A Critical Review of Pool and Flow Boiling Heat Transfer of Dielectric Fluids On Enhanced Surfaces
A Critical Review of Pool and Flow Boiling Heat Transfer of Dielectric Fluids On Enhanced Surfaces
A Critical Review of Pool and Flow Boiling Heat Transfer of Dielectric Fluids On Enhanced Surfaces
sg)
Nanyang Technological University, Singapore.
2016
Leong, K. C., Ho, J. Y., & Wong, K. K. (2017). A critical review of pool and flow boiling heat
transfer of dielectric fluids on enhanced surfaces. Applied Thermal Engineering, 112,
999‑1019. doi:10.1016/j.applthermaleng.2016.10.138
https://hdl.handle.net/10356/144567
https://doi.org/10.1016/j.applthermaleng.2016.10.138
© 2016 Elsevier Ltd. All rights reserved. This paper was published in Applied Thermal
Engineering and is made available with permission of Elsevier Ltd.
A critical review of pool and flow boiling heat transfer of dielectric fluids on
enhanced surfaces
K.C. Leong∗, J.Y. Ho, K.K. Wong
Singapore Centre for 3D Printing, School of Mechanical and Aerospace Engineering, Nanyang
Technological University, 50 Nanyang Avenue, Singapore 639798, Republic of Singapore
ABSTRACT
Pool and flow boiling of dielectric fluids are efficient direct liquid cooling techniques with immense
potential in the thermal management of electronic/electrical components. However, due to the
increasing demand for higher rate of heat flux dissipation, many enhanced surfaces have been developed
to further augment the boiling performance of dielectric fluids. This has resulted in large collections of
experimental data and predictive models being reported in the recent years. The present review seeks
to consolidate and highlight the recent developments in pool and flow boiling of dielectric fluids with
enhanced surfaces. The various models developed to characterize the nucleate boiling curves and to
predict critical heat fluxes of plain and enhanced surface are critically examined. The effects of
enhanced surfaces on the bubble dynamics in dielectric fluids are also examined and the fundamental
studies on boiling found in the recent literature are provided. In addition, attempts are made to
categorize the various enhanced surfaces based on their fabrication techniques and their heat transfer
performances and to elucidate the thermal transport mechanisms involved. Based on the literature
surveyed, the various experimental results are compared, existing shortfalls are identified and areas
which require further investigations are proposed.
KEYWORDS: Heat transfer; pool boiling; flow boiling; dielectric fluids; enhanced surfaces
1. Introduction
The advent in nanotechnology along with the continuous miniaturization of electronics and
increasing transistor packing densities have resulted in the exponential increase in chip-level
heat flux. It was projected that heat dissipated from the microprocessor could reach 200 W/cm2
to 300 W/cm2 and in some cases, on-chip hot spot heat fluxes as high as 1.5 kW/cm2 were
reported [1]. Due to this increase, efficient thermal management solutions are crucial to
maintain the electronic devices within the operating temperature limits. Pool and flow boiling
are direct liquid cooling techniques which have demonstrated significantly higher heat transfer
performance as compared to single-phase cooling schemes [2, 3]. These cooling techniques
involve phase change in the thermal transport process and reduce the interfacial resistance
otherwise present in indirect cooling. Due to their immense cooling potential without the need
for elaborate operating facilities, pool and flow boiling have been employed in the cooling of
high heat flux electronic and electrical devices such as high performance electronic computers
[4], batteries in electric vehicles [5] and avionics components [6]. As the coolant fluids are in
direct contact with the electronic heat sources, dielectric fluids such as fluorocarbon fluids (FC-
72, FC-87 and PF-5060) and hydrofluoroethers (HFE-7000, HFE-7100, HFE-7300) are often
the fluids of choice due to their chemical compatibility, electrical inertness and high dielectric
strength. In addition, due to their low boiling points, dielectric fluids are also highly desirable
for boiling heat transfer applications to preserve the die temperatures within the recommended
range of 80ºC to 130ºC [1]. However, the low surface tension and high-wetting nature of these
∗ Corresponding author
E-mail address: [email protected] (K.C. Leong)
1
Pre-publication version
K.C. Leong et al. Applied Thermal Engineering 112 (2017) 999–1019
http://dx.doi.org/10.1016/j.applthermaleng.2016.10.138
fluids which enable the fluids to penetrate deep into the cavity sites, have resulted in noticeably
higher incipient superheats as compared to conventional fluids such as water [7]. As shown in
Table 1, their thermophysical properties such as latent heat of vaporization and liquid specific
heat are also poorer than those of water.
Nomenclature
2
Pre-publication version
K.C. Leong et al. Applied Thermal Engineering 112 (2017) 999–1019
http://dx.doi.org/10.1016/j.applthermaleng.2016.10.138
To overcome these drawbacks, many surface modification techniques with the aim of
enhancing boiling heat transfer with dielectric fluids have been developed and extensively
investigated. Direct coating methods such as electrochemical deposition, chemical vapor
deposition and direct powder sintering have been studied and remarkable enhancements in
boiling heat transfer coefficients and critical heat fluxes were reported. On the other hand,
intrinsic surface features developed using MEMS/NEMS technologies, CNC machining,
additive manufacturing (AM) and other advanced manufacturing techniques have also
achieved significant boiling heat transfer augmentation. Recently, there are several review
articles on boiling heat transfer. Ciloglu and Bolukbasi [8] presented the potential heat transfer
applications of nanofluids and provided a detailed survey of existing pool boiling results with
nanofluids. Subsequently, a review on flow boiling heat transfer with nanofluids was
undertaken by Fang et al. [9]. On the other hand, various porous surface fabrication techniques
for enhancing pool boiling were reviewed by Patil and Kandlikar [10] and more recently,
Shojaeian and Kosar [11] and Kim et al. [12] presented comprehensive data of pool and flow
boiling with micro- and nanostructured surfaces. References [10 – 12] focus mainly on the
surface modification techniques for enhancing boiling.
