 kg/m2 s to 600 kg/m2 s.

W. Li et al. / International Journal of Heat and Mass Transfer 116 (2018) 208–217 211

by flow visualization. The two-phase HTC (�htp) considering fin effi-ciency is used to evaluate flow boiling heat transfer performance inthe two-phase region (Ltp) by excluding the weight of the single-phase heat transfer from the overall temperature. The single-phase effective HTC considering fin efficiency is calculated from,

�hsp ¼ Peff

½At � NAf ð1� gf Þ�½�Tsp � ðTi þ TeÞ=2�ð6Þ

where the fin efficiency (gf ) is estimated from

gf ¼ tanhðmHÞ=mHand m ¼ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi2�hspðLþWf Þ=ksWf L

q. The overall tem-

perature in single-phase region is the mean value of the inlet andexit surface temperatures in this region.

�Tsp ¼ Tin;sp þ Texit;sp

2ð7Þ

In the above equation, the inlet and the exit temperatures are esti-mated as,

Tin;sp ¼ Ti þ Peff

�hspAsand Texit;sp ¼ Tsat þ Peff

�hspAsð8Þ

The overall temperature of the two-phase heat transfer region(�Ttp) is obtained by a weighted overall method [10] in terms ofthe single-phase and overall wall temperatures:

�Ttp ¼�TwallL� �TspLsp

Ltpð9Þ

These length values (Lsp and Ltp) are measured through visual-ization. Considering fin effects on a single microchannel, the effec-tive two-phase HTC considering fin efficiency is calculated as,

�htp ¼ Peff

ðPðWLþ 2HLgf ÞÞð�Ttp � TsatÞð10Þ

Because the thermal conductivity of Pyrex glass is approxi-mately 1% of silicon, the interface between the microchannel wallsand the cover glass is assumed to be thermally insulated in the finapproximation. Then, �htpwas iteratively obtained from Eq. (1)through Eq. (10). Additionally, the exit vapor quality was calcu-lated with mass flow rate and input power according to:

v ¼ Peff � Qsensible

_mhfgð11Þ

Then, the contribution of latent heat during phase-change heattransfer was derived as,

Qlatent ¼ Peff � Qsensible ð12Þ

where Peff ¼ P � Qloss, Qsensible is the sensible heat resulting from thesubcooled liquid. It was derived as,

Qsensible ¼ GAcCpðTsat � TiÞ ð13Þ

In above equations, t and ks are the substrate thickness, the thermalconductivity of silicon, respectively. Major physical properties of DI-water are given in atmosphere. Tsat is a function of working pressure(p) in the middle of a microchannel, which is estimated as,

p ¼ pi �Dp2

ð14Þ

where pi is the inlet pressure and Dp is the pressure drop.In addition, to evaluate the overall heat transfer performance of

the device, the overall HTC based on the heating area is estimatedas

�h ¼ Peff

Abð�Twall � TsatÞð15Þ

3.2. Uncertainty analysis

The uncertainty of mass flux, electrical voltage, electrical cur-rent, electrical power, ambient temperature and average wall tem-perature are ±1%, ±0.5%, ±0.5%, ±0.7%, ±0.5 �C and ±0.86%,respectively. Uncertainty propagations are calculated using meth-ods developed by Kline and McClintock [23]. For results of overallHTC, the uncertainty is ±6%.

4. Results and discussion

4.1. Flow boiling curves

4.1.1. Flow boiling curves based on heating areaFig. 2 shows the overall HTCs based on the heating area as a

function of heat flux at mass flux ranging from 120 kg/m2 s to600 kg/m2 s. The overall HTCs increase with increasing mass flux.Fig. 2 indicates that the slope of each HTC curve initially declinessharply with the increase in effective heat flux and exit vapor qual-ity, and then becomes flat. During the early boiling stage, nucleateboiling is dominant associated with a low vapor quality. The dis-persed bubbles in the channel may promote the convection heattransfer by increasing the liquid superficial velocity. Hence, theoverall HTC shows the highest value after onset of nucleate boiling(ONB). With further increase in heat flux and thus, the vaporquality, the role of nucleate boiling in determining heat transfergradually decreases while convective boiling and thin film

Page 5: International Journal of Heat and Mass Transfer - me.sc.edu Flow Boiling in... · Enhanced flow boiling in microchannels using auxiliary channels and multiple micronozzles (I): Characterizations

Fig. 4. Significant mixing is not observed in the downstream at a low heat flux of117 W/cm2 and a mass flux of 150 kg/m2 s (Scale bar is 200 lm).

