International Journal of Heat and Mass...

19
Micro-channel evaporator for space applications – 2. Assessment of predictive tools Hyoungsoon Lee a , Ilchung Park a , Issam Mudawar a,, Mohammad M. Hasan b a Purdue University Boiling and Two-Phase Flow Laboratory (PU-BTPFL), School of Mechanical Engineering, 585 Purdue Mall, West Lafayette, IN 47907, USA b NASA Glenn Research Center, 21000 Brookpark Road, Cleveland, OH 44135, USA article info Article history: Received 29 January 2014 Received in revised form 2 June 2014 Accepted 3 June 2014 Available online 1 July 2014 Keywords: Flow boiling Micro-channel Pressure drop Flow orientation Reduced gravity abstract This study is the second part of a two-part study addressing the effectiveness of micro-channel evapora- tors for space applications. The first part provided pressure drop and heat transfer data for FC-72 that were acquired with a test module containing 80 of 231 lm wide 1000 lm deep micro-channels. The tests were performed in three flow orientations: horizontal, vertical upflow and vertical downflow over broad ranges of mass velocity and heat flux. The present part uses these experimental results to assess the accuracy of published predictive tools. The two-phase heat transfer coefficient data are compared to predictions of 15 popular correlations, and pressure drop data to 7 mixture viscosity relations used in conjunction of the Homogeneous Equilibrium Model (HEM), and 18 correlations based on the Separated Flow Model (SFM). These models and correlations are carefully assessed in pursuit of identify- ing the most accurate tools. In addition, three important criteria for implementing micro-channel flow boiling in space systems are proposed: avoiding large pressure drop, avoiding critical heat flux (CHF), and negating the influence of body force. It is shown that micro-channels require significantly smaller mass velocities to negate body force effects than macro-channels, and are therefore very effective for space applications. Ó 2014 Elsevier Ltd. All rights reserved. 1. Introduction Pressure drop and heat transfer characteristics of single-phase micro-channel heat sinks have been the subject of extensive study spanning three decades, especially in conjunction with electronics cooling [1–6]. Aside from their very compact and lightweight design, these heat sinks provide unprecedented enhancement in heat transfer coefficient. Their thermal attributes are readily recog- nized for laminar flow, where the heat transfer coefficient is inver- sely proportional to hydraulic diameter. This implies the high heat transfer coefficient can be increased simply by decreasing hydrau- lic diameter. The advantages of utilizing a small diameter are also achieved with turbulent flow. But because these heat sinks rely on sensible heat rise of the coolant for heat dissipation, they typically incur large temperature gradients in both coolant and device being cooled. This shortcoming is largely responsible for shifting focus in recent years from single-phase to two-phase micro-channel heat sinks. With phase change, far greater heat transfer coefficients are achieved by capitalizing upon the coolant’s sensible and latent heat rather than sensible heat alone. This helps greatly reduce coolant flow rate when dissipating the same amount of heat as from a single-phase heat sink, let alone the reduction in coolant inventory for the cooling system at large. Reliance on latent heat exchange also enables two-phase heat sinks to achieve superior temperature uniformity. For several decades, flow boiling has been examined in predom- inantly macro-channel geometries as well as a variety of cooling configurations, including pool boiling [7–10], channel flow boiling [11–13], jet [14–17] and spray [18–21], as well as enhanced surfaces [22–24]. Despite many similarities between the boiling phenomena associated with these different configurations, extend- ing this understanding to flow boiling in micro-channels is by no means straightforward. Because of small hydraulic diameter, bub- ble nucleation, departure and coalescence are far more influenced by surface tension and wall confinement effects than macro- channels. These influences have a profound effect on two-phase regime transitions, pressure drop, heat transfer coefficient and critical heat flux (CHF). This paper is the second part of a two-part study concerning the influence of orientation on pressure drop and heat transfer http://dx.doi.org/10.1016/j.ijheatmasstransfer.2014.06.008 0017-9310/Ó 2014 Elsevier Ltd. All rights reserved. Corresponding author. Tel.: +1 (765) 494 5705; fax: +1 (765) 494 0539. E-mail address: [email protected] (I. Mudawar). URL: https://engineering.purdue.edu/BTPFL (I. Mudawar). International Journal of Heat and Mass Transfer 77 (2014) 1231–1249 Contents lists available at ScienceDirect International Journal of Heat and Mass Transfer journal homepage: www.elsevier.com/locate/ijhmt

Transcript of International Journal of Heat and Mass...

Page 1: International Journal of Heat and Mass Transferatsla.cau.ac.kr/wp-content/uploads/2018/05/J5.-Lee-et-al... · 2018-05-04 · 1232 H. Lee et al./International Journal of Heat and Mass

Micro-channel evaporator for space applications – 2. Assessmentof predictive tools

Hyoungsoon Lee a, Ilchung Park a, Issam Mudawar a,⇑, Mohammad M. Hasan b

a Purdue University Boiling and Two-Phase Flow Laboratory (PU-BTPFL), School of Mechanical Engineering, 585 Purdue Mall, West Lafayette, IN 47907, USAb NASA Glenn Research Center, 21000 Brookpark Road, Cleveland, OH 44135, USA

a r t i c l e i n f o

Article history:Received 29 January 2014Received in revised form 2 June 2014Accepted 3 June 2014Available online 1 July 2014

Keywords:Flow boilingMicro-channelPressure dropFlow orientationReduced gravity

a b s t r a c t

This study is the second part of a two-part study addressing the effectiveness of micro-channel evapora-tors for space applications. The first part provided pressure drop and heat transfer data for FC-72 thatwere acquired with a test module containing 80 of 231 lm wide ! 1000 lm deep micro-channels. Thetests were performed in three flow orientations: horizontal, vertical upflow and vertical downflow overbroad ranges of mass velocity and heat flux. The present part uses these experimental results to assessthe accuracy of published predictive tools. The two-phase heat transfer coefficient data are comparedto predictions of 15 popular correlations, and pressure drop data to 7 mixture viscosity relations usedin conjunction of the Homogeneous Equilibrium Model (HEM), and 18 correlations based on theSeparated Flow Model (SFM). These models and correlations are carefully assessed in pursuit of identify-ing the most accurate tools. In addition, three important criteria for implementing micro-channel flowboiling in space systems are proposed: avoiding large pressure drop, avoiding critical heat flux (CHF),and negating the influence of body force. It is shown that micro-channels require significantly smallermass velocities to negate body force effects than macro-channels, and are therefore very effective forspace applications.

! 2014 Elsevier Ltd. All rights reserved.

1. Introduction

Pressure drop and heat transfer characteristics of single-phasemicro-channel heat sinks have been the subject of extensive studyspanning three decades, especially in conjunction with electronicscooling [1–6]. Aside from their very compact and lightweightdesign, these heat sinks provide unprecedented enhancement inheat transfer coefficient. Their thermal attributes are readily recog-nized for laminar flow, where the heat transfer coefficient is inver-sely proportional to hydraulic diameter. This implies the high heattransfer coefficient can be increased simply by decreasing hydrau-lic diameter. The advantages of utilizing a small diameter are alsoachieved with turbulent flow. But because these heat sinks rely onsensible heat rise of the coolant for heat dissipation, they typicallyincur large temperature gradients in both coolant and device beingcooled.

This shortcoming is largely responsible for shifting focus inrecent years from single-phase to two-phase micro-channel heatsinks. With phase change, far greater heat transfer coefficients

are achieved by capitalizing upon the coolant’s sensible and latentheat rather than sensible heat alone. This helps greatly reducecoolant flow rate when dissipating the same amount of heat asfrom a single-phase heat sink, let alone the reduction in coolantinventory for the cooling system at large. Reliance on latent heatexchange also enables two-phase heat sinks to achieve superiortemperature uniformity.

For several decades, flow boiling has been examined in predom-inantly macro-channel geometries as well as a variety of coolingconfigurations, including pool boiling [7–10], channel flow boiling[11–13], jet [14–17] and spray [18–21], as well as enhancedsurfaces [22–24]. Despite many similarities between the boilingphenomena associated with these different configurations, extend-ing this understanding to flow boiling in micro-channels is by nomeans straightforward. Because of small hydraulic diameter, bub-ble nucleation, departure and coalescence are far more influencedby surface tension and wall confinement effects than macro-channels. These influences have a profound effect on two-phaseregime transitions, pressure drop, heat transfer coefficient andcritical heat flux (CHF).

This paper is the second part of a two-part study concerning theinfluence of orientation on pressure drop and heat transfer

http://dx.doi.org/10.1016/j.ijheatmasstransfer.2014.06.0080017-9310/! 2014 Elsevier Ltd. All rights reserved.

⇑ Corresponding author. Tel.: +1 (765) 494 5705; fax: +1 (765) 494 0539.E-mail address: [email protected] (I. Mudawar).URL: https://engineering.purdue.edu/BTPFL (I. Mudawar).

International Journal of Heat and Mass Transfer 77 (2014) 1231–1249

Contents lists available at ScienceDirect

International Journal of Heat and Mass Transfer

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

Page 2: International Journal of Heat and Mass Transferatsla.cau.ac.kr/wp-content/uploads/2018/05/J5.-Lee-et-al... · 2018-05-04 · 1232 H. Lee et al./International Journal of Heat and Mass

associated with flow boiling in micro-channels. This study is partof NASA’s Flow Boiling and Condensation Experiment (FBCE),which is projected for deployment in the International Space Sta-tion (ISS) in 2017. In part 1 [25], the pressure drop and heat trans-fer characteristics were examined experimentally for horizontalflow, vertical upward and vertical downflow for different massvelocities and orientations. The micro-channel module consists of80 parallel 231 lm wide ! 1000 lm deep micro-channels, andoperating conditions are provided in Table 1. This part will assesspredictive tools for pressure drop and heat transfer coefficient inpursuit of identifying the most accurate tools. Also discussed are

criteria for negating the influence of body force when implement-ing the micro-channel heat sink in reduced gravity space systems.