In this paper, recent developments in augmenting pool and flow boiling heat transfer of
dielectric fluids with enhanced surfaces are critically reviewed. The various methods employed
to fabricate these surfaces and their levels of enhancements are elaborated. The effects of
enhanced surfaces on the bubble dynamics in dielectric fluids are examined and the
fundamental studies on boiling found in recent literature are also provided. Based on the
literature surveyed, existing shortfalls and areas which require further investigations are
identified.
2. Pool boiling
2.1 Overview
The Nukiyama boiling curve [14] (Fig. 1) which relates the heat flux dissipated from the heated
surface to its wall superheat is typically observed in most boiling processes and can be
classified into four distinct regimes, namely, natural convection, nucleate boiling, transition
boiling and film boiling. For practical applications, it is ideal to operate within the nucleate
boiling regime to capitalize on the high heat removal rate at low surface temperature. In
addition, the ability to predict the critical heat flux (CHF), which denotes the departure from
nucleate boiling regime, is also essential to prevent system burnout.
The classical boiling heat transfer model proposed by Rohsenow [15] and shown in Eq. (1)
assumes that the heat removed from the surface is primarily due to the fluid motion induced by
the departing bubbles or bubble microconvection. Hence, the single-phase force convection
correlation was adopted where the bubble departure diameter and velocity were used as the
characteristic length and velocity in the correlation. However, the effects of phase change were
not considered.
3
Pre-publication version
K.C. Leong et al. Applied Thermal Engineering 112 (2017) 999–1019
http://dx.doi.org/10.1016/j.applthermaleng.2016.10.138
In a separate model, Mikic and Rohsenow [16] assumed that the departing bubble resulted in
the removal of localized superheated liquid layer allowing cooler bulk fluid to replenish and
cool the surface. Commonly known as the transient conduction model, this mechanism-based
correlation shown in Eq. (2) includes the effects of the boiling surface active nucleation site
density but does not consider interactions between neighboring bubbles.
𝑞𝑞𝑡𝑡 𝑞𝑞 𝐴𝐴𝑛𝑛𝑛𝑛 𝑞𝑞
=� � + � � (2a)
𝐴𝐴𝑡𝑡 𝐴𝐴 𝑏𝑏 𝐴𝐴𝑡𝑡 𝐴𝐴 𝑛𝑛𝑛𝑛
𝑞𝑞 𝐾𝐾 2
� � = �𝜋𝜋𝑘𝑘𝑙𝑙 ρ𝑙𝑙 𝐶𝐶𝑝𝑝𝑝𝑝 𝑓𝑓𝐷𝐷𝑏𝑏2 𝑁𝑁(𝑇𝑇𝑤𝑤 − 𝑇𝑇∞ ) (2b)
𝐴𝐴 𝑏𝑏 2
1/4
𝑞𝑞 γ𝑔𝑔(𝑇𝑇𝑤𝑤 − 𝑇𝑇∞ ) 5 α3 (2c)
� � = 0.54ρ𝑙𝑙 𝐶𝐶𝑝𝑝𝑝𝑝 � �
𝐴𝐴 𝑛𝑛𝑛𝑛 �𝐴𝐴𝑡𝑡𝑡𝑡𝑡𝑡𝑡𝑡𝑡𝑡 ν
1/3 (2d)
𝑞𝑞 γ𝑔𝑔(𝑇𝑇𝑤𝑤 − 𝑇𝑇∞ ) 4 α2
� � = 0.14ρ𝑙𝑙 𝐶𝐶𝑝𝑝𝑝𝑝 � �
𝐴𝐴 𝑛𝑛𝑛𝑛 ν
Cooper and Lloyd [17] measured the surface temperatures of glass and ceramic substrates
during the bubble ebullition cycle and recorded a rapid decline in temperature during the initial
growth period of the vapor bubble. The temperature then quickly recovered following the
formation of dry spot. This observation was consistent with the hypotheses of Snyder and
Edward [18] and Moore and Mesler [19] and suggested that the dominant heat transfer
mechanism during the ebullition cycle was the evaporation of the fluid microlayer beneath the
growing bubble. To account for the effects of microlayer evaporation, Judd and Hwang [20]
included an additional term shown in Eq. (3) to the right hand side of Eq. (2). Benjamin and
Balakrishnan [21] subsequently derived an expression which relates the volume of the
microlayer evaporation (𝑉𝑉�𝑀𝑀𝑀𝑀 ) to the bubble growth time.
𝑞𝑞
� � = ρ𝑙𝑙 ℎ𝑓𝑓𝑓𝑓 𝑁𝑁𝑁𝑁𝑉𝑉�𝑀𝑀𝑀𝑀 (3)
𝐴𝐴 𝑀𝑀𝑀𝑀
On the other hand, Wayner et al. [22] showed that the significant heat transfer rate during the
ebullition cycle was due to the evaporation of the thin liquid film at the three-phase contact
line. Fig. 2 shows the liquid meniscus at the base of the bubble where the three-phase contact
line exists. In the absorbed film region, the liquid film is of nanoscale thickness and cannot
evaporate due to molecular adhesion force whereas in the micro region, maximum evaporation
and heat transfer rates take place by heat conduction across the thin liquid film. Stephan and
Hammer [23] included the effects of capillary pressure and interface curvature at the micro
region and developed a model to predict the nucleate boiling heat transfer coefficient. From
their numerical results, it was shown that the percentage of heat transfer from the micro region
increases with increasing wall superheat and the maximum heat flux in the micro region was
about 100 times larger than the burnout heat flux. Subsequently, Kern and Stephan [24]
extended the model to binary fluid mixtures.