Fig. 5. Stronger mixing and higher frequency two-phase oscillations were observedin upstream at a higher heat flux of 298 W/cm2 and a mass flux of 325 kg/m2 s(Scale bar is 200 lm).

212 W. Li et al. / International Journal of Heat and Mass Transfer 116 (2018) 208–217

evaporation prevail. Fig. 2(b) shows two distinct bands of HTCcurves as indicated by exit vapor quality among different flowrates. At low mass fluxes between 120 kg/m2 s and 180 kg/m2 s,the HTCs dramatically decrease among exit vapor quality from0.25 to 0.6. At a higher mass flux ranging from 325 kg/m2 s to600 kg/m2 s, the HTCs gradually decrease among exit vapor qualityfrom 0.1 to 0.35. Previous studies demonstrate that the HTC des-cends with the increase in exit vapor quality due to the increasingheat flux [4,24,25].

A visualization study would help explain different trends offlow boiling curves between the low and high mass fluxes. Fig. 3shows that TPOs present in upstream at a low mass flux of150 kg/m2 s. The confined bubble shrinks due to interfacial con-densation. The effect of two-phase mixing induced by jetting flowsis limited in enhancing HTC at low heat flux and mass flux as indi-cated in Fig. 3 and Fig. 4. In another word, nucleate boiling is dom-inant. As mass flux increases, for example, at 325 kg/m2 s asillustrated in Fig. 5, strong mixing induced by jetting flow and TPOsare well established, indicating the dominance of convective boil-ing. Compared to upstream, however, Fig. 4 and Fig. 6 show thatthe mixing is not as obvious and frequent as observed in the down-stream. Partial dryout or liquid detaching from walls starts todevelop there. The force analysis of flow boiling in microchannelsby Alam et al. [26] has demonstrated that evaporation momentumforce increases with increasing heat flux and then dominates overshear force at high heat flux ranges, which makes it difficult tospread liquid on walls and hence, leading to liquid detaching asillustrated in Figs. 4 and 6. This could explain that a reduced HTCwith the increasing heat flux is a result of a reduced effective heattransfer areas.

4.1.2. Flow boiling curves considering the fin effectCompared to overall HTCs, effective HTCs considering the fin

effect could provide more insights into the heat transfer mecha-nisms. Figs. 7 and 8 shows the effective HTCs versus effective heatflux and vapor quality at low and high mass fluxes, respectively.Effective HTCs can better evaluate the heat transfer mechanismssince it considers all effective heat transfer areas. Fig. 7 shows thata peak value of effective HTCs exists at low mass fluxes of 120 kg/m2 s, 150 kg/m2 s and 180 kg/m2 s. The effective HTCs initiallyincrease with the increase in effective heat flux until the establish-ment of the fully developed boiling; and then gradually decreasewith increasing heat flux. With an increase in effective heat flux,dryout was observed in the downstream as illustrated in Fig. 4 ata mass flux of 150 kg/m2 s and a heat flux of 117W/cm2, resultingin descending effective HTC curves. In the effective HTC curves ver-sus exit vapor quality, the effective HTCs reach their maximum val-ues at exit vapor qualities between 0.3 and 0.45 when boiling isfully established in the entire channel, as shown in Fig. 7. For

Fig. 3. Two-phase oscillations were observed in the upstream at a heat flux of117 W/cm2 and a mass flux of 150 kg/m2 s in the present four-nozzle configurationmicrochannels. The jetting flow from nozzle is not strong and effective in collapsingconfined bubbles at a low heat flux (Scale bar is 200 lm).

Fig. 6. Compared to upstream, the mixing is not obvious in the downstream at aheat flux of 298 W/cm2 and a mass flux of 325 kg/m2 s (Scale bar is 200 lm).

example, at the low mass flux of 150 kg/m2 s, nucleate boiling pre-vails; while the confined bubble shrinks in upstream owing tointerfacial condensation. Although TPOs were observed, the inten-sity in terms of frequency and magnitude is low. The effective HTCssharply drop as the exit vapor quality higher than 0.4 because theincrease of exit vapor quality leads to even weaker mixing (as indi-cated in Figs. 3 and 4) and hence, the local dryout starts to developfrom outlet region. For example, Fig. 6 shows that mixing is notobvious and intensive in the downstream at a mass flux of325 kg/m2 s and a heat flux of 298W/cm2.