2. Assessment of heat transfer correlations

2.1. Heat transfer data reduction

As discussed in part 1 of this study [25], the local heat transfercoefficient is calculated by applying energy balance to the controlvolume of a unit cell consisting of a single micro-channel and

Nomenclature

b constant in Eq. (13)Bd Bond numberBo Boiling numberCc contraction coefficientcp specific heat at constant pressureDh hydraulic diameterE coefficient in two-phase heat transfer coefficient corre-

lationsf Fanning friction factorFq fraction of wall heat flux consumed in converting near-

wall liquid to vaporFr Froude numberG mass velocityg gravitational accelerationgn component of gravity normal to heated wallH mean thickness of phase layerh heat transfer coefficientHb distance between upper and lower thermocouplesHch height of micro-channelHp height of inlet/outlet plenumsHt distance between bottom wall of micro-channel and

upper thermocouplek thermal conductivityL length; length of micro-channelM molecular weightm fin parameter_m mass flow rate for single micro-channel

Nch number of micro-channelsNconf confinement numberNu Nusselt numberq00 heat fluxq00eff effective heat flux based on width of unit cell containing

single micro-channelq00m critical heat flux based on heated perimeterP pressure, perimeterDP pressure dropPR reduced pressurePr Prandtl numberRe Reynolds numberS suppression coefficient in two-phase heat transfer coef-

ficient correlationsSu Suratman numberT temperatureU mean velocity of phase layerv specific volumeWch width of single micro-channelWe Weber numberWp width of inlet/outlet plenumsWW width between micro-channelsX Lockhart–Martinelli parameterxe thermodynamic equilibrium quality

Xtt Lockhart–Martinelli parameter for turbulent liquid andturbulent vapor flows

z stream-wise coordinatez0 axial location where vapor layer velocity just exceeds

liquid layer velocityz⁄ axial location parameter in Eq. (13)

Greek Symbolsa void fractionb aspect ratio of micro-channel (Wch/Hch)d vapor layer thicknessg fin efficiencyh flow orientation angle from horizontal flow, percentage

predicted within ±30%kc critical wavelengthl dynamic viscosityn percentage predicted within ±50%q densityr surface tensionrc contraction ratio/ two-phase pressure drop multiplier

SubscriptsA accelarationalb bottom thermocouple planec contractioncb convective boilingch micro-channelcir circumferentiald developinge expansionexp experimentally-determinedF frictional, wettedf liquid, bulk fluidfd fully-developedfo liquid onlyG gravitationalg vaporH heatednb nucleate boilingpred predicteds solidsat saturatedsc subcooled boilingsp single-phasesub subcoolingt top thermocouple planetot totaltp two-phasew wall

1232 H. Lee et al. / International Journal of Heat and Mass Transfer 77 (2014) 1231–1249

Page 3: International Journal of Heat and Mass Transferatsla.cau.ac.kr/wp-content/uploads/2018/05/J5.-Lee-et-al... · 2018-05-04 · 1232 H. Lee et al./International Journal of Heat and Mass

two half sidewalls, and applying the fin analysis method. Themethod used here uses experimentally determined effective heatflux, q00eff , and difference between the local micro-channel bottomwall temperature, Tw,b, and local bulk fluid temperature, Tf.

h ¼q00eff ðWch þWwÞ

ðTw;b & Tf ÞðWch þ 2gHchÞ; ð1Þ

where q00eff , g, and m are the effective heat flux, fin efficiency and finparameter, which are defined, respectively, as q00eff ¼ ksðTb & TtÞ=Hb,

g ¼ tanhðmHchÞ=mHch, and m ¼ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi2h=ksWW

p. In these relations, Tb

and Tt are bottom and top temperatures measured by pairs ofthermocouples inserted in the copper block beneath themicro-channel at five axial locations. As discussed in [25], themicro-channel bottom wall temperature, Tb,w, and the bulk fluidtemperature, Tf, are given by Tw;b ¼ Tt & q00eff Ht=ks andTf ;nþ1 ¼ Tf ;n þ q00eff ðWch þWWÞDz=ð _mcp;f Þ for xe < 0, Tf = Tsat for 0 6 xe

6 1, and Tf ;nþ1 ¼ Tf ;n þ q00eff ðWch þWWÞDz=ð _mcp;gÞ for 1 < xe, wherethe saturation temperature, Tsat, is determined from the localsaturation pressure. The thermodynamic equilibrium quality alongthe channel is determined according to

xe ¼ &cp;f ðTsat & Tf Þ

hfgfor xe < 0; ð2aÞ

xe;nþ1 ¼ xe;n þq00eff ðWch þWWÞDz

_mhfgfor 0 6 xe 6 1; ð2bÞ

and

xe ¼ 1þcp;gðTf & TsatÞ

hfgfor 1 < xe: ð2cÞ

2.2. Assessment of correlations

The experimentally determined local two-phase heat transfercoefficient, htp, is identified for the spatial span where xe > 0 andcompared to predictions of several popular correlations. As shownin Table 2, there correlations are categorized into three differenttypes: (1) those based on either a nucleate boiling (nb) relationor convective boiling (cb) relation [26–33], (2) those based on themaximum of value determined from nb and cb relations [34–36],and (3) correlations based on superpositioning of nb and cb rela-tions [37,39–42]. It should be emphasized that, excepting the cor-relation of Kim and Mudawar [42], all other correlations weredeveloped for circular channels with uniform circumferential heat-ing, or rectangular channels with uniform four-sided heating.When assessing the predictive accuracy of these correlationsagainst the present data for rectangular channels with three-sidedheating, the relation htp ¼ htp;cirðNu3=Nu4Þ is used [42], where htp,cir

is the value predicted from the original correlation, and Nu3 and

Nu4 are the single-phase Nusselt numbers for laminar flow withthree-sided and four-sided heating, respectively.

Nu3 ¼ 8:235ð1& 1:883bþ 3:767b2 & 5:814b3 þ 5:361b4 & 2:0b5Þð3aÞ

and

Nu3 ¼ 8:235ð1& 1:883bþ 3:767b2 & 5:814b3 þ 5:361b4 & 2:0b5Þ:ð3bÞ

Table 2 shows several of the htp correlations employ a relationfor the single-phase heat transfer relation, hsp, which is determinedfrom hsp ¼ Nuðkf =DhÞ, where, for three-sided and four-sided heat-ing, respectively,

Nu ¼ Nu3 or Nu ¼ Nu4 for Ref < 2000 ð4aÞ

and

Nu ¼ 0:023Re0:8f Pr0:4

f for Ref > 2000 ð4bÞ

The multiplier Nu3/Nu4 is not applied to the correlation of Kimand Mudawar [42], which accounts for partial or full circumferen-tial heating by the ratio of heated perimeter to wetted perimeter,PH/PF.

Figs. 1–3 compare the experimentally determined local two-phase heat transfer coefficient corresponding to xe > 0 at fivedifferent axial locations where the copper block temperatures aremeasured. The predictive accuracy of a correlation is determinedby mean absolute error, which is defined as

MAE ¼ 1N

X htp;pred & htp;exp

htp;exp

""""

""""! 100%: ð5Þ

Two additional parameters, h and n, indicate the percentages ofdata points predicted within ±30% and ±50%, respectively. Sub-scripts T1 to T5 indicate the five axial locations, z = 12.7, 44.5,76.2, 108.0 and 137.7 mm, where the copper block temperaturesare measured. The open symbols represent two-phase heat trans-fer coefficient data measured during tests incurring severe pressuredrop oscillation [25]; these data points are excluded from calcula-tions of MAE, h and n since the correlations are not intended forunstable conditions.

Fig. 1 compares the experimentally-determined local two-phase heat transfer coefficient, htp, at the five different axial loca-tions with predictions of the first group of correlations in Table 2that are based on a nucleate boiling (nb) or convective boiling(cb) relation. Separate comparisons are shown for the horizontalflow, vertical upflow and vertical downflow data. Notice that thedata designated by open symbols, which correspond to severepressure drop oscillation and are excluded from calculations ofMAE, h and n, show large deviations from both the other dataand predictions. Overall, the correlations in this group give poor

Table 1Operating conditions of present study.

Horizontal Vertical upward Vertical downward

Pressure drop Heat transfer Pressure drop Heat transfer Pressure drop Heat transfer

G (kg/m2 s) 155.9–792.0 196.6–792.0 151.5–834.5 259.2–834.5 175.4–772.3 175.4–772.3G/qf (m/s) 0.10–0.50 0.13–0.50 0.10–0.53 0.17–0.53 0.11–0.49 0.11–0.49q00eff (W/cm2) 2.19–9.46 3.10–9.41 2.14–9.56 3.43–9.56 2.47–9.89 3.28–9.30Pin (kPa) 188.1–297.3 188.1–297.3 187.8–292.2 187.8–292.2 189.3–306.5 189.3–306.5Tsat ("C) 76.5–93.1 76.5–93.1 76.4–92.5 76.4–92.5 76.7–94.3 76.7–94.3Tin ("C) 60.5–83.2 60.5–81.0 60.2–83.3 60.2–81.8 60.8–84.1 60.8–84.1xe,in &0.29 to &0.09 &0.29 to &0.09 &0.28 to &0.10 &0.28 to &0.11 &0.29 to &0.09 &0.29 to &0.09PR 0.09–0.16 0.10–0.16 0.10–0.16 0.10–0.16 0.09–0.17 0.10–0.17Data points 97 65 80 65 97 86

H. Lee et al. / International Journal of Heat and Mass Transfer 77 (2014) 1231–1249 1233

Hyoungsoon
Highlight
Hyoungsoon
Highlight
Hyoungsoon
Highlight
Hyoungsoon
Highlight
Hyoungsoon
Rectangle
Hyoungsoon
Typewriter
Nu4
Page 4: International Journal of Heat and Mass Transferatsla.cau.ac.kr/wp-content/uploads/2018/05/J5.-Lee-et-al... · 2018-05-04 · 1232 H. Lee et al./International Journal of Heat and Mass

Table 2Saturated boiling heat transfer correlations.

Author(s) Correlation Remarks

Correlations based on either a nucleate boiling (nb) relation or convective boiling (cb) relationLazarek and Black

[26]htp ¼ 30Re0:857

fo Bo0:714 kfDh

, Refo ¼ GDhlf

, Bo ¼ q00HGhfg

nb-Based, single circular copper tubes, Dh = 3.15 mm,R133, vertical upflow and downflow, G = 125–750 kg/m2 s, q00H ¼0–40 W/cm2, xe = 0–0.8

Cooper [27] htp ¼ 55P0:12R ð&log10PRÞ&0:55M&0:5q000:67

Hnb-Based, pool boiling, stainless steel, copper, platinum,nickel, aluminum, brass, circular, rectangular, wire,hydrogen, deuterium, helium, methane, water, neon,nitrogen, ethane, methanol, oxygen, propane, ethanol, n-butane, benzene, R11, R12, R13, R113, R114, R115, R21,R22, R13B1, R226, RC318, SF6, >6000 data points fromover100 sources

Tran et al. [28]htp ¼ 8:4! 105ðBo2WefoÞ

0:3 qgqf

# $0:4, Wefo ¼ G2 Dh

qf rnb-Based, single circular/rectangular tubes, Dh = 2.4 mmfor rectangular, 2.46, 2.92 mm for circular, R12, R113,horizontal, G = 44–832 kg/m2 s, q00H ¼0.36–12.9 W/cm2,xe = 0.2–0.94, 249 data points

Lee and Lee [29] htp ¼ Ehsp cb-Based, single rectangular, Dh = 0.784–3.636 mm,R113, horizontal, stainless steel, G = 50–200 kg/m2 s,q00H ¼1.5 W/cm2, xe = 0.15–0.75

hsp ¼ 0:023Re0:8fo Pr0:4

fkfDh

, E ¼ 10:3b0:398/0:598f

/f ¼ 1þ CX þ

1X2

# $0:5, C ¼ 6:185! 10&2Re0:726

fo , Refo ¼ GDhlf

Warrier et al. [30] htp ¼ Ehsp nb-Based, multi (N = 5) rectangular aluminum,Dh = 0.75 mm, FC84, horizontal, G = 557–1600 kg/m2 s,q00H ¼0–5.99 W/cm2, xe = 0–0.5