4
Pre-publication version
K.C. Leong et al. Applied Thermal Engineering 112 (2017) 999–1019
http://dx.doi.org/10.1016/j.applthermaleng.2016.10.138
Building on the above earlier concepts, more comprehensive analytical and numerical boiling
heat transfer models have subsequently been developed. For instance, Zhao et al. [25]
suggested that the individual bubble growth can be divided into two distinct growth periods,
viz., initial and final growth periods. In the initial growth period, the bubble grows from the
evaporation of the microlayer whereas in the final growth period, the macrolayer (liquid layer
thicker than the microlayer) forms underneath the bubble and between adjacent bubbles.
Consequently, Das et al. [26] developed a mechanistic model which incorporated the
mechanisms of near-wall micro-/macrolayer evaporation and transient conduction within the
liquid layer. The model adopted several empirical constants but showed good agreement with
experimental data obtained from the literature. More recently, a semi-analytical boiling heat
transfer model was developed by Li et al. [27] for hydrophilic surfaces in which the influences
of contact angle on the microlayer dynamics, bubble departure diameter and nucleation site
density were considered. For hydrophilic surfaces with solid-liquid contact angles from 0° to
50°, the model showed reasonably good agreement with the experimental data (within ±30%
error).
In order to investigate the nucleate boiling process, several numerical methods have also been
introduced. Son et al. [28, 29] successfully implemented the level set (LS) method to simulate
a single bubble growth process in which the LS formulation was coupled with the phase change
equations to track the bubble liquid-vapor interface. On the other hand, Kunkelmann and
Stephan [30] numerically studied the boiling of HFE-7100 using the Volume-of-Fluid (VOF)
method. The effects of microscale heat transfer at the three-phase contact line was incorporated
in the model and it was reported that the simulated bubble growth time and departure diameter
were in good agreement with experimental observations. Recently, a modified VOF model was
also developed by Jia et al. [31] through the introduction of an improved Height Function
algorithm. Apart from the LS and VOF methods, the lattice Boltzmann method (LBM) provides
another approach to solve the complex interfacial dynamics in the phase change process. LBM
is a statistical mesoscopic method based on the kinetic theory where the system behavior is
described by a distribution function. Gong and Cheng [32] proposed a LBM model for
multiphase flow where the equation of state for real gas, which determined the system phase
change, was incorporated in the body force term of the evolution equation. The ebullition cycle
of a single bubble was simulated and results showed good agreement with existing correlations.
The model was subsequently implemented to investigate boiling heat transfer on smooth
surfaces with mixed wettability [33].
5
Pre-publication version
K.C. Leong et al. Applied Thermal Engineering 112 (2017) 999–1019
http://dx.doi.org/10.1016/j.applthermaleng.2016.10.138
mechanism for bubble heat transfer. Subsequently, Myers et al. [35] extended the work to
measurements of bubble nucleation under constant wall heat flux and determined that
microlayer evaporation and contact line heat transfer account for not more than 23% of the
total heat transfer from the surface. Wagner and Stephan [36], on the other hand, measured the
evolution of a bubble from a single nucleation site using IR thermography. At 500 mbar, the
′′
ratio of the micro region heat flux (𝑞𝑞𝑚𝑚𝑚𝑚𝑚𝑚,𝑚𝑚 ) to the total bubble evaporation heat flux (𝑞𝑞𝐵𝐵′′ ) was
30% for FC-84 and was reduced to 13.2% for FC-84/FC-3284 binary mixtures. More recently,
Moghaddam and Kiger [37, 38] investigated saturated boiling of FC-72 of a single bubble from
an artificial cavity and obtained temperature measurements with spatial resolutions of 22 to 40
µm. From the temperatures recorded and the high speed images of the bubble evolution (Fig.
4), it was identified that the transient conduction mechanism was mainly limited to the bubble-
to-surface contact area whereas microconvection was localized outside the contact area. In
addition, heat transfer from microlayer evaporation only constituted between 16.3% and 28.8%
of the total heat transfer.
Based on the published research highlighted above, it has been relatively well accepted that
transient conduction and/or microconvection are the dominant mechanisms in the heat transfer
of isolated bubbles. However, under high heat fluxes, significantly larger number of bubbles
can be generated over the boiling surface which would result in interaction and coalescence of
neighboring bubbles. Bubble interaction may alter the bubble dynamics such as bubble growth
rate and departure frequency which may in turn affect the associated mode of heat transfer.
Using a microheater array, Chen and Chung [39] generated two bubbles growing adjacent to
each other and at close proximity such that the bubbles were allowed to touch and coalesce
during the ebullition cycle. Their measurements indicated that bubble coalescence have
increased the bubble departure frequency and also substantially enhanced heat transfer (up to
70%). On the other hand, from the observation of the high speed images taken (Fig. 5), Seidel
et al. [40] reasoned that the occurrence of pressure imbalance at the instance of bubble
coalescence created shock waves which propagated along two fronts. It was suggested that
these capillary induced waves, instead of vapor production during bubble merger, resulted in
the removal of the macrolayer beneath the coalesced bubbles. More recently, Bi et al. [41]
studied the effects of coalescence of (1) two bubbles of identical sizes and (2) two bubbles of
different sizes in FC-72. It was observed that the heat flux enhancement due to bubble
coalescence was significantly larger than bubble growth and accounted for approximately 90%
of the total increase in heat flux. Bubble coalescence also increased the bubble departure
frequencies by more than two times as compared to single bubble growth and hence, resulted
in the increase in average heat flux.