Page 6: International Journal of Heat and Mass Transfer - me.sc.edu Flow Boiling in... · Enhanced flow boiling in microchannels using auxiliary channels and multiple micronozzles (I): Characterizations

Fig. 7. Effective HTCs against (a) effective heat flux and (b) exit vapor quality areplotted at low mass fluxes.

Fig. 8. Effective HTCs against (a) effective heat flux and (b) exit vapor quality areplotted at high mass fluxes.

W. Li et al. / International Journal of Heat and Mass Transfer 116 (2018) 208–217 213

Fig. 8 demonstrates that the effective HTCs initially increasewhen nucleate boiling is dominant and then gradually decline afterthe full establishment of mixing and TPOs for higher mass fluxes.Specifically, Fig. 8 shows that the effective HTC keeps a relativehigh value as the boiling process at a mass flux of 600 kg/m2 s asso-ciated with exit vapor quality lower than 0.4. The maximum valueis �115 kW/m2 K during fully developed flow boiling owing tostrong mixing.

4.2. Enhanced flow boiling

Fig. 9 shows that the overall HTCs based on the heating area aresignificantly enhanced compared to our previous two-nozzle con-figuration [4]. An enhancement up to 83.7% is achieved at a lowmass flux of 150 kg/m2 s (Fig. 9(a)). With an increase in massfluxes, the enhancements of overall HTCs decrease to �14.5% at amoderate mass flux of 325 kg/m2 s (Fig. 9(b)) and �30.5% at a mod-erate mass flux of 430 kg/m2 s (Fig. 9(c)). The larger enhancementof overall HTC in low mass fluxes may be attributed to theenhanced nucleate boiling primarily owing to increased nucleationsite density compared to that in the two-nozzle configuration [15].

4.2.1. Enhanced nucleate boilingTo fairly compare the boiling performance in these two

microchannel configurations, Fig. 10 compares effective HTCs formass flux ranging from 150 kg/m2 s to 430 kg/m2 s. Thoughenhancement varies with mass fluxes, the effective HTC has been

substantially enhanced under all three compared mass fluxeswhen the superheat is less than �10 K, at which nucleate boilingis believed to prevail. Specifically, at low mass fluxes, such as150 kg/m2 s, an enhancement of �67% is achieved (Fig. 10(a))and then the enhancement drops to �31% at the peak value. Assuperheat increasing, the enhancement is further reduced. Thetrend of the two boiling curves is similar. The effective HTC firstlyincreases and then decreases with an increase in effective heat flux.The peak effective HTC is achieved at the onset of fully developedboiling (OFB). However, the enhancement decreases to �54% as themass flux increases to 325 kg/m2 s in nucleate boiling region, asshown in Fig. 10(b). The trend of boiling curves between the pre-sent study and the two-nozzle configuration is quite different asmass flux reaches 325 kg/m2 s or higher. In this study, the peakeffective HTCs occurred after ONB and then effective HTCs gradu-ally decline with an increase in heat flux for mass fluxes over325 kg/m2 s (as shown in Fig. 10(b, c)). In contrast, the peak effec-tive HTC for the two-nozzle configuration happens when bubblesreach the inlet for all the mass fluxes [4]. The superheat is �15 K(in Fig. 10(b)) at a mass flux of 325 kg/m2 s when boiling is fullyestablished. Our previous study [4] has demonstrated that thegradually decline of effective HTCs was linked to the formation ofpersistent vapor slugs in the section between the cross-junctionand the inlet. Finally, the results in Fig. 10 show that the effectiveHTC is significantly enhanced compared to the two-nozzle config-uration, including significant enhancement in the whole boilingprocess at a mass flux of 150 kg/m2 s as well as in the nucleate

Page 7: International Journal of Heat and Mass Transfer - me.sc.edu Flow Boiling in... · Enhanced flow boiling in microchannels using auxiliary channels and multiple micronozzles (I): Characterizations

Fig. 9. Overall HTC based on heating area in the present study is compared to thatin the two-nozzle microchannel configuration. Fig. 10. The effective HTCs are compared to these in two-nozzle microchannel

configuration.