E ¼ 1:0þ 6:0Bo1=16 & 5:3ð1& 855BoÞx0:65e , htp ¼ 0:023Re0:8

fo Pr0:4f

kfDh

Agostini andBontemps [31]

for xe 6 0:43, htp ¼ 28q002=3H G&0:26x&0:10

enb-Based, multi (N = 11) rectangular aluminum,Dh = 2.01 mm, R134a, vertical upflow, G = 90–295 kg/m2 s, q00H ¼0.6–3.16 W/cm2, xe = 0.26–1

for xe > 0.43, htp ¼ 28q002=3H G&0:64x&2:08

e

Li and Wu [32]htp ¼ 334Bo0:3 BdRe0:36

f

# $0:4 kfDh

# $, Bd ¼ gðqf&qg ÞD

2h

rnb-Based, single/multi circular/rectangular, Dh = 0.16–3.01 mm, R11, R12, R123, R134a, R141b, R22, R236fa,R245fa, FC77, FC84, water, CO2, propane, horizontal,vertical upflow, G = 23.4–3570.0 kg/m2 s, q00H ¼0–115.0 W/cm2, xe = 0–1, 3744 data points from 26 sources

Oh and Son [33]htp ¼ 0:034Re0:8

f Pr0:3f 1:58 1

Xtt

# $0:87% &

kfDh

# $ cb-Based, single circular copper, Dh = 1.77, 3.36,5.35 mm, R134a, R22, horizontal

Correlations based on maximum of values determined from nucleate boiling (nb) and convective boiling (cb) relationsShah [34,35] htp ¼ Max ðE; SÞhsp Single, circular, R11, R113, R12, R22, water, cyclohexane,

horizontal, vertical upflow/downflow, 780 data pointsfrom 19 sources

S = 1.8/N0.8, hsp ¼ 0:023Re0:8f Pr0:4

fkfDh

for N > 1.0, E ¼ 1þ 46Bo0:5 for Bo < 3 ! 10&5

E ¼ 230Bo0:5 for Bo > 3 ! 10&5

for 0:1 < N 6 1:0, E ¼ FBo0:5 expð2:74N&0:1Þfor N 6 0:1, E ¼ F Bo0:5 expð2:47N&0:15ÞF = 14.7 for Bo P 11! 10&4 or F = 15.43 for Bo < 11 ! 10&4

N = Co for vertical tube,N = Co for horizontal tube with Frf P 0:04

N ¼ 0:38Fr&0:3f Co for horizontal tube with Frf < 0.04

Ref ¼Gð1&xÞDh

lf, Co ¼ 1&x

x

' (0:8 qgqf

# $0:5, Frf ¼ G2

q2f gDh

Ducoulomnieret al. [36]

htp ¼ Max ðhnb; hcbÞ Single, circular stainless steel, Dh = 0.529 mm, CO2,horizontal, G = 200–1200 kg/m2 s, q00H ¼1–3 W/cm2, 2710data points

hnb ¼ 131P&0:0063R ð&log10PRÞ&0:55M&0:5q000:58

H ,for Bo > 1.1 ! 10&4,

hcb ¼ 1:47! 104 Boþ 0:93 1Xtt

# $2=3) *

hsp;fo

for Bo < 1.1 ! 10&4, hcb ¼ 1þ 1:80 1Xtt

# $0:986) *

hsp

hsp ¼ 0:023Re0:8fo Pr0:4

fkfDh

Correlations based on superpositioning of nucleate boiling (nb) and convective boiling (cb) relationsChen [37] htp ¼ Ehsp þ Shnb Water, methanol, cyclohexan, pentane, heptane,

benzene, vertical upflow/downflow, over 600 datapoints from 10 sourceshnb ¼ 0:00122

k0:79f c0:45

p;f v0:24g

r0:5l0:29f h0:24

fg v0:49f

+ ,DT0:24

sat DP0:75sat

E ¼ ð1þ 1X0:5Þ

1:78, S ¼ 0:9622& 0:5822 tan&1 Ref E1:25

6:18!104

# $

E, S from Edelstein et al. [38]

Gungor andWinterton [39]

htp ¼ Ehsp þ Shnb , hsp ¼ 0:023Re0:8fo Pr0:4

fkfDh

Circular, Dh = 2.95–32 mm, R11, R113, R114, R12, R22,water, ethylene glycol, horizontal, vertical upflow/downflow, G = 12.4–8179.3 kg/m2 s, q00H ¼0.03–262 W/cm2, 4300 data points from 28 sources

E ¼ 1þ 24000Bo1:16 þ 1:37 1Xtt

# $0:86, S ¼ 1þ 1:15! 10&6E2 Re1:17

f

# $&1

hnb ¼ htp;Cooper

Liu and Winterton[40]

htp ¼ ½ðEhspÞ2 þ ðShnbÞ2(

0:5, hsp ¼ 0:023Re0:8

fo Pr0:4f

kfDh

Same data as Gungor and Winterton [39]

for horizontal tube with Frf 6 0:05

1234 H. Lee et al. / International Journal of Heat and Mass Transfer 77 (2014) 1231–1249

Hyoungsoon Lee
Hyoungsoon Lee
Hyoungsoon
Highlight
Page 5: International Journal of Heat and Mass Transferatsla.cau.ac.kr/wp-content/uploads/2018/05/J5.-Lee-et-al... · 2018-05-04 · 1232 H. Lee et al./International Journal of Heat and Mass

predictions of the closed symbols, corresponding to tests free fromsevere pressure drop oscillation, with those of Lazarek and Black[26], Fig. 1(a), Cooper [27], Fig. 1(b), Tran et al. [28], Fig. 1(c), Leeand Lee [29], Fig. 1(d), and Warrier et al. [30], Fig. 1(e), underpre-dicting the present data. The correlations by Oh and Son [33],Fig. 1(g), and of Li and Wu [32], Fig. 1(h), show weak trends, evi-denced by large MAE values. Only the correlation by Agostini andBontemps [31], Fig. 1(f), which was developed from vertical upflowmulti-channel data gives relatively good predictions, withMAE = 17.2%, 16.1%, and 13.4%, for the present horizontal flow, ver-tical upflow, and vertical downflow data, respectively.

Fig. 2 compares the experimentally determined local two-phaseheat transfer coefficient with predictions of the second group ofcorrelations, where the heat transfer coefficient is based on themaximum value from nucleate boiling (nb) and convective boiling(cb) relations. Both correlations yield fair accuracy, albeit with sig-nificant scatter, with that of Shah [34,35], Fig. 2(a), yielding MAEsof 27.8%, 26.8% and 24.1% for horizontal flow, vertical upflow andvertical downflow, respectively, and corresponding MAEs of thecorrelation by Ducoulomnier et al. [36], Fig. 2(b), of 25.7%, 24.8%and 21.7%.

Fig. 3 compares the experimentally determined local two-phaseheat transfer coefficient with predictions of the third group of cor-relations, which involve superpositioning of nucleate boiling (nb)and convective boiling (cb) relations. The correlations in this groupshow relatively good predictions for the different orientations, withthe recent correlation by Kim and Mudawar [42], Fig. 3(e), yieldingthe best predictions, with MAEs of 19.0%, 19.9% and 16.9% for hor-izontal flow, vertical upflow and vertical downflow, respectively.The superior accuracy of this correlation can be traced to its relianceon the largest consolidated database consisting of 10,805 mini/micro-channel data points amassed from 31 sources, including 18working fluids, hydraulic diameters of 0.19–6.5 mm, massvelocities of 19–1608 kg/m2 s, liquid-only Reynolds numbers of57–49,820, qualities of 0–1, and reduced pressures of 0.005–0.69.

Figs. 4–6 compare, for six narrow ranges of mass velocity, thevariations of experimentally determined local heat transfer coeffi-cient with quality for horizontal flow, vertical upflow and vertical

downflow with the predictions of different correlations. Fig. 4compares data with the first group of correlations based on nb orcb relations. Fig. 5 shows comparisons with the second group ofcorrelations based on the maximum of values determined fromnb and cb relations. Fig. 6 compares data with the third group ofcorrelations based on superpositioning of nb and convective boil-ing cb relations. Excepting the highest mass velocity range ofG = 607.6–644.5 kg/m2 s, the data show increasing xe causes an ini-tial upstream decrease in htp, followed by an increase in htp down-stream. However, Figs. 4 and 5 show that most correlations fromthe first two groups of correlations fail to capture the trend ofdecreasing htp for low xe. Correlations from the first group thatdo capture this trend are those of Warrier et al. [30] and Agostiniand Bontemps [31], the latter being the one that proved most accu-rate in the prediction of htp as shown in Fig. 1(f). Fig. 5 shows thatthe correlations by Shah [34,35] and Ducoulomnier et al. [36] alsofail to capture the trend of decreasing htp for low xe. Fig. 6 showscorrelations from the third group are also unable to capture thistrend. Overall, the third group of correlations shows better predic-tions with increasing xe and increasing G.

3. Assessment of pressure drop correlations

3.1. Pressure drop determination

As described in part 1 of this study [25], the total pressure drop,DPtot, is determined from the difference in pressures measured byabsolute pressure transducers connected to the inlet and outlet ofthe test module. The total pressure drop consists of several compo-nents, including upstream contraction, DPc, upstream subcooledliquid region, DPsp,f, where xe < 0, saturated two-phase region, DPtp,where 0 < xe < 1, superheated single-phase vapor region, DPsp,g,where xe > 0, and downstream expansion, DPe. It should be empha-sized that not all components are encountered in a test. For exam-ple, DPsp,f = 0 when the flow enters the channels in saturated liquidstate or as a two-phase mixture. Additionally, for most operatingconditions, DPsp,g = 0, since tests are terminated before wall dryoutor critical heat flux (CHF) are incurred.