Apart from plain surfaces, the need for higher heat flux removal have also resulted in the
introduction of numerous enhanced boiling surfaces such as surfaces with artificial fins,
cavities and grooves. The various surface modification techniques and the associated heat
transfer characteristics of these surfaces will be discussed in detail in Section 2.2. On the
visualization of bubble dynamics, the presence of surface features have been shown to alter the
bubble behavior and increased the complexity of transport mechanisms involved. For instance,
artificial surface pores of various pore diameters (0.12 to 0.2 mm) with interconnected network
channels were fabricated by Ramaswamy et al. [42]. In their bubble visualization studies in
FC-72, for each surface of the same pore size and pore separation, data of approximately 50
bubbles were collected, from which the average bubble departure diameters (Db) were
computed. It was determined that Db increases whereas the bubble departure frequency
decreases with increasing pore diameter. In addition, the bubble growth rates were seen to
increase linearly with time for the entire growth period, suggesting that the bubble growth was
6
Pre-publication version
K.C. Leong et al. Applied Thermal Engineering 112 (2017) 999–1019
http://dx.doi.org/10.1016/j.applthermaleng.2016.10.138
inertia driven. Subsequent boiling visualization by Ghiu and Joshi [43] of similar enhanced
surfaces made from transparent material (quartz) further determined that the evaporation
process within the interconnected channels also contributed significantly to the overall heat
dissipated from the surface. More recently, Teodori et al. [44] combined high speed
visualization with Particle Imaging Velocimetry (PIV) and substrate surface temperature
measurements to determine the effects of artificial cavity spacing on boiling. Enhanced
surfaces with arrays of micro-sized square cavities (52 µm width and 20 µm depth) with
different spacings were fabricated and tested in water, ethanol and HFE-7100. From their
measurements, the bubble departure frequency was observed to increase with decreasing cavity
spacing. In addition, for boiling from surfaces with closely spaced cavities in HFE-7000, stable
increase in the heat transfer coefficient was observed with increasing bubble characteristics
velocities. However, for surfaces with micro-cavities, horizontal coalescence of bubbles at high
heat flux were suggested to have deteriorated heat transfer coefficients as the bubbles formed
a vapor layer which insulated the surface and prevented liquid replenishments. Subsequent
studies suggested that the horizontal bubble coalescence can be inhibited with the use of surface
enhanced micro-sized pillars which delay the formation of the vapor blanket [45].
Despite coexisting on the continuous boiling curve as shown in Fig. 1, nucleate boiling heat
transfer and CHF are conventionally viewed as disjoint processes dominated by different
mechanisms. One of the earliest CHF model which was developed by Zuber [46] assumes that
CHF occurs when the columns of vapor jets escaping from the boiling surface and the resupply
fluid in the counterflow direction reached a hydrodynamically unstable state that inhibited
further vapor outflow. Based on the Taylor and Helmholtz instabilities, Eq. (4) which relates
CHF to the fluid thermophysical properties was derived where the constant, C, was taken to be
0.131. On the other hand, Gaertner [47] conducted visualization studies on the boiling
phenomenon and reported the existence of a macrolayer, comprising columns of vapor stems
and a liquid film between the heated surface and a hovering vapor mushroom at high heat flux.
Assuming that the thickness of the macrolayer was related to Helmholtz instabilities at the
interface of the vapor-liquid stems and that CHF occurs when the macrolayer dries out before
the departure of the vapor mushroom, Haramura and Katto [48] developed a hydrodynamic
model which relates CHF to the vapor stem-to-heater surface area ratio as shown in Eq. (5).
The hydrodynamic models of Zuber [46] and Haramura and Katto [48] provide reasonably
accurate predictions for a wide range of applications. Furthermore, Eq. (4) also allows the
flexibility for C to be varied so as to correlate with the different boiling conditions. For instance,
El-Genk and Parker [49] and Ho et al. [50] proposed C as a function of the surface orientation
and correlated with the experimental CHF values of enhanced surfaces in dielectric fluids.