214 W. Li et al. / International Journal of Heat and Mass Transfer 116 (2018) 208–217

boiling region at mass flux of 325 kg/m2 s and 430 kg/m2 s. At ahigher mass flux of 430 kg/m2 s, the effective HTC values of thepresent study in the fully developed boiling region are lower thanthat of two-nozzle configuration, as indicated in Fig. 10(c).

As aforementioned, effective HTCs are significantly enhancedduring nucleate boiling region compared to the two-nozzle config-uration. Fig. 11 indicates that significant reduction of ONB isobserved owing to the enhanced nucleation because of an increaseof nucleation site density. The overall wall superheat at ONB isapproximately 15 �C lower compared to that of the two-nozzleconfiguration. Fig. 11 also shows that the ONB gradually increases

with the increase of mass flux, requiring higher average wall tem-perature to trigger nucleate boiling.

4.2.2. Enhanced mixingDue to the small hydraulic diameter, laminar flow is dominant

during flow boiling in conventional microchannels in most work-ing conditions [3,5,15,16]. Improving mixing can be an effectiveway to enhance flow boiling heat transfer in microchannels. Mix-ing is highly efficient, but hard to be passively achieved during flowboiling in microchannels. Active methods such as impingementjets [27–29] can enhance flow boiling by promoting fluid mixing.

Page 8: International Journal of Heat and Mass Transfer - me.sc.edu Flow Boiling in... · Enhanced flow boiling in microchannels using auxiliary channels and multiple micronozzles (I): Characterizations

Fig. 11. Significant reduction of ONB in the present configuration compared to thetwo-nozzle one.

Fig. 13. Persistent vapor slugs near the inlet section of each main channel after OFBwere observed at a mass flux of 430 kg/m2s and a heat flux of 399 W/cm2 in thetwo-nozzle configuration microchannels.

W. Li et al. / International Journal of Heat and Mass Transfer 116 (2018) 208–217 215

Our previous study has demonstrated that embedded staggeredherringbone mixers can significantly improve flow boiling perfor-mance through enhancing mixing [30]. Recently, mixing was alsoachieved using the self-sustained high frequency TPOs inducedby rapid thermal bubble growth-collapse [4]. However, the mixingis only limited to the central section (Fig. 12(c)) and there are noTPOs in the upstream (Fig. 13). In this study, the extendedintense-mixing (as indicated in Fig. 12(a, b)) due to high frequencyjetting flows could be another responsible factor for the enhancedeffective HTCs of the present configuration, as shown inFigs. 7 and 8.

As depicted in Fig. 14, the frequency of TPOs is at a magnitude of100 Hz in both configurations. Due to four nozzles shared one aux-iliary channel, the average frequency of TPOs in the present studyis about 2 times lower than that in the two-nozzle configuration atthe same working conditions, in which one auxiliary channel isconnected to the main channel by one nozzle. In the present study,the order of magnitude of TPOs is approximately 100 Hz and the

Fig. 12. (a, b) Extended mixing in the entire channel is achieved in the four-nozzlemicrochannel configuration at a mass flux of 600 kg/m2 s and a heat flux of 439 W/cm2. (c) Mixing was observed at the central section of channel in two-nozzleconfiguration microchannel at 430 kg/m2 s and a heat flux of 106 W/cm2 (Number1–4 represents the four locations of micronozzles from inlet to outlet. Scale bars are200 lm).

highest frequency as measured is around 571 Hz as illustrated inFig. 14.

Although the frequency of TPOs is much lower in this studycompared to that of the two-nozzle configuration in similar work-ing conditions (i.e., effective heat fluxes) as shown in Fig. 14,intense mixing induced by high frequency jetting flows is success-fully extended to the entire channel by integrating evenly-distributed four nozzles in each main channel as indicated inFig. 12(a and b). Fig. 15 shows the high frequency of jetting flowsfrom one of the four micronozzles. Thereby, the heat transfer rateof the present study is comparable to that of two-nozzle configura-tion after OFB for mass flux of 150 kg/m2 s and 325 kg/m2 s; whilethe effective HTCs are lower at a mass flux of 430 kg/m2 s in theregion of OFB. In summary, our studies show that at low massfluxes such as 150 kg/m2 s, the primary enhancement mechanismis the augmented nucleate site density, particularly with the super-heat less than 10 K. However, in high mass flux conditions,although the mixing has been successfully extended to the entirechannel, the HTC enhancement is shown to be compromised dueto the reduced overall two-phase oscillation frequency as well asthe mild mixing (or even liquid detachment) as observed in thedownstream in the present configuration.