Table 2 (continued)

Author(s) Correlation Remarks

E ¼ Frð0:1&2Frf Þf

h i1þ xePrf

qfqg& 1

# $h i0:35,

S ¼ ðFr0:5f Þ 1þ 0:055E0:1 Re0:16

fo

# $&1

for vertical tube and horizontal tube with Frf > 0:05

E ¼ 1þ xe Prfqfqg& 1

# $h i0:35, S ¼ 1þ 0:055E0:1Re0:16

fo

# $&1

hnb ¼ htp;Cooper

Bertsch et al. [41] htp ¼ Ehcb þ Shnb Circular/rectangular, Dh = 2.95–32 mm, R11, R113, R114,R12, R22, water, ethylene glycol, horizontal, verticalupflow/downflow, G = 20–3000 kg/m2 s, q00H ¼0.4–115 W/cm2, xe = 0–1, 3899 data points from 14 sources

hcb ¼ hsp;foð1& xeÞ þ hsp;goxe

E ¼ 1þ 80 x2e & x6

e' (

expð&0:6Nconf Þ, S ¼ 1& xe

hsp;fo ¼ 3:66þ 0:668DhL Refo Prf

1þ0:04DhL Refo Prf

- .2=3

!kfDh

,

hsp;go ¼ 3:66þ 0:668DhL Rego Prg

1þ0:04DhL Rego Prg

- .2=3

!kgDh

Nconf ¼ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

rgðqf&qg ÞD

2h

q, Refo ¼ GDh

lf, Rego ¼ GDh

lg

Kim and Mudawar[42]

htp ¼ h2nb þ h2

cb

# $0:5 Circular/rectangular, Dh = 0.19–6.5 mm, FC72, R11, R113,R123, R1234yf, R1234ze, R134a, R152a, R22, R236fa,R245fa, R32, R404A, R407C, R410A, R417A, CO2, water,horizontal, vertical upflow/downflow, G = 19–1608 kg/m2 s, xe = 0–1, 12,974 data points from 31 sources

hnb ¼ 2345 Bo PH PF

# $0:70P0:38

R ð1& xeÞ&0:51% &

0:023Re0:8f Pr0:4

fkfDh

# $

hcb ¼ 5:2 Bo PHPF

# $0:08We&0:54

fo þ 3:5 1Xtt

# $0:94 v fvg

# $0:25% &

0:023Re0:8f Pr0:4

fkfDh

# $

Bo ¼ q00HGhfg

, PR ¼ PPcrit

, Ref ¼Gð1&xeÞDh

lf, Wefo ¼ G2Dh

qf r,

Xtt ¼lflg

# $0:11&xe

xe

# $0:9 qgqf

# $0:5

H. Lee et al. / International Journal of Heat and Mass Transfer 77 (2014) 1231–1249 1235

Hyoungsoon Lee
Hyoungsoon Lee
Hyoungsoon Lee
Dh = 0.16 - 2.92mm
Hyoungsoon
Rectangle
Page 6: International Journal of Heat and Mass Transferatsla.cau.ac.kr/wp-content/uploads/2018/05/J5.-Lee-et-al... · 2018-05-04 · 1232 H. Lee et al./International Journal of Heat and Mass

Depending on operating conditions, pressure drop for theupstream subcooled region where xe < 0 can be further divided intothree sub-regions: single-phase liquid developing region, DPsp,d,single-phase liquid fully-developed region, DPsp,fd, and subcooledboiling region, DPsc. Pressure drop in the saturated region, DPtp,is comprised of three components: accelerational, DPtp,A, gravita-tional, DPtp,G, and frictional, DPtp,F .

For the conditions of the present study, the single-phase liquiddeveloping and subcooled portions of the subcooled regionaccount for only a small fraction of the total length of the micro-channel. Therefore, DPsp,d and DPsc are combined into DPsp,f, andtotal pressure drop is represented by

DPtot ¼ DPc þ DPsp;f þ ðDtp;A þ DPtp;G þ DPsp;g þ DPtp;FÞ þ DPsp;g þ DPe:

ð6Þ

Part 1 of the present study included relations used to calculateDPc, DPsp,f, DPsp,g, DPe, DPtp,A and DPtp,G, which are also included inTable 3. Discussed below are details of the two-phase frictionalpressure drop component, DPtp,F, which accounts for the largestfraction of DPtot.

3.2. Two-phase pressure drop models

Two general approaches are used to determine pressure dropfor the saturated two-phase flow region: Homogeneous Equilibrium

Model (HEM) and Separated Flow Model (SFM). Following is a briefdiscussion of these models and associated assumptions.

3.2.1. Pressure drop data comparison with predictions HomogeneousEquilibrium Model (HEM)

By employing a variety of averaging techniques to evaluatemixture properties, HEM treats a two-phase mixture as a pseudofluid that obeys standard conservation relations similar to thoseemployed with single-phase flow. Based on the assumptions ofnegligible kinetic and potential energy changes and negligibleflashing and compressibility, the two-phase pressure gradient fora channel with a constant flow area can be expressed as the sumof frictional, accelerational and gravitational components.

& dPdz

+ ,

tp¼ 2

DhftpG2v f ð1þ xe

v fg

v fÞ þ G2v fg

dxe

dzþ g sinh

v f þ xev fg; ð7Þ

The two-phase friction factor, ftp, in Eq. (7) is obtained from[42,43]

ftpRetp ¼ 24 1& 1:3553bþ 1:9467b2 & 1:7012b3-

þ0:9564b4 & 0:2537b5. for Retp < 2000 ð8aÞ

ftp ¼ 0:079Re&0:25tp for 2000 < Retp < 20;000; ð8bÞ

ftp ¼ 0:046Re&0:2tp for 20;000 < Retp; ð8cÞ

Fig. 1. Comparison of experimentally determined local heat transfer coefficient at five different axial locations for horizontal flow, vertical upflow and vertical downflow withpredictions based on nucleate boiling (nb) or convective boiling (cb) correlations of (a) Lazarek and Black [26], (b) Cooper [27], (c) Tran et al. [28], (d) Lee and Lee [29], (e)Warrier et al. [30], (f) Agostini and Bontemps [31], (g) Li and Wu [32], and (h) Oh and Son [33]. Open symbols correspond to data for severe pressure oscillation, which are notincluded in the statistical assessment of correlations.

1236 H. Lee et al. / International Journal of Heat and Mass Transfer 77 (2014) 1231–1249

Page 7: International Journal of Heat and Mass Transferatsla.cau.ac.kr/wp-content/uploads/2018/05/J5.-Lee-et-al... · 2018-05-04 · 1232 H. Lee et al./International Journal of Heat and Mass

where Retp = GDh/ltp, based on mixture viscosity ltp. Table 4 pro-vides a summary of two-phase viscosity relations evaluated in thepresent study. When calculating the two-phase frictional pressuregradient in order to determine DPtp, Eq. (7) is integrated betweenz = 0 and L.

Fig. 7(a)–(g) show comparisons of the experimental data withpredictions of the HEM using the seven mixture viscosity relationsfrom Table 4. The predictive accuracy is determined by mean abso-lute error, which is defined as

MAE ¼ 1N

X DPtot;pred & DPtot;exp

DPtot;exp

""""

""""! 100%; ð9Þ

as well as h and n, which are defined as the percentages of datapoints predicted within ±30% and ±50%, respectively. Predictionsusing the viscosity relations of McAdams et al. [46], Akers et al.[47], Dukler et al. [50] and Lin et al. [52] show fairly good agreementwith the experimental data, while those using the relations ofCicchiti et al. [48], Owens [49] and Beattie and Whalley [51] highlyoverpredict the data especially for low mass velocities, but showimproved accuracy at high mass velocities. Of all seven viscosityrelations used in conjunction with the HEM, the relation by Linet al. shows the best accuracy, with MAEs of 22.0%, 28.9% and22.8% for horizontal flow, vertical upflow and vertical downflow,respectively.

3.2.2. Pressure drop data comparison with predictions of SeparateFlow Model (SFM)

Unlike the HEM, SFMs allow for differences between liquid andvapor velocities, with each phase occupying a separate portion ofthe flow area. Based on the SFM, the total pressure gradient forsteady flow in a channel with a constant flow area is given by [53]

& dPdz

+ ,

tp¼ & dP

dz

+ ,

tp;F& dP

dz

+ ,

tp;A& dP

dz

+ ,

tp;G; ð10Þ

where

& dPdz

+ ,

tp;A¼ G2 d

dzx2

evg

a þð1& xeÞ2v f

1& a

" #ð11aÞ

and

& dPdz

+ ,

tp;G¼ ½aqg þ ð1& aÞqf (g sinh: ð11bÞ

The void fraction, a, in the accelerational and gravitationalterms can be determined using Zivi’s correlation [45],

a ¼ 1þ 1& xe

xe

+ , qg

qf

!2=32

4

3

5&1

: ð12Þ

The total pressure gradient, DPtp, is obtained by integrating Eq.(10) between z = 0 and L using an appropriate correlation for thefrictional pressure gradient.

Table 5 summarizes popular correlations for the two-phase fric-tional pressure gradient based on the SFM. The first eleven correla-tions [54–64] were not specifically developed, but have beenwidely used for micro-channel applications, while the last sevencorrelations [53,65–70] were developed from mini/micro-channelsexperimental data. Comparisons between the experimental dataand predictions of these correlations are provided in Figs. 8–10.The predictions using the first eleven correlations are shown inFigs. 8 and 9. The correlations by Lockhart and Martinelli [54],Friedel [55], Müller-Steinhagen and Heck [56], Mishima and Hibiki[58], Yang and Webb [59], Wang et al. [60], and Yan and Lin [61]show fair predictions of the data, with MAEs in the range of28.7–39.5%, 22.3–42.3%, and 31.5–40.0% for horizontal flow,vertical upflow and vertical downflow, respectively. On the otherhand, the correlations by Jung and Radermacher [57], Tran et al.[62], Chen et al. [63] and Yu et al. [64] yield far less favorablepredictions. Notice that these correlations show no obvious supe-rior accuracy relative to flow orientation.

Fig. 2. Comparison of experimentally determined local heat transfer coefficient at five different axial locations for horizontal flow, vertical upflow and vertical downflow withpredictions based on maximum of value from relations for nucleate boiling (nb) and convective boiling (cb) according to (a) Shah [34,35] and (b) Ducoulomnier et al. [36].Open symbols correspond to data for severe pressure oscillation, which are not included in the statistical assessment of correlations.

H. Lee et al. / International Journal of Heat and Mass Transfer 77 (2014) 1231–1249 1237

Page 8: International Journal of Heat and Mass Transferatsla.cau.ac.kr/wp-content/uploads/2018/05/J5.-Lee-et-al... · 2018-05-04 · 1232 H. Lee et al./International Journal of Heat and Mass

Fig. 3. Comparison of experimentally determined local heat transfer coefficient at five different axial locations for horizontal flow, vertical upflow and vertical downflow withpredictions of correlations based on superpositioning of nucleate boiling (nb) and convective boiling (cb) relations by (a) Chen [37], (b) Gungor and Winterton [39], (c) Liu andWinterton [40], (d) Bertsch et al. [41], and (e) Kim and Mudawar [42]. Open symbols correspond to data for severe pressure oscillation, which are not included in thestatistical assessment of correlations.

1238 H. Lee et al. / International Journal of Heat and Mass Transfer 77 (2014) 1231–1249

Page 9: International Journal of Heat and Mass Transferatsla.cau.ac.kr/wp-content/uploads/2018/05/J5.-Lee-et-al... · 2018-05-04 · 1232 H. Lee et al./International Journal of Heat and Mass

Fig. 4. Comparison of variation of experimentally determined local heat transfer coefficient with quality for horizontal flow, vertical upflow and vertical downflow withpredictions of nucleate boiling (nb) or convective boiling (cb) correlations for (a) G = 180.2–199.9 kg/m2 s, (b) G = 267.3–302.1 kg/m2 s, (c) G = 357.6–362.1 kg/m2 s, (d)G = 455.0–456.6 kg/m2 s, (e) G = 550.8–573.0 kg/m2 s, and (f) G = 607.6–644.5 kg/m2 s.