′′ 1/2 1/4
𝑞𝑞𝐶𝐶𝐶𝐶𝐶𝐶 = 𝐶𝐶ρ𝑔𝑔 ℎ𝑓𝑓𝑓𝑓 �𝑔𝑔σ�ρ𝑙𝑙 − ρ𝑔𝑔 �� (4)
5
5 ⎡ ρ ⎤
16
5 � 𝑙𝑙 + 1�
𝐴𝐴𝑣𝑣 8 𝐴𝐴𝑣𝑣 16 ⎢ ρ𝑔𝑔
⎥ 1/4
′′
𝑞𝑞𝐶𝐶𝐶𝐶𝐶𝐶 = 0.72 � � �1 − � ⎢ �ρ𝑔𝑔 ℎ𝑓𝑓𝑓𝑓 �σg(ρ𝑙𝑙 − ρ𝑔𝑔 ) � � (5)
𝐴𝐴𝑤𝑤 𝐴𝐴𝑤𝑤 3/5 ⎥
⎢�11 ρ𝑙𝑙 + 1� ⎥
⎣ 16 ρ𝑔𝑔 ⎦
However, the hydrodynamic models were also commonly criticized for not having explicit
expressions to account for effects such as surface topology, surface wettability and material
thermal conductivity. This led to several improved models being proposed in recent years. Arik
7
Pre-publication version
K.C. Leong et al. Applied Thermal Engineering 112 (2017) 999–1019
http://dx.doi.org/10.1016/j.applthermaleng.2016.10.138
and Bar-Cohen [51] suggested that boiling surface properties such as heater thickness and
thermal conductivity have direct influence on CHF and can be characterized using the thermal
effusivity relation. Under high heat flux, surfaces of high effusivity facilitate the temperature
rise at localized dry spots and caused heat to be conducted to regions of the surface where
nucleate boiling still prevails. This prevents global dryout and delays CHF. The effusivity-
based CHF correlation for dielectric fluids was developed, as shown in Eq. (6), and provided
accurate predictions for a range of surfaces in dielectric fluids. On the other hand, the effects
of surface-liquid interactions were considered in the CHF model proposed by Kandlikar [52]
through the introduction of the contact angle (β) as shown in Eq. (7). The model suggests that
CHF was caused by the uncontrolled lateral expansion of the bubble which eventually
blanketed the heater surface and was derived by balancing the vapor recoil forces against the
counteracting surface tension and gravitational forces acting on the bubble. The model was
validated with fluids of different contact angles and surface orientations with reasonably
accurate predictions.
′′
𝜋𝜋 1/2 1/4 𝑆𝑆
𝑞𝑞𝐶𝐶𝐶𝐶𝐶𝐶 = ℎ𝑓𝑓𝑓𝑓 ρ𝑔𝑔 �σg(ρ𝑙𝑙 − ρ𝑔𝑔 ) � � � × �1 + 〈0.3014 − 0.01507𝐿𝐿′(𝑃𝑃)〉�
24 𝑆𝑆 + 0.1
ρ𝑙𝑙
0.75
C𝑝𝑝𝑝𝑝 (6)
× �1 + 0.030 �� � � ∆𝑇𝑇𝑠𝑠𝑠𝑠𝑠𝑠 �
ρ𝑔𝑔 h𝑓𝑓𝑓𝑓
1/2
′′ 1/2 1 + cosβ 2 𝜋𝜋
(7)
𝑞𝑞𝐶𝐶𝐶𝐶𝐶𝐶 = ℎ𝑓𝑓𝑓𝑓 ρ𝑔𝑔 � � � + (1 + 𝑐𝑐𝑐𝑐𝑐𝑐β)cosϕ� [σ(ρ𝑙𝑙 − ρ𝑣𝑣 ) ]1/4
16 𝜋𝜋 4
Certain micro/nanostructured enhanced surfaces exhibit very high surface wettability with
apparent contact angles ranging from 0° to 20°. For these surfaces, substantial deviations
between measured CHF values and predicted values from Eq. (7) were observed [53, 54]. Kim
and Kim [53] characterized surface wettability using the capillarity of the surface. Increased
surface capillarity facilitates the resupply of liquid to dry regions under high heat flux which
further delays CHF. To account for the enhancement of CHF due to capillary wicking effects,
Ahn et al. [54] modified Kandlikar’s CHF model by including a scaling factor (A1) to Eq. (7)
and an additional term for capillarity effects, as shown in Eq. (8). On the other hand, Chu et al.
[55] included Wenzel’s equation [56] in their CHF model to account for the microscopic
surface topology of silicon surfaces fabricated with arrays of micro-sized pillars. As shown in
Eq. (9), CHF is related to the surface roughness factor (r), which can be calculated from the
geometries of the surface micro-features, and the liquid receding angle on the corresponding
smooth surface (θrec). Other models which account for the effects of microstructures topology
have also been proposed by Zou and Maroo [57] and Rahman et al. [58]. More recently, Quan
et al. [59] obtained a modified Taylor instability wavelength for microstructured surfaces by
analyzing the change in surface free energy. Adopting the force balance approach of Kandlikar
[52] and by including the effects of capillary wicking force and the modified Taylor
wavelength, a CHF model as shown in Eq. (10) was developed. The model indicates the
dependence of CHF on the microstructures surface features which are characterized by the
surface roughness factor (r) and solid fraction (φs). The model was compared against existing
experimental data for enhanced surfaces in water, ethanol and FC-72, and most of the predicted
CHF values were found to be within ±25% of the experimental results.