0 200 400 600 8000

200

400

600

800Two-nozzle

150 kg/m2s 241 kg/m2s 330 kg/m2s

TPO

freq

uenc

y (H

z)

Effective heat flux (W/cm2)

Present study 150 kg/m2s 430 kg/m2s 600 kg/m2s

Fig. 14. A comparison of two-phase oscillation (TPO) frequency between the twoconfigurations. The frequency is calculated based on the time interval of the bubble-growth-collapse cycle.

Page 9: International Journal of Heat and Mass Transfer - me.sc.edu Flow Boiling in... · Enhanced flow boiling in microchannels using auxiliary channels and multiple micronozzles (I): Characterizations

0 200 400 600 8000

100

200

300

400

Freq

uenc

y (H

z)

Effective heat flux (W/cm2)

Present study 150 kg/m2 s 430 kg/m2 s 600 kg/m2 s

Fig. 15. Frequency of jetting flows versus effective heat flux. The frequency iscalculated based on the time interval of sequent jetting flows from one nozzle.

100 200 300 400 5000.0

0.4

0.8

1.2

1.6

2.0

Stan

dard

dev

iatio

n of

Tw

all

Effective heat flux (W/cm2)

Two-nozzle Present study

G=380 kg/m2s

Fig. 16. Standard deviation of wall temperature fluctuation is compared betweenthe present configuration and the two-nozzle configuration at a mass flux of 380 kg/m2 s.

216 W. Li et al. / International Journal of Heat and Mass Transfer 116 (2018) 208–217

4.2.3. Enhanced flow boiling stabilityThe enhanced removal of confined bubbles of flow boiling in

microchannels is critical in reducing two-phase flow instabilitiesin terms of both wall temperature and pressure drop fluctuations.Numerous techniques have been explored to remove or at leastmanage confined bubbles, for example, using taper geometry[21]. In our previous study, the two auxiliary microchannels hasbeen designed to connect with a main microchannel throughtwo-microscale nozzles. This configuration has been demonstratedin effectively stabilizing flow boiling in terms of both wall temper-ature and pressure drop fluctuations by inducing high frequencyTPOs compared to microchannels with IRs [3]. The removal of con-fined bubbles can also lead to significantly enhanced heat transferrate and CHF. In the present study, the extended range of intensemixing to the entire channel can further manage two-phase flowinstabilities.

Specifically, the wall temperature fluctuation in the presentdesign is substantially reduced compared to that of the two-nozzle configuration. Standard deviation of wall temperaturebetween the present study and the two-nozzle microchannel con-figuration at a mass flux of 380 kg/m2 s was compared in Fig. 16.As shown in Fig. 16, the difference of standard deviation of walltemperature varies with increasing effective heat flux. At low heatflux, the standard deviation of wall temperature of the two-nozzle configuration is �150% larger than that of the presentstudy since confined bubbles still persist in the downstreamchannel. In contrast, the increased number of micronozzles couldgenerate jetting flow near outlet section to facilitate the removalof confined bubbles. Consequently, the wall temperature fluctua-tion is reduced owing to enhanced mixing range. As the increasein effective heat flux, the standard deviations of wall temperatureare dramatically reduced on both configurations owing to intensi-fied mixing (as shown in Fig. 12), which is induced by high fre-quency TPOs. The frequency of TPOs of the two-nozzleconfiguration is much higher, as shown in Fig. 14. But theincreased number of jetting flow can still increase the mixingeffect in two aspects. First, jetting flows from four evenly-distributed micronozzles can enhance bubble removal, leadingto enhanced oscillations in main channels. Second, jetting flowsthemselves would induce mixing around the micro-nozzle area.As the further increase in effective heat flux, confined slugs havebeen better managed in the present configuration and hence,leading to smaller deviation in Fig. 16.