H. Lee et al. / International Journal of Heat and Mass Transfer 77 (2014) 1231–1249 1239

Page 10: International Journal of Heat and Mass Transferatsla.cau.ac.kr/wp-content/uploads/2018/05/J5.-Lee-et-al... · 2018-05-04 · 1232 H. Lee et al./International Journal of Heat and Mass

Fig. 5. Comparison of variation of experimentally determined local heat transfer coefficient with quality for horizontal flow, vertical upflow and vertical downflow withpredictions of correlations based on maximum of value from relations for nucleate boiling (nb) and convective boiling (cb) for (a) G = 180.2–199.9 kg/m2 s, (b) G = 267.3–302.1 kg/m2 s, (c) G = 357.6–362.1 kg/m2 s, (d) G = 455.0–456.6 kg/m2 s, (e) G = 550.8–573.0 kg/m2 s, and (f) G = 607.6–644.5 kg/m2 s.

1240 H. Lee et al. / International Journal of Heat and Mass Transfer 77 (2014) 1231–1249

Page 11: International Journal of Heat and Mass Transferatsla.cau.ac.kr/wp-content/uploads/2018/05/J5.-Lee-et-al... · 2018-05-04 · 1232 H. Lee et al./International Journal of Heat and Mass

(a)

(b)

(c)

(d)

(e)

(f)Fig. 6. Comparison of variation of experimentally determined local heat transfer coefficient with quality for horizontal flow, vertical upflow and vertical downflow withpredictions of correlations based on superpositioning of nucleate boiling (nb) and convective boiling (cb) relations for (a) G = 180.2–199.9 kg/m2 s, (b) G = 267.3–302.1 kg/m2 s, (c) G = 357.6–362.1 kg/m2 s, (d) G = 455.0–456.6 kg/m2 s, (e) G = 550.8–573.0 kg/m2 s, and (f) G = 607.6–644.5 kg/m2 s.

H. Lee et al. / International Journal of Heat and Mass Transfer 77 (2014) 1231–1249 1241

Page 12: International Journal of Heat and Mass Transferatsla.cau.ac.kr/wp-content/uploads/2018/05/J5.-Lee-et-al... · 2018-05-04 · 1232 H. Lee et al./International Journal of Heat and Mass

The predictions of the correlations developed from mini/micro-channel data are shown in Fig. 10. Overall, these correlations showbetter accuracy than those in Figs. 8 and 9. The correlations of Leeand Lee [65], Hwang and Kim [66] and Zhang et al. [69] slightlyunderpredict the data, while those of Sun and Mishima [67] and Liand Wu [68,70] overpredict. These correlations to do not appear toprovide superior accuracy relative to any particular flow orientation.

Among all the correlations examined in Figs. 8–10, Kim andMudawar’s [53] shows the best predictive capability, with MAEsof 24.2%, 30.3% and 24.1% for horizontal flow, vertical upflow, andvertical downflow, respectively. The superior accuracy of this corre-lation can be attributed to reliance on the most recent and mostcomprehensive consolidated database consisting of 2378 mini/micro-channel data points from 16 sources. The database encom-passes 9 working fluids, hydraulic diameters from 0.349 to5.35 mm, mass velocities from 33 to 2738 kg/m2 s, liquid-only Rey-nolds numbers from 156 to 28,010, qualities from 0 to 1, reducedpressures from 0.005 to 0.78, and both single-channel and multi-channel data.

4. Design criteria for reduced gravity space systems

4.1. Rationale

Three important criteria for implementing micro-channel flowboiling in space systems are:

(1) Avoiding the large pressure drops associated with two-phase flow boiling in micro-channels to minimize powerconsumption in a space vehicle’s thermal control system(TCS).

(2) Avoiding critical heat flux (CHF) in the TCS evaporator.(3) Negating the influence of body force on two-phase flow and

heat transfer.

Discussed below are means to simultaneously achieve thesecriteria.

4.2. Avoiding high pressure drop

High pressure drop is a primary concern when implementingflow boiling in micro-channels. This concern stems from both theuse of small hydraulic diameter, and appreciable vapor productionwithin the micro-channel, especially when dissipating high heatfluxes.

In the first part of this study [25], pressure drop data were pre-sented for flow boiling in micro-channels in three different orien-tations. Two-phase friction, followed by two-phase accelerationand single-phase liquid friction, were identified as dominant com-ponents of pressure drop. As shown in [25] and Eq. (8) of the pres-ent study, these three components of pressure drop areproportional to G2. In addition, both the two-phase frictional andaccelerational components increase with increasing xe, i.e., increas-ing q00eff . Compounding operation at high G and q00eff are two-phaseinstabilities. Given the predictive tools presented in this study fortwo-phase pressure drop, the designer of a space vehicle’s TCScan conduct a parametric study of evaporator design by exploringdifferent combinations of operating conditions, overall evaporatorsurface area, and micro-channel shape and size to minimize pres-sure drop and to avoid instabilities.

4.3. Avoiding critical heat flux (CHF)

Critical heat flux (CHF) is unquestionably the most importantdesign parameter for any two-phase system that is subjected to aprescribed heat load, and avoiding CHF requires maintaining heatflux safely below CHF. Recent studies at the Purdue University

Table 3Pressure drop components [44].

Correlation

Area changesDPc ¼

G2v f2

1Cc& 1

# $2þ 1& r2

c' (% &

1þ v fg xe;inv f

h i

DPe ¼ G2rcðrc & 1Þv f 1þ v fg xe;out

v f

h i

Cc ¼ 1& 1&rc2:08ð1&rc Þþ0:5371, rc ¼ Wch Hch Nch

WpHp

Single-phase liquid and vapor regions DPsp;k ¼2Lsp;k

Dhfsp;kG2vk

for Rek < 2000 : fsp;k Rek ¼ 24½1& 1:3553bþ 1:9467b2 & 1:7012b3 þ 0:9564b4 & 0:2537b5(for 2000 < Rek < 20;000 : fsp;k ¼ 0:079Re&0:25

k

for 20;000 < Rek : fsp;k ¼ 0:046Re&0:2k

Rek ¼ GDhlk

k = f for liquid, k = g for vapor

Two-phase region Accelerational:

DPtp;A ¼R Ltp

0 &dPdz

' (Adz

& dPdz

' (A ¼ G2 d

dzx2

e vga þ

ð1&xeÞ2v f1&a

j k

Gravitational:

DPtp;G ¼R Ltp

0 &dPdz

' (Gdz

& dPdz

' (G ¼ !qg sin h ¼ a

vgþ ð1&aÞ

v f

h ig sin h

Void fraction:

a ¼ 1þ 1&xexe

# $v fvg

# $h i&1when used with HEM, a ¼ 1þ 1&xe

xe

# $v fvg

# $2=3% &&1

when used with SFM [45]

Table 4Mixture viscosity relations used in HEM.

Viscosity relation

McAdams et al. [46] 1ltp¼ xe

lgþ 1&xe

lf

Akers et al. [47] ltp ¼lf

ð1&xeÞþxevgvf

# $0:5% &

Cicchitti et al. [48] ltp ¼ xelg þ ð1& xeÞlf

Owens [49] ltp ¼ lf

Dukler et al. [50] ltp ¼xevglgþð1&xe Þv f lf

xevgþð1&xeÞv f

Beattie and Whalley [51] ltp ¼ xlg þ ð1&xÞð1þ 2:5xÞlf

x ¼ xevgv fþxev fg

Lin et al. [52] ltp ¼lf lg

lgþx1:4e ðlf&lg Þ

1242 H. Lee et al. / International Journal of Heat and Mass Transfer 77 (2014) 1231–1249

Page 13: International Journal of Heat and Mass Transferatsla.cau.ac.kr/wp-content/uploads/2018/05/J5.-Lee-et-al... · 2018-05-04 · 1232 H. Lee et al./International Journal of Heat and Mass

Boiling and Two-Phase Flow Laboratory (PU-BTPFL) have yielded adetailed model for CHF in both Earth gravity [71–76] and micro-gravity [77,78] with both subcooled and saturated inlet conditions.Using the mechanism of interfacial lift-off as a basis for CHF, Zhanget al. [78] developed the following relation for CHF based on heatedperimeter of the channel.

q00m ¼bFq

qgðhfg þ cp;f DTsub;oÞ4prd

qgbk2c

sinðpbÞ !"""""

z)

" #1=2

; ð13Þ

where b = 0.2, d is the mean thickness of the vapor layer formed alongthe heated walls of the channel. In Eq. (13), Fq is the fraction of thewall heat flux that is consumed in converting near-wall liquid tovapor (remaining fraction is consumed in overcoming liquid subcooling),

Fq ¼ 1&qf

qg

cp;f DTsub;o

hfg0:00285

qf U2Dh

r

!0:22

4

3

5: ð14Þ

Also in Eq. (13), kc is the critical wavelength of instability of theinterface between a vapor layer of thickness Hg (=d) and velocityUg, and liquid layer of thickness Hf and velocity Uf,

2pkc¼

q00f q00gðUg & Uf Þ2

2r q00f þ q00g# $ þ

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiq00f q00gðUg & Uf Þ2

2r q00f þ q00g# $

2

4

3

52

þðqf & qgÞgn

r

vuuut ; ð15Þ

where q00f ¼ qf cothð2pHf =kcÞ and q00g ¼ qgcothð2pHg=kcÞ and gn is thecomponent of gravity perpendicular to heated wall. A separated

flow model is used to determine the axial variations of Uf, Ug andd along the channel. Hf for a rectangular channel with three-sidedheating can be derived from Hf ¼ ½ð1& aÞHchWch(=½2fðHch & dÞþðWch & 2dÞg(, and q00m from Eqs. (13) and (14) based on the valuesof d and kc at z) ¼ z0 þ kcðz)Þ, where z0 is the location where thevapor layer velocity just exceeds the liquid layer velocity. Detailsof the separated flow model are provided in [75,78].

4.4. Negating influence of body force

Zhang et al. [72] derived two dimensionless parameters, Bo/We2

and 1/Fr, for macro-channel flows whose magnitude has to bemaintained below specific values to negate body force effects per-pendicular to, and parallel to the heated wall, respectively. The cri-teria developed by Zhang et al. are valid for subcooled inletconditions (xe,in = 0). More recently, Konishi et al. [76] extendedthese criteria to conditions involving a saturated liquid–vapor mix-ture at the inlet; i.e., for xe,in > 0. Using xe,in = 0 in the Konishi et al.criteria gives

BoWe2 ¼

ðqf & qgÞðqf þ qgÞ2rg

q2f q2

gðG=qf Þ4 6 0:232 ð16Þ

and

1jFrj¼ðqf & qgÞgDh

qf ðG=qf Þ2 6 0:02: ð17Þ

Fig. 7. Comparison of measured total pressure drop for horizontal flow, vertical upflow and vertical downflow with predictions of Homogeneous Equilibrium Model (HEM)using two-phase viscosity models of (a) McAdams et al. [46], (b) Akers et al. [47], (c) Cicchitti et al. [48], (d) Owens [49], (e) Dukler et al. [50], (f) Beattie and Whalley [51], and(g) Lin et al. [52].