8
Pre-publication version
K.C. Leong et al. Applied Thermal Engineering 112 (2017) 999–1019
http://dx.doi.org/10.1016/j.applthermaleng.2016.10.138
1/2
′′ 1/2 1 + cosβ 2 𝜋𝜋
𝑞𝑞𝐶𝐶𝐶𝐶𝐶𝐶 = 𝐴𝐴1 × ℎ𝑓𝑓𝑓𝑓 ρ𝑔𝑔 � � � + (1 + 𝑐𝑐𝑐𝑐𝑐𝑐β)cosϕ� [σ(ρ𝑙𝑙 − ρ𝑣𝑣 ) ]1/4
16 𝜋𝜋 4
�����������
𝜀𝜀𝜀𝜀𝜌𝜌𝑙𝑙 ℎ𝑓𝑓𝑓𝑓 𝐴𝐴2 𝑑𝑑𝐴𝐴 (8)
𝑤𝑤𝑤𝑤𝑤𝑤𝑤𝑤𝑤𝑤𝑤𝑤
+
𝐴𝐴ℎ𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒 𝑑𝑑𝑑𝑑
1/2
′′ 1/2 1+ cosβ 2(1 + 𝑟𝑟cos𝜃𝜃𝑟𝑟𝑟𝑟𝑟𝑟 ) 𝜋𝜋
𝑞𝑞𝐶𝐶𝐶𝐶𝐶𝐶 = ℎ𝑓𝑓𝑓𝑓 ρ𝑔𝑔 � �� + (1 + 𝑐𝑐𝑐𝑐𝑐𝑐β)cosϕ� [σ(ρ𝑙𝑙
(9)
16 𝜋𝜋(1 + cosβ) 4
1/4
− ρ𝑣𝑣 ) ]
′′ 1/2 1 + cosβ
𝑞𝑞𝐶𝐶𝐶𝐶𝐶𝐶 = ℎ𝑓𝑓𝑓𝑓 ρ𝑔𝑔 [σ(ρ𝑙𝑙 − ρ𝑣𝑣 ) ]1/4 � �
16
1/2 (10)
2 𝑟𝑟 + cosβ 𝜋𝜋 1
× � (1 − �φ𝑠𝑠 ) + (1 − �𝜑𝜑𝑠𝑠 )2 (1 + 𝑐𝑐𝑐𝑐𝑐𝑐β)cosϕ�
𝜋𝜋 1 + cosβ 4
Over the years, numerous enhanced surfaces were fabricated for improving nucleate boiling
heat transfer in dielectric fluids. These surfaces can be broadly classified into (1) coated
surfaces – surface features are directly coated onto the substrate surfaces, (2) intrinsic surfaces
– surface features that are intrinsically part of the material’s surface and (3) hybrid surfaces –
surface features which consist of combinations of (1) and/or (2). In this section, the recent
progress in nucleate pool boiling of dielectric fluids with enhanced surfaces is reviewed. An
attempt is made to classify the surfaces based on their fabrication techniques. In addition, the
various fabrication techniques employed are briefly summarized and the corresponding heat
transfer performances of the enhanced surface are examined and compared.
Epoxy binding
One of the initial methods employed in the fabrication of enhanced surfaces is the direct coating
of micro-sized particles onto plain surfaces using adhesive epoxies. Synthetic diamond
particles were often used as the coating material due to its superior thermal conductivity. The
coating forms porous structures with cavities which could increase the active nucleation site
density [60]. Chang and You [61] investigated the pool boiling performances of surfaces coated
with different sizes of diamond particles (2 µm to 70 µm) in saturated FC-72 and reported that
the surface coated with 20 µm diamond particles exhibited the highest heat transfer coefficients
under high heat fluxes (> 2.5 W/cm2). Rainey and You [60] conducted comprehensive studies
on the effects of surface orientation and heater size in saturated FC-72 and concluded that the
coated surfaces were insensitive to these parameters. A subsequent analysis on the bubble
dynamics by Kim et al. [62] on wire heaters showed that the diamond particle coating resulted
in higher bubble departure frequencies and smaller bubble departure diameter as compared to
an untreated surface. It was further suggested that the coated surfaces enhanced heat transfer
through the increase in latent heat transfer at low heat fluxes and increased convection heat
transfer at high heat fluxes. Arik et al. [63] showed that the effusivity-based CHF correlation
[Eq. (6)] provided good predictions for diamond coated surfaces tested at various pressure and
subcooling temperatures, with a deviation of ±15%.
9
Pre-publication version
K.C. Leong et al. Applied Thermal Engineering 112 (2017) 999–1019
http://dx.doi.org/10.1016/j.applthermaleng.2016.10.138
Takata et al. [64] demonstrated the use of dipping method in the coating of substrate surface
with micro/nano-sized particles. The coating was achieved by polishing the copper substrate
surface to mirror finish and subsequently cleaning the surface with hydrochloric acid and
ethanol before dipping the substrate in TiO2 suspension and heating the samples at 150ºC. Wu
et al. [65], on the other hand, coated TiO2 nanoparticles of 10 nm particle size using the dripping
process. One drop of diluted TiO2-ethanol solution was deposited and spread over the substrate
surface. The substrate was then heated to 200ºC which evaporated the ethanol leaving 1 µm
thickness of TiO2 coating. TiO2 was known for its high affinity for water and superhydrophilic
surfaces with near 0º water-surface contact angle could be produced using TiO2 coating.
Enhanced boiling performances of TiO2 coated surfaces in water were suggested to be due to
the surface’s increased water affinity [64]. However, even for highly wetting fluids such as FC-
72, Wu et al. [65] demonstrated that significant boiling performances can be achieved with
TiO2 coating where improvements in heat transfer coefficients and CHF were up to 91.2% and
38.2% as compared to plain copper. The effective liquid-solid interactions induced by
superhydrophilic TiO2 coating was suggested for the enhancements observed.