5. Conclusions

In our previous study, the two-nozzle microchannel configura-tion [4] has demonstrated the effectiveness in enhancing flow boil-ing owing to generation of intense mixing induced by highfrequency TPOs and jetting flows. However, mixing cannot be gen-erated in the entire channel length. In this study, we show thatthrough an improved microchannel configuration, such an enhan-cer can be extended to the entire channel. In the part (I) of thisstudy, heat transfer performance is further enhanced due toextended intense mixing induced by four evenly-distributed noz-zles. Meanwhile, nucleation is substantially enhanced because ofincreased nucleation sites. The main conclusions are summarizedbelow:

(a) Mixing is extended to the entire microchannel length.(b) An enhancement of overall HTCs up to �83.7% is achieved at

a mass flux of 150 kg/m2 s. The enhancement of overall HTCsdecreases to �14.5% at a moderate mass flux of 325 kg/m2 s.To fairly compare the enhanced mechanism, effective HTCsconsidering fin efficiency are examined. As shown inFig. 10, an enhancement of �67% is achieved at a mass fluxof 430 kg/m2 s due to the enhanced nucleate boiling, whichis consistent with the 14% lower ONB over the two-nozzleconfiguration.

(c) Effective HTCs curves are converging during fully developedboiling. Meanwhile, the effectiveness of TPOs decreases withthe increase in exit vapor quality.

(d) Flow boiling stability in terms of wall temperature is notice-ably enhanced in this present configuration.

Acknowledgments

This work was supported by the U.S. Department of Defense,Office of Naval Research under the Grant N000141210724 andN000141612307 (Program Officer Dr. Mark Spector). Devices werefabricated at Institute of Electronics and Nanotechnology (IEN) inGeorgia Tech, which is supported by the National Science Founda-tion under the Grant ECS-0335765. SEM images were taken at USCMicroscopy Center.

Page 10: International Journal of Heat and Mass Transfer - me.sc.edu Flow Boiling in... · Enhanced flow boiling in microchannels using auxiliary channels and multiple micronozzles (I): Characterizations

W. Li et al. / International Journal of Heat and Mass Transfer 116 (2018) 208–217 217

Conflict of interest

The authors declare that there are no conflicts of interest.

References

[1] S.G. Kandlikar, History, advances, and challenges in liquid flow and flowboiling heat transfer in microchannels: a critical review, J. Heat Transfer-Trans.Asme 134 (3) (2012) 034001.

[2] A.E.B.J.A. Boure, L.S. Tong, Review of two-phase flow instability, Nucl. Eng. Des.25 (2) (1973) 165–192.

[3] A. Kos�ar, C.-J. Kuo, Y. Peles, Suppression of boiling flow oscillations in parallelmicrochannels by inlet restrictors, J. Heat Transfer 128 (3) (2006) 251–260.

[4] F. Yang, X. Dai, C.-J. Kuo, Y. Peles, J. Khan, C. Li, Enhanced flow boiling inmicrochannels by self-sustained high frequency two-phase oscillations, Int. J.Heat Mass Transf. 58 (1–2) (2013) 402–412.

[5] Y. Zhu, D.S. Antao, K.-H. Chu, S. Chen, T.J. Hendricks, T. Zhang, E.N. Wang,Surface structure enhanced microchannel flow boiling, J. Heat Transfer 138 (9)(2016) 091501.

[6] C.T. Lu, C. Pan, A highly stable microchannel heat sink for convective boiling, J.Micromechan. Microeng. 19 (5) (2009) 055013.

[7] B. Schneider, A. Kos�ar, C.-J. Kuo, C. Mishra, G.S. Cole, R.P. Scaringe, Y. Peles,Cavitation enhanced heat transfer in microchannels, J. Heat Transfer 128 (12)(2006) 1293.

[8] C. Li, Z. Wang, P.I. Wang, Y. Peles, N. Koratkar, G.P. Peterson, Nanostructuredcopper interfaces for enhanced boiling, Small 4 (8) (2008) 1084–1088.

[9] W. Li, F. Yang, T. Alam, J. Khan, C. Li, Experimental and theoretical studies ofcritical heat flux of flow boiling in microchannels with microbubble-excitedhigh-frequency two-phase oscillations, Int. J. Heat Mass Transf. 88 (2015) 368–378.

[10] A. Kos�ar, C.-J. Kuo, Y. Peles, Boiling heat transfer in rectangular microchannelswith reentrant cavities, Int. J. Heat Mass Transf. 48 (23–24) (2005) 4867–4886.