H. Lee et al. / International Journal of Heat and Mass Transfer 77 (2014) 1231–1249 1243

Page 14: International Journal of Heat and Mass Transferatsla.cau.ac.kr/wp-content/uploads/2018/05/J5.-Lee-et-al... · 2018-05-04 · 1232 H. Lee et al./International Journal of Heat and Mass

Table 5Correlations for two-phase frictional pressure gradient.

Author(s) Correlation Remarks

Correlations not specifically developed for mini/micro-channel flowsLockhart and

Martinelli [54]dPdz

' (F ¼

dPdz

' (f /

2f

Adiabatic, circular, Dh = 1.49–25.83 mm, air–water, oils,hydrocarbons, horizontal

/2f ¼ 1þ C

X þ1

X2, X2 ¼ ðdP=dzÞfðdP=dzÞg

h i

Cvv ¼ 5, Ctv ¼ 10, Cvt ¼ 12, Ctt ¼ 20

Friedel [55] dPdz

' (F ¼

dPdz

' (fo/

2fo

Dh > 4 mm, R12, air–water, air–oil, horizontal, vertical upflow, 25,000data points

/2fo ¼ ð1& xeÞ2 þ x2

eqfqg

# $ffofgo

# $þ * * *

3:24x0:78e ð1& xeÞ0:224 qf

qg

# $0:91 lflg

# $0:191& lf

lg

# $0:7Fr&0:045

H We&0:035H

FrH ¼ G2

gDhq2H, WeH ¼ G2Dh

rqH, qH ¼ 1

xevgþð1&xe Þv f

Müller-Steinhagenand Heck [56]

dPdz

' (F ¼

dPdz

' (fo þ 2 dP

dz

' (go &

dPdz

' (fo

h ixe

n oð1& xeÞ1=3 þ dP

dz

' (gox3 Adiabatic, Dh = 4–392 mm, R11, R12, R22, air–water, steam–water,

oil, argon, hydrocarbons, neon, horizontal, vertical upflow/downflow,9300 data points

Jung andRadermacher[57]

dPdz

' (F ¼

dPdz

' (fo/

2fo , /fo ¼ 12:82X&1:47

tt ð1& xeÞ1:8 Circular stainless steel, Dh = 9.1 mm, R113, R12, R22, R152a,horizontal, more than 600 data points

Xtt ¼lflg

# $0:11&xe

xe

# $0:9 qgqf

# $0:5

Mishima andHibiki [58]

dPdz

' (F ¼

dPdz

' (f /

2f , /2

f ¼ 1þ CX þ

1X2

Circular Pyrex glass, adiabatic, Dh = 1.05–4.08 mm, air–water, verticalupflow, 299 data pointsfor rectangular channels, C ¼ 21½1& expð&319DhÞ(

for circular channels, C ¼ 21½1& expð&319DhÞ(Dh in mm

Yang and Webb[59]

dPdz

' (F ¼ &0:87Re0:12

eq ffoG2

eqv f

Dh

Adiabatic, multi-channel (N = 4), rectangular aluminum, Dh = 1.56,2.64 mm, R12, horizontal

Reeq ¼ Geq Dhlf

, Geq ¼ G ð1& xeÞ þ xeqfqg

# $0:5% &

Wang et al. [60] for G P 200 kg/m2 s, Adiabatic, circular Pyrex glass, Dh = 6.5 mm, R134a, R22, R407c,horizontaldP

dz

' (F ¼

dPdz

' (g/

2g , /2

g ¼ 1þ 9:4X0:62 þ 0:564X2:45

for G < 200 kg/m2 s,dPdz

' (F ¼

dPdz

' (f /

2f , /2

f ¼ 1þ CX þ

1X2 ,

C ¼ 4:566! 10&6X0:128Re0:938fo

v fvg

# $2:15 lflg

# $5:1

Yan and Lin [61] dPdz

' (F ¼ &0:22Re&0:1

eqG2

eqv f

Dh

Circular copper, Dh = 2.0 mm, R134a, horizontal

Tran et al. [62] dPdz

' (F ¼

dPdz

' (fo/

2fo/fo ¼ 1þ 4:3 ðdP=dzÞgo

ðdP=dzÞfo& 1

h iNconf x0:875

e ð1& xeÞ0:875 þ x1:75e

h iAdiabatic, circular/rectangular stainless steel, brass, Dh = 2.40, 2.46,2.92 mm, R113, R12, R134a, horizontal

Nconf ¼ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

rgðqf&qg ÞD

2h

q

Chen et al. [63] dPdz

' (F ¼

dPdz

' (fo;Friedel !X Adiabatic, circular copper, Dh = 1.02–9 mm, R410a, air–water,

horizontal, 886 data pointsfor Bd) < 2:5, X ¼ 0:0333Re0:45

fo

Re0:09g ð1þ0:4 expð&Bd) ÞÞ

for Bd) P 2:5, X ¼ We0:2H

ð2:5þ0:06Bd) Þ, Bd) ¼ gðqf&qg ÞðDh=2Þ2

r

Yu et al. [64]dPdz

' (F ¼

dPdz

' (f /

2f , /2

f ¼ 18:65 qgqf

# $0:51&xe

xe

# $Re0:1

g

Re0:5f

% &&1:9 Circular stainless steel, Dh = 2.98 mm, water, horizontal, 327 datapoints

Correlations specifically developed for mini/micro-channel flows

Lee and Lee [65] dPdz

' (F ¼

dPdz

' (f /

2f , /2

f ¼ 1þ CX þ

1X2

Adiabatic, rectangular acrylic, Dh = 1.02–9 mm, air–water, horizontal,305 data points

Cvv ¼ 6:833! 10&8k&1:317w0:719Re0:557fo , Ctv ¼ 3:627 Re0:174

fo

Cvt ¼ 6:185! 10&2Re0:726fo , Ctt ¼ 0:048Re0:451

fo

w ¼ lf jfr , k ¼

l2f

qf rDh

Hwang and Kim[66]

dPdz

' (F ¼

dPdz

' (f /

2f , /2

f ¼ 1þ CX þ

1X2

Adiabatic, circular stainless steel, Dh = 0.244, 0.430, 0.792 mm, R134a,horizontal

C ¼ 0:227Re0:452fo X&0:32N&0:82

conf

Sun and Mishima[67]

dPdz

' (F ¼

dPdz

' (f /

2f

Adiabatic/diabatic, circular/rectangular, Dh = 0.506–12 mm, R123,R134a, R22, R236ea, R245fa, R404a, R407C, R410a, R507, air–water,CO2, horizontal, vertical, 2092 data points from 18 sources

for Ref < 2000 and Reg < 2000,

/2f ¼ 1þ C

X þ1

X2, C ¼ 24 1þ Ref1000

# $1& exp &0:153

0:27Nconfþ0:8

# $h i

for Ref P 2000 and Reg P 2000,

/2f ¼ 1þ C

X1:19 þ 1X2 , C ¼ 1:79 Reg

Ref

# $0:41&xe

xe

# $0:5

Li and Wu [68] dPdz

' (F ¼

dPdz

' (f /

2f , /2

f ¼ 1þ CX þ

1X2

Adiabatic, circular/rectangular, Dh = 0.148–3.25 mm, R12, R134a, R22,R236ea, R245fa, R32, R404a, R410a, R422d, ammonia, propane,nitrogen, horizontal, 769 data pointsfor Bd 6 1:5, C ¼ 11:9Bd0:45

for 1:5 < Bd 6 11, C ¼ 109:4 BdRe0:5f

# $&0:56

for Bd > 11, HEM using Beattie and Whalley [51] mixture viscosity modelis recommended

1244 H. Lee et al. / International Journal of Heat and Mass Transfer 77 (2014) 1231–1249

Hyoungsoon Lee
Page 15: International Journal of Heat and Mass Transferatsla.cau.ac.kr/wp-content/uploads/2018/05/J5.-Lee-et-al... · 2018-05-04 · 1232 H. Lee et al./International Journal of Heat and Mass

Table 5 (continued)

Author(s) Correlation Remarks

Zhang et al. [69] dPdz

' (F ¼

dPdz

' (f /

2f , /2

f ¼ 1þ CX þ

1X2

Adiabatic/diabatic, circular/rectangular, Dh = 0.07–6.25 mm, R12,R113, R22, R134a, R404a, water, ammonia, air, N2, 2201 data pointsfrom 13 sources

C ¼ 21½1& expð&0:142=Nconf Þ(

Li and Wu [70] for Bd < 0:1, Adiabatic, circular/rectangular, Dh = 0.148–3.25 mm, R12, R134a, R22,R236ea, R245fa, R32, R404a, R410a, R422d, ammonia, propane,nitrogen, horizontal, 769 data points

dPdz

' (F ¼

dPdz

' (f /

2f , /2

f ¼ 1þ CX þ

1X2 , C ¼ 5:60Bd0:28

for 0:1 6 Bd and BdRe0:5f 6 200, dP

dz

' (F ¼

dPdz

' (fo/

2fo ,

/2fo ¼ ð1& xeÞ2 þ 2:87x2

e P&1R þ 1:54Bd0:19 qf&qg

qH

# $0:81

for 200 < BdRe0:5f , HEM using Beattie and Whalley [49] mixture viscosity

model is recommended

Kim and Mudawar[53]

dPdz

' (F ¼

dPdz

' (f /

2f , /2

f ¼ 1þ CX þ

1X2 , X2 ¼ ðdP=dzÞf

ðdP=dzÞgAdiabatic/diabatic, circular/rectangular, Dh = 0.349–5.35 mm, R12,R134a, R245fa, R410a, FC72, ammonia, CO2, water, horizontal,vertical, 2378 data points from 16 sources& dP

dz

' (f ¼

2f f v f G2ð1&xeÞ2

Dh, & dP

dz

' (g ¼

2f g vg G2 x2e

Dh

for Rek < 2000,f k ¼ 16Re&1

k for circularf kRek ¼ 24½1& 1:3553bþ 1:9467b2 & 1:7012b3 þ 0:9564b4 & 0:2537b5(

for rectangularfor 2000 6 Rek < 20;000, fk ¼ 0:079Re&0:25

k

for 20;000 6 Rek , fk ¼ 0:046Re&0:2k

Ref ¼Gð1&xÞDh

lf, Reg ¼ GxDh

lg

for Ref < 2000 and Reg < 2000 (vv)

C ¼ 3:5! 10&5Re0:44fo Su0:50

goqfqg

# $0:481þ 530We0:52

fo Bo PHPF

# $1:09% &

for Ref < 2000 and Reg P 2000 (vt)

C ¼ 0:0015Re0:59fo Su0:19

goqfqg

# $0:361þ 530We0:52

fo Bo PHPF

# $1:09% &

for Ref P 2000 and Reg < 2000 (tv)

C ¼ 8:7! 10&4Re0:17fo Su0:50

goqfqg

# $0:141þ 60We0:32

fo Bo PHPF

# $0:78% &

for Ref P 2000 and Reg P 2000 (tt)

C ¼ 0:39Re0:03fo Su0:10

goqfqg

# $0:351þ 60We0:32

fo Bo PHPF

# $0:78% &

Refo ¼ GDhlf

, Sugo ¼qgrDh

l2g

, Wefo ¼ G2Dhqf r

, Bo ¼ q00HGhfg

Fig. 8. Comparison of measured total pressure drop for horizontal flow, vertical upflow and vertical downflow with predictions of Separated Flow Models (SFMs) by (a)Lockhart and Martinelli [54], (b) Friedel [55], and (c) Müller-Steinhagen and Heck [56], (d) Jung and Radermacher [57], (e) Mishima and Hibiki [58], and (f) Yang and Webb[59].