Spray coating
In this method, coating is achieved by accelerating the coating particles at high velocities and
impinging them directly onto the substrate surfaces. Due to the high impact experienced by the
particles as they bombarded the surfaces, the particles undergo plastic deformation and adhere
to the surfaces. In some instances, the coating precursors are also pre-heated or melted before
spraying onto the surfaces. For instance, Dewangan et al. [66] coated copper powder onto
horizontal copper tubes using a gas flame spraying technique and tested the surfaces in pure
and quasi-azeotropic refrigerants. On the other hand, using the cold gas dynamic spraying
method, Pialago et al. [67, 68] fabricated surfaces coated with carbon nanotubes (CNT)-
metallic powder composites and performed their boiling experiments in R134a. Recently, Sahu
et al. [69] coated polymer nanofibers onto the copper substrate by employing the electrically-
assisted supersonic blowing technique. Air was blown out of a nozzle at supersonic velocity
and 6 wt% PAN solution in DMF was pumped through a gage needle. An electric potential
was set up between the needle and the nozzle to attract the polymer towards the nozzle which
was then carried and impinged onto the substrate. To improve thermal conduction,
electroplating was further carried out on the coated surfaces. These processes resulted in the
formation of a copper-plated nanofiber mat as shown in Fig. 6. Employing a similar coating
method, Sahu et al. [70] investigated the pool boiling heat transfer of copper-plated nanofiber
coated surface and their results in HFE-7300 showed that CHF was 33% higher and significant
enhancements in heat transfer coefficients were obtained as compared to an untreated copper
surface. In addition, prolonged boiling has also resulted in a slight reduction in the nanofiber
surface roughness although no change in the heat transfer characteristics of the surface was
observed. Using supersonic spraying, a surface coated with reduced graphene oxide (rGO) was
produced by An et al. [71] and their boiling experiments were performed in FC-72 at system
pressure of 0.24 bar and fluid temperature of 20ºC. In comparison with a bare copper surface,
the coating of rGO film improved heat transfer coefficients and CHF by 1.5 times and 0.5 times,
respectively. The improved wettability and large number of boiling nucleation sites of the rGO
film were suggested to have resulted in the improved heat transfer characteristics.
10
Pre-publication version
K.C. Leong et al. Applied Thermal Engineering 112 (2017) 999–1019
http://dx.doi.org/10.1016/j.applthermaleng.2016.10.138
The use of free particles in enhancing nucleate boiling heat transfer was demonstrated by Kim
et al. [72]. In this method, the particles are free to move on the boiling surface. Copper particles
were selected as the high density of copper makes the particles more resilient to removal from
the surface by the departing bubbles during boiling. The narrow corner cavities formed by the
free particles were suggested to have increased the number of surface nucleation sites. Various
sizes of particles were investigated in FC-72 and an enhancement in the heat transfer coefficient
of 76.3% and CHF of 160 kW/m2 were achieved with the 10 µm particle size. Subsequently,
the boiling performances of surfaces with sintered and free particles were studied by Sarangi
et al. [73]. Free particles deposited surfaces similar to those reported by Kim et al. [72] were
produced, while sintered coatings were fabricated by placement of copper substrates onto
graphite molds packed with copper particles. The particles were then sintered at an elevated
temperature of 950ºC (Fig. 7). Experiments in FC-72 showed that surfaces with sintered
coating resulted in significantly higher heat transfer coefficients as compared to free particles
and plain surfaces. For instance, the best performing sintered coating provided 95% reduction
in wall superheat for particle sizes ranging from 90-106 µm. An investigation of sintered
powder coated surfaces was also carried out by Thiagarajan et al. [74]. Copper-rich
microparticles of 5-20 µm diameters were sintered onto copper substrates to produce different
coating thicknesses. Experiments were conducted in HFE-7100 and enhancements in the heat
transfer coefficient by 50-270% and CHF by 33-60% as compared to a plain surface were
obtained. Based on bubble visualization performed and a heat flux partition model, it was
suggested that the heat transfer enhancement was due to the significantly higher active
nucleation site density on the sintered surfaces.
Electrochemical deposition
El-Genk and Ali [75, 76] and Ali and El-Genk [77, 78] employed the electrochemical process
to fabricate copper microporous coated surfaces. In this fabrication technique, the copper
substrate to be coated was positioned as the cathode in the electrolysis process, where an
electrolyte solution of sulfuric acid and copper sulfate were used. A high current density (3
A/cm2) was transmitted through the cathode and anode resulting in the deposition of copper
nano-dendrites on the substrate. Due to the generation and departure of hydrogen bubbles
during the chemical process, pores were formed and the coating thickness could be varied by
controlling the duration of the process. Images illustrating the microporous structures obtained
from the scanning electron microscope (SEM) are depicted in Fig. 8. El-Genk and Ali [75]
performed saturated pool boiling experiments with porous surfaces of different coating
thicknesses in PF-5060. For all the porous surfaces investigated, CHF was found to range from
22.7 W/cm2 to 27.8 W/cm2 with the surface superheats between 2.16 K to 12.89 K. The porous
surface with the medium range coating thickness (171.1 μm) was recorded to have the highest
CHF (27.8 W/cm2) at the lowest wall superheat (2.16 K). Ali and El-Genk [77] subsequently
studied the pool boiling heat transfer performance of microporous surfaces at different surface
orientations. The experimental results revealed that for all surfaces, CHF and the corresponding
wall superheats reduced with increasing surface inclinations. However the CHF for porous
surfaces still outperformed plain copper surface at all surface orientations. For instance, at 180°
surface inclination, the CHF values for 80 μm and 137 μm coating thicknesses were
respectively 74% and 87% higher than the plain copper surface. Similarly, by using
electrochemical deposition, Furberg and Palm [79] fabricated microporous structures and
investigated their pool boiling heat transfer performance in both R134a and FC-72.
Visualization of the bubble departure phenomenon using high speed imaging revealed that
bubble coalescence took place between adjacent pores at medium to high heat fluxes (> 5
11
Pre-publication version
K.C. Leong et al. Applied Thermal Engineering 112 (2017) 999–1019
http://dx.doi.org/10.1016/j.applthermaleng.2016.10.138
W/cm2) whereas relatively low bubble coalescence was observed at low heat fluxes (2 W/cm2).