[11] R. Chen, M.C. Lu, V. Srinivasan, Z. Wang, H.H. Cho, A. Majumdar, Nanowires forenhanced boiling heat transfer, Nano letters 9 (2) (2009) 548–553.

[12] J. Bi, K. Vafai, D.M. Christopher, Heat transfer characteristics and CHFprediction in nanofluid boiling, Int. J. Heat Mass Transf. 80 (2015) 256–265.

[13] Y.H. Jeong, M.S. Sarwar, S.H. Chang, Flow boiling CHF enhancement withsurfactant solutions under atmospheric pressure, Int. J. Heat Mass Transf. 51(7–8) (2008) 1913–1919.

[14] T.I. Kim, Y.H. Jeong, S.H. Chang, An experimental study on CHF enhancement inflow boiling using Al2O3 nano-fluid, Int. J. Heat Mass Transf. 53 (5–6) (2010)1015–1022.

[15] C.J. Kuo, Y. Peles, Local measurement of flow boiling in structured surfacemicrochannels, Int. J. Heat Mass Transf. 50 (23–24) (2007) 4513–4526.

[16] F. Yang, X. Dai, Y. Peles, P. Cheng, J. Khan, C. Li, Flow boiling phenomena in asingle annular flow regime in microchannels (I): characterization of flowboiling heat transfer, Int. J. Heat Mass Transf. 68 (2014) 703–715.

[17] Y.M. Chen, F. Mayinger, Measurement of heat transfer at the phase interface ofcondensing bubbles, Int. J. Multiph. Flow 18 (6) (1992) 877–890.

[18] I. Ueno, Y. Hattori, R. Hosoya, Condensation and collapse of vapor bubblesinjected in subcooled pool, Microgravity Sci. Technol. 23 (1) (2010) 73–77.

[19] H.Y.a.A. Prosperetti, The pumping effect of growing and collapsing bubbles in atube, J. Micromech. Microeng. (1999).

[20] Z. Yin, A. Prosperetti, J. Kim, Bubble growth on an impulsively poweredmicroheater, Int. J. Heat Mass Transf. 47 (5) (2004) 1053–1067.

[21] S.G. Kandlikar, T. Widger, A. Kalani, V. Mejia, Enhanced flow boiling over openmicrochannels with uniform and tapered gap manifolds, J. Heat Transfer 135(6) (2013) 061401.

[22] F. Yang, X. Dai, C. Li, High frequency microbubble-switched oscillationsmodulated by microfluidic transistors, Appl. Phys. Lett. 101 (7) (2012).

[23] S.J. Kline, F.A. McClintock, Describing uncertainties in single-sampleexperiments, Mech. Eng. 75 (1) (1953) 3–8.

[24] W. Qu, I. Mudawar, Flow boiling heat transfer in two-phase micro-channelheat sinks––I. Experimental investigation and assessment of correlationmethods, Int. J. Heat Mass Transfer 46 (15) (2003) 2755–2771.

[25] W. Qu, I. Mudawar, Flow boiling heat transfer in two-phase micro-channelheat sinks––II. Annular two-phase flow model, Int. J. Heat Mass Transfer 46(15) (2003) 2773–2784.

[26] T. Alam, W. Li, F. Yang, W. Chang, J. Li, Z. Wang, J. Khan, C. Li, Force analysis andbubble dynamics during flow boiling in silicon nanowire microchannels, Int. J.Heat Mass Transf. 101 (2016) 915–926.

[27] L.Z. Evelyn, N. Wang, Linan Jiang, Jae-Mo Koo, James G. Maveety, Eduardo A.Sanchez, Micromachined Jets for Liquid Impingement Cooling of VLSI Chips, J.Microelectromech. Syst. 13 (2004) 833–841.

[28] M. Fabbri, V.K. Dhir, Optimized heat transfer for high power electronic coolingusing arrays of microjets, J. Heat Transfer 127 (7) (2005) 760.

[29] Y.C. Chen, C.F. Ma, M. Qin, Y.X. Li, Theoretical study on impingement heattransfer with single-phase free-surface slot jets, Int. J. Heat Mass Transf. 48(16) (2005) 3381–3386.

[30] F.H. Yang, M. Alwazzan, W.M. Li, C. Li, Single- and two-phase thermal transportin microchannels with embedded staggered herringbone mixers, J.Microelectromech. Syst. 23 (6) (2014) 1346–1358.