H. Lee et al. / International Journal of Heat and Mass Transfer 77 (2014) 1231–1249 1245

Hyoungsoon Lee
Hyoungsoon Lee
Dh = 0.007~
Page 16: International Journal of Heat and Mass Transferatsla.cau.ac.kr/wp-content/uploads/2018/05/J5.-Lee-et-al... · 2018-05-04 · 1232 H. Lee et al./International Journal of Heat and Mass

Fig. 9. Comparison of measured total pressure drop for horizontal flow, vertical upflow and vertical downflow with predictions of Separated Flow Models (SFMs) by (a) Wanget al. [60], (b) Yan and Lin [61], (c) Tran et al. [62], (d) Chen et al. [63], and (e) Yu et al. [64].

1246 H. Lee et al. / International Journal of Heat and Mass Transfer 77 (2014) 1231–1249

Page 17: International Journal of Heat and Mass Transferatsla.cau.ac.kr/wp-content/uploads/2018/05/J5.-Lee-et-al... · 2018-05-04 · 1232 H. Lee et al./International Journal of Heat and Mass

Both criteria must be tested simultaneously to determine thecorresponding range for G/qf; the dominant criterion for specificoperating conditions is the one that yields the larger G/qf value.Eqs. (16) and (17) can also be used to determine G/qf for say Lunargravity (0.17ge) or Martian gravity (0.38ge). Using these two equa-tions for Earth gravity yields G/qf P 0.98 m/s and G/qf P 0.43 m/s,respectively. Therefore, gravitational effects for macro-channelflow can be eliminated with G/qf P 0.98 m/s.

As shown in Fig. 5g of part 1 of this study [25], orientationeffects are negated altogether for the present micro-channel flowwith G/qf P 0.22 m/s, which is significantly smaller than the valueof 0.98 m/s required for macro-channels. This demonstrates thesuperiority of micro-channels at negating body force effects forspace systems compared to macro-channels. In effect, the Bo/We2

and 1/Fr criteria given by Eqs. (16) and (17) provide a high upperlimit for the value of G/qf required to negate body force effects.

Therefore, using the predictive tools presented above, optimumdesign of a micro-channel evaporator for a space vehicle can beachieved through a trade study aimed at simultaneously (1) reduc-ing pressure drop, (2) maintaining heat flux safely below CHF, and(3) negating the influence of body force.

5. Conclusions

Using data for pressure drop and heat transfer coefficient forflow boiling in micro-channels at different orientations from part1 [25], the present part of a two-part study examined the accuracy

of published predictive tools. Also examined is the effectiveness ofmicro-channels at negating the influence of body force in reducedgravity space systems. Key findings from the study are as follows:

(1) The two-phase heat transfer coefficient data are comparedwith predictions of 15 popular correlations. These correla-tions are grouped into three distinct types: (i) correlationsbased on nucleate boiling (nb) or convective boiling (cb)relations, (ii) correlations based on maximum value pre-dicted using nb and cb relations, and (iii) correlations involv-ing superpositioning of nb and cb relations. Overall,correlations of the first type show poor predictions, withthe exception of Agostini and Bontemps’ [31], which yieldsMAEs of 17.2%, 16.1%, and 13.4%, for horizontal flow, verticalupflow, and vertical downflow, respectively. Correlations ofthe second type provide fair accuracy, but with appreciablescatter. And correlations of the third type generally providefair to good predictions, with Kim and Mudawar’s [42] yield-ing the best predictions, with MAEs of 19.0%, 19.9% and16.9% for horizontal flow, vertical upflow and vertical down-flow, respectively.

(2) The pressure drop data are compared with the predictions ofthe Homogeneous Equilibrium Model (HEM) in conjunctionwith 7 mixture viscosity relations, and 18 Separate FlowModels (SFMs). Using the HEM, the relation by Lin et al.[52] shows the best accuracy of all 7 mixture viscosityrelations, with MAEs of 22.0%, 28.9% and 22.8% for horizontal

Fig. 10. Comparison of measured total pressure drop for horizontal flow, vertical upflow and vertical downflow with predictions of mini/micro-channel Separated FlowModels (SFMs) by (a) Lee and Lee [65], (b) Hwang and Kim [66], (c) Sun and Mishima [67], (d) Li and Wu [68], (e) Zhang et al. [69], (f) Li and Wu [70], and (g) Kim andMudawar [53].

H. Lee et al. / International Journal of Heat and Mass Transfer 77 (2014) 1231–1249 1247

Hyoungsoon Lee
Hyoungsoon Lee
Hyoungsoon Lee
Hyoungsoon Lee
Page 18: International Journal of Heat and Mass Transferatsla.cau.ac.kr/wp-content/uploads/2018/05/J5.-Lee-et-al... · 2018-05-04 · 1232 H. Lee et al./International Journal of Heat and Mass

flow, vertical upflow and vertical downflow, respectively. Ofthe 18 SFM-based correlations, Kim and Mudawar’s [53]shows the best accuracy, with MAEs of 24.2%, 30.3% and24.1% for horizontal flow, vertical upflow and vertical down-flow, respectively.

(3) Three important criteria for implementing micro-channelflow boiling in space systems are proposed: (i) avoidinglarge pressure drop, (ii) avoiding critical heat flux (CHF),and (iii) negating the influence of body force. By requiringsigificantly smaller mass velocities to negate body forceeffects, it is shown that micro-channels are far more effec-tive for space applications than macro-channels.

Conflict of interest

None declared.

Acknowledgement

The authors are grateful for the support of the National Aero-nautics and Space Administration (NASA) – United States underGrant No. NNX13AC83G.

References

[1] I. Mudawar, Two-phase micro-channel heat sinks: theory, applications anlimitations, J. Electron. Pack. – Trans. ASME 133 (2011) 041002.

[2] D.B. Tuckerman, R.F.W. Pease, High-performance heat sinking for VLSI,Electron. Dev. Lett. – Trans. IEEE 2 (1981) 126–129.

[3] A.G. Fedorov, R. Viskanta, Three-dimensional conjugate heat transfer in themicrochannel heat sink for electronic packaging, Int. J. Heat Mass Transfer 43(2000) 399–415.

[4] W. Qu, I. Mudawar, Experimental and numerical study of pressure drop andheat transfer in a single-phase micro-channel heat sink, Int. J. Heat MassTransfer 45 (2000) 2549–2565.

[5] M.K. Sung, I. Mudawar, Experimental and numerical investigation of single-phase heat transfer using a hybrid jet-impingement/micro-channel coolingscheme, Int. J. Heat Mass Transfer 49 (2006) 682–694.

[6] M.K. Sung, I. Mudawar, Single-phase hybrid micro-channel/jet impingementcooling, Int. J. Heat Mass Transfer 51 (2008) 4342–4352.

[7] C.Y. Han, P. Griffith, The mechanism of heat transfer in nucleate pool boiling,Int. J. Heat Mass Transfer 8 (1965) 887–903.

[8] P.J. Marto, V.J. Lepere, Pool boiling heat transfer from enhanced surfaces todielectric fluids, J. Heat Transfer – Trans. ASME 104 (1982) 292–299.

[9] T.M. Anderson, I. Mudawar, Microelectronic cooling by enhanced pool boilingof a dielectric fluorocarbon liquid, J. Heat Transfer – Trans. ASME 111 (1989)752–759.

[10] I. Mudawar, A.H. Howard, C.O. Gersey, An analytical model for near-saturatedpool boiling CHF on vertical surfaces, Int. J. Heat Mass Transfer 40 (1997)2327–2339.

[11] S. Yilmaz, J.W. Westwater, Effect of velocity on heat transfer in boiling Freon-l13, J. Heat Transfer 102 (1) (1980) 26–31.

[12] Y. Katto, Critical heat flux, Adv. Heat Transfer 17 (1985) 1–64.[13] T.C. Willingham, I. Mudawar, Forced-convection boiling and critical heat flux

from a linear array of discrete heat sources, Int. J. Heat Mass Transfer 35 (1992)2879–2890.

[14] Y. Katto, M. Kunihiro, Study of the mechanism of burn-out in boiling system ofhigh burn-out heat flux, Bull. JSME 16 (1973) 1357–1366.

[15] M. Monde, T. Inoue, Critical heat flux in saturated forced convective boiling ona heated disk with multiple impinging jets, J. Heat Transfer – Trans. ASME 113(1991) 722–727.

[16] D.C. Wadsworth, I. Mudawar, Enhancement of single-phase heat transfer andcritical heat flux from an ultra-high-flux-source to a rectangular impinging jetof dielectric liquid, J. Heat Transfer – Trans. ASME 114 (1992) 764–768.

[17] M.E. Johns, I. Mudawar, An ultra-high power two-phase jet-impingementavionic clamshell module, J. Electron. Pack. – Trans. ASME 118 (1996) 264–270.

[18] S. Toda, A study in mist cooling (1st report: investigation of mist cooling),Trans. JSME 38 (1972) (1972) 581–588.

[19] L. Lin, R. Ponnappan, Heat transfer characteristics of spray cooling in a closedloop, Int. J. Heat Mass Transfer 46 (2003) 3737–3746.

[20] J.R. Rybicki, I. Mudawar, Single-phase and two-phase cooling characteristics ofupward-facing and downward-facing sprays, Int. J. Heat Mass Transfer 49(2006) 5–16.

[21] M. Visaria, I. Mudawar, Theoretical and experimental study of the effects ofspray orientation on two-phase spray cooling and critical heat flux, Int. J. HeatMass Transfer 51 (2008) 2398–2410.

[22] W. Nakayama, T. Nakajima, S. Hirasawa, Heat sink studs having enhancedboiling surfaces for cooling of microelectronic components, ASME Paper 84-WA/HT-89, 1984.

[23] R.L. Webb, The evolution of enhanced surface geometries for nucleate boiling,Heat Transfer Eng. 2 (1981) 46–69.

[24] V. Khanikar, I. Mudawar, T. Fisher, Effects of carbon nanotube coating on flowboiling in a micro-channel, Int. J. Heat Mass Transfer 52 (2009) 3805–3817.

[25] H. Lee, I. Park, I. Mudawar, M.M. Hasan, Micro-channel evaporator for spaceapplications – 1. Experimental pressure drop and heat transfer results fordifferent orientations in Earth gravity, Int. J. Heat Mass Transfer (2014), http://dx.doi.org/10.1016/j.ijheatmasstransfer.2014.06.012.

[26] G.M. Lazarek, S.H. Black, Evaporative heat transfer, pressure drop and criticalheat flux in a small vertical tube with R-113, Int. J. Heat Mass Transfer 25(1982) 945–960.

[27] M.G. Cooper, Saturation nucleate pool boiling – a simple correlation, Ind.Chem. Eng. Symp. Ser. 86 (1984) 785–793.

[28] T.N. Tran, M.W. Wambsganss, D.M. France, Small circular- and rectangularchannel boiling with two refrigerants, Int. J. Multiphase Flow 22 (1996) 485–498.

[29] H.J. Lee, S.Y. Lee, Heat transfer correlation for boiling flows in small rectangularhorizontal channels with low aspect ratios, Int. J. Multiphase Flow 27 (2001)2043–2062.

[30] G.R. Warrier, V.K. Dhir, L.A. Momoda, Heat transfer and pressure drop innarrow rectangular channels, Exp. Therm. Fluid Sci. 26 (2002) 53–64.

[31] B. Agostini, A. Bontemps, Vertical flow boiling of refrigerant R134a in smallchannels, Int. J. Heat Fluid Flow 26 (2005) 296–306.

[32] W. Li, Z. Wu, A general correlation for evaporative heat transfer in micro/minichannels, Int. J. Heat Mass Transfer 53 (2010) 1778–1787.

[33] H.K. Oh, C.H. Son, Flow boiling heat transfer and pressure drop characteristicsof CO2 in horizontal tube of 4.57-mm inner diameter, Appl. Therm. Eng. 31(2011) 163–172.

[34] M.M. Shah, A new correlation for heat transfer during boiling flow throughpipes, ASHRAE Trans. 82 (1976) 66–86.

[35] M.M. Shah, Chart correlation for saturated boiling heat transfer: equations andfurther study, ASHRAE Trans. 88 (1982) 185–196.

[36] M. Ducoulombier, S. Colasson, J. Bonjour, P. Haberschill, Carbon dioxide flowboiling in a single microchannel – Part II: Heat transfer, Exp. Therm. Fluid Sci.35 (2011) 597–611.

[37] J.C. Chen, Correlation for boiling heat transfer to saturated fluids in convectiveflow, Ind. Eng. Chem. Process Des. Dev. 5 (1966) 322–329.

[38] S. Edelstein, A.J. Perez, J.C. Chen, Analytic representation of convective boilingfunctions, AIChE J. 30 (1984) 840–841.

[39] K.E. Gungor, R.H.S. Winterton, A general correlation for flow boiling in tubesand annuli, Int. J. Heat Mass Transfer 29 (1986) 351–358.

[40] Z. Liu, R.H.S. Winterton, A general correlation for saturated and subcooled flowboiling in tubes and annuli, based on a nucleate pool boiling equation, Int. J.Heat Mass Transfer 34 (1991) 2759–2766.

[41] S.S. Bertsch, E.A. Groll, S.V. Garimella, A composite heat transfer correlation forsaturated flow boiling in small channels, Int. J. Heat Mass Transfer 52 (2009)2110–2118.

[42] S.M. Kim, I. Mudawar, Universal approach to predicting saturated flow boilingheat transfer in mini/micro-channels – Part II. Two-phase heat transfercoefficient, Int. J. Heat Mass Transfer 64 (2013) 1239–1256.

[43] R.K. Shah, A.L. London, Laminar flow forced convection in ducts: a source bookfor compact heat exchanger analytical data – Suppl. 1, Academic Press, NewYork, 1978.

[44] S.M. Kim, I. Mudawar, Consolidated method to predicting pressure drop andheat transfer coefficient for both subcooled and saturated flow boiling inmicro-channel heat sinks, Int. J. Heat Mass Transfer 55 (2012) 3720–3731.

[45] S.M. Zivi, Estimation of steady-state steam void-fraction by means of theprinciple of minimum entropy production, J. Heat Transfer – Trans. ASME 86(1964) 247–252.

[46] W.H. McAdams, W.K. Woods, L.C. Heroman, Vaporization inside horizontaltubes, II. Benzene–oil mixture, Trans. ASME 64 (1942) 193–200.

[47] W.W. Akers, H.A. Deans, O.K. Crosser, Condensing heat transfer withinhorizontal tubes, Chem. Eng. Prog. 54 (1958) 89–90.

[48] A. Cicchitti, C. Lombardi, M. Silvestri, G. Soldaini, R. Zavalluilli, Two-phasecooling experiments-pressure drop, heat transfer and burnout measurements,Energia Nucl. 7 (1960) 407–425.

[49] W.L. Owens, Two-phase pressure gradient, in: Int. Dev. Heat Transfer, Part II,ASME, New York, 1961.

[50] A.E. Dukler, M. Wicks, R.G. Cleaveland, Pressure drop and hold up in two-phaseflow, AIChE J. 10 (1964) 38–51.

[51] D.R.H. Beattie, P.B. Whalley, A simple two-phase frictional pressure dropcalculation method, Int. J. Multiphase Flow 8 (1982) 83–87.

[52] S. Lin, C.C.K. Kwok, R.Y. Li, Z.H. Chen, Z.Y. Chen, Local frictional pressure dropduring vaporization of R-12 through capillary tubes, Int. J. Multiphase Flow 17(1991) 95–102.

[53] S.M. Kim, I. Mudawar, Universal approach to predicting two-phase frictionalpressure drop for mini/micro-channel saturated flow boiling, Int. J. Heat MassTransfer 58 (2013) 718–734.

[54] R.W. Lockhart, R.C. Martinelli, Proposed correlation of data for isothermal two-phase, two-component flow in pipes, Chem. Eng. Prog. 45 (1949) 39–48.

[55] L. Friedel, Improved friction pressure drop correlations for horizontal andvertical two-phase pipe flow, in: European Two-Phase Group Meeting, Ispra,Italy, 1979, Paper E2.

1248 H. Lee et al. / International Journal of Heat and Mass Transfer 77 (2014) 1231–1249

Hyoungsoon Lee
Page 19: International Journal of Heat and Mass Transferatsla.cau.ac.kr/wp-content/uploads/2018/05/J5.-Lee-et-al... · 2018-05-04 · 1232 H. Lee et al./International Journal of Heat and Mass

[56] H. Müller-Steinhagen, K. Heck, A simple friction pressure drop correlation fortwo-phase flow in pipes, Chem. Eng. Process 20 (1986) 297–308.

[57] D.S. Jung, R. Radermacher, Prediction of pressure drop during horizontalannular flow boiling of pure and mixed refrigerants, Int. J. Heat Mass Transfer32 (1989) 2435–2446.

[58] K. Mishima, T. Hibiki, Some characteristics of air–water two-phase flow insmall diameter vertical tubes, Int. J. Multiphase Flow 22 (1996) 703–712.

[59] C.Y. Yang, R.L. Webb, Friction pressure drop of R-12 in small hydraulicdiameter extruded aluminum tubes with and without micro-fins, Int. J. HeatMass Transfer 39 (1996) 801–809.

[60] C.C. Wang, C.S. Chiang, D.C. Lu, Visual observation of two-phase flow pattern ofR-22, R-134a, and R-407C in a 6.5-mm smooth tube, Exp. Therm. Fluid Sci. 15(1997) 395–405.

[61] Y.Y. Yan, T.F. Lin, Evaporation heat transfer and pressure drop of refrigerant R-134a in a small pipe, Int. J. Heat Mass Transfer 41 (1998) 4183–4194.

[62] T.N. Tran, M.C. Chyu, M.W. Wambsganss, D.M. France, Two-phase pressuredrop of refrigerants during flow boiling in small channels: an experimentalinvestigation and correlation development, Int. J. Multiphase Flow 26 (2000)1739–1754.

[63] I.Y. Chen, K.S. Yang, Y.J. Chang, C.C. Wang, Two-phase pressure drop of air–water and R-410A in small horizontal tubes, Int. J. Multiphase Flow 27 (2001)1293–1299.

[64] W. Yu, D.M. France, M.W. Wambsganss, J.R. Hull, Two-phase pressure drop,boiling heat transfer, and critical heat flux to water in a small-diameterhorizontal tube, Int. J. Multiphase Flow 28 (2002) 927–941.

[65] H.J. Lee, S.Y. Lee, Pressure drop correlations for two-phase flow withinhorizontal rectangular channels with small heights, Int. J. Multiphase Flow 27(2001) 783–796.

[66] Y.W. Hwang, M.S. Kim, The pressure drop in microtubes and the correlationdevelopment, Int. J. Heat Mass Transfer 49 (2006) 1804–1812.

[67] L. Sun, K. Mishima, Evaluation analysis of prediction methods for two-phaseflow pressure drop in mini-channels, Int. J. Multiphase Flow 35 (2009) 47–54.

[68] W. Li, Z. Wu, A general correlation for adiabatic two-phase pressure drop inmicro/mini-channels, Int. J. Heat Mass Transfer 53 (2010) 2732–2739.

[69] W. Zhang, T. Hibiki, K. Mishima, Correlations of two-phase frictional pressuredrop and void fraction in mini-channel, Int. J. Heat Mass Transfer 53 (2010)453–465.

[70] W. Li, Z. Wu, Generalized adiabatic pressure drop correlations in evaporativemicro/mini-channels, Exp. Therm. Fluid Sci. 35 (2011) 866–872.

[71] H. Zhang, I. Mudawar, M.M. Hasan, Experimental and theoretical study oforientation effects on flow boiling CHF, Int. J. Heat Mass Transfer 45 (2002)4463–4477.

[72] H. Zhang, I. Mudawar, M.M. Hasan, A method for assessing the importance of bodyforce on flow boiling CHF, J. Heat Transfer – Trans. ASME 126 (2004) 161–168.

[73] C.R. Kharangate, I. Mudawar, M.H. Hasan, Experimental and theoretical studyof critical heat flux in vertical upflow with inlet vapor void, Int. J. Heat MassTransfer 55 (2012) 360–374.

[74] C.R. Kharangate, I. Mudawar, M.H. Hasan, Photographic study and modeling ofcritical heat flux in horizontal flow boiling with inlet vapor void, Int. J. HeatMass Transfer 55 (2012) 4154–4168.

[75] C. Konishi, I. Mudawar, M.H. Hasan, Investigation of the influence oforientation on critical heat flux for flow boiling with two-phase inlet, Int. J.Heat Mass Transfer 61 (2013) 176–190.

[76] C. Konishi, I. Mudawar, M.H. Hasan, Criteria for negating the influence ofgravity on flow boiling critical heat flux with two-phase inlet conditions, Int. J.Heat Mass Transfer 65 (2013) 203–218.

[77] H. Zhang, I. Mudawar, M.M. Hasan, Flow boiling CHF in microgravity, Int. J.Heat Mass Transfer 48 (2005) 3107–3118.

[78] H. Zhang, I. Mudawar, M.M. Hasan, CHF model for subcooled flow boiling inEarth gravity and microgravity, Int. J. Heat Mass Transfer 50 (2007) 4039–4051.

H. Lee et al. / International Journal of Heat and Mass Transfer 77 (2014) 1231–1249 1249