In addition, the bubble diameter from the porous surface was also found to be 10 to 30% smaller
than plain surface whereas higher bubble departure frequency was detected. It was suggested
that the enhanced heat transfer was due to evaporation of fluid within the porous structures and
the interconnected porous network served as effective vapor escape paths.
Chemical vapor deposition (CVD) enables highly ordered structures to be synthesized onto the
substrate surface. A catalyst is usually deposited onto the surface and the synthesis process is
typically performed by exposing the surface to volatile precursor and carrier gas. Several
researchers have investigated the pool boiling performances of vertically aligned CNT coated
surfaces produced by CVD technique. For instance, Ujereh et al. [80] fabricated multi-walled
carbon nanotubes (MWCNT) arrays of different CNT densities and coating patterns. The
MWCNTs are approximately 50 nm in diameter and 20 to 30 μm in height. Saturated pool
boiling of FC-72 enhanced CHF by 45% and the fully-coated CNT surface also exhibited 450%
enhancement in heat transfer coefficient as compared to bare silicon. Subsequently,
Sathyamurthi et al. [81] studied the performance of MWCNT in saturated and subcooled pool
boiling of PF-5060 dielectric fluid. CNT coatings of 9 and 25 µm thicknesses were synthesized
on atomically smooth silicon substrates where moderate enhancements in heat transfer
coefficients were observed as compared to bare silicon. More recently, Ho et al. [50] reported
a detailed investigation on pool boiling of CNT-coated surface under different surface
orientations in saturated FC-72. Using iron as catalyst and pure acetylene as the carbon source,
highly aligned CNTs of 215 µm thickness were produced (Fig. 9). Enhancements in the average
heat transfer coefficient and CHF of up to 86% and 42% were respectively achieved as
compared to bare silicon. In addition, the increase in surface orientation was also shown to
reduce the average heat transfer coefficient of the CNT-coated surface. Finally, a modified
Rohsenow correlation was proposed to characterize the pool boiling curves of the surfaces at
different orientations.
Selected publications on pool boiling with coated surfaces reviewed above are summarized in
Table 2 and comparisons of the boiling curves of the better performing surfaces from the
selected publications are reproduced in Fig. 10. The reported enhancements in the heat transfer
coefficients of the coated surfaces vary from 32% to more than 17 times as compared to
plain/reference surfaces whereas CHF increases between 10% and 87%. From this review, it
can be seen that the numerous surface coating techniques presented have resulted in
significantly enhanced boiling performance which made these techniques attractive for many
cooling applications. However, as the coatings are not intrinsically part of the substrate surface,
the associated thermal contact resistance between the coating and substrate and the tendency
of the coating to degrade and peel over time are some of the disadvantages. However, these
drawbacks could be overcome by intrinsic surface feature designs, where the microstructures
are part of the material surface. In the following section, the techniques employed in the
fabrication of intrinsic surfaces and their accompanying heat transfer performances shall be
discussed.
12
Pre-publication version
K.C. Leong et al. Applied Thermal Engineering 112 (2017) 999–1019
http://dx.doi.org/10.1016/j.applthermaleng.2016.10.138
Electrical discharge machining (EDM) uses electrical discharges or sparks to remove material
from a workpiece and thereby obtain the desired shapes and geometries. Guglielmini et al. [82]
uses EDM to fabricate fin arrays of different configurations. Fin lengths of 3 mm or 6 mm and
fin thicknesses of 0.4 mm or 1.0 mm were produced. Experiments were performed to examine
the pool boiling characteristics of these surfaces in FC-72 under different saturation pressures.
The maximum overall heat transfer coefficients were observed to increase with fin thickness
and spacing but decrease with fin length. Similarly, Yu and Lu [83] also produced millimeter-
size rectangular fin arrays of different fin dimensions and arrangements using EDM. Boiling
experiments and visualizations were performed in saturated FC-72. The results indicated that
closely-spaced and longer fins caused high flow resistance to the departing bubbles along the
fin and delay departure of the bubbles at the base of the fin. In addition, the overall heat transfer
coefficient also decayed rapidly as fin spacing decreased or fin length increased. A maximum
CHF of 9.8 × 105 W/m2 was achieved with 0.5 mm fin spacing and 4.0 mm fin length, which
was 5 times higher than that for the plain surface.
Jones et al. [84] produced samples of different surface roughness using EDM. The samples
were characterized using various surface roughness parameters and have average roughness
(Ra) ranging from 1.08 µm to 10.0 µm. Experiments were conducted with FC-77 and water.
The results indicated stronger influence of Ra on heat transfer coefficients with FC-77 as
compared to water. The roughest surface achieved 210% higher heat transfer coefficients as
compared to a polished surface in FC-77. High speed visualization experiments on the boiling
phenomenon of polished and EDM-roughened surface were subsequently carried out by
McHale and Garimella [85]. They reported that the EDM-roughened surface resulted in smaller
bubble departure diameters and larger departure frequencies as compared to a polished surface
for a given heat flux.
The effects of surface roughness were also explored by El-Genk and Suszko [86] where the
copper samples were prepared by ablating the surfaces with emery papers of different grit
counts. Samples with Ra ranging from 0.039 µm to 1.79 µm were tested at different surface
orientations in saturated PF-5060. For all the surfaces, the maximum heat transfer coefficients
(hMNB) for the downward-facing orientation were found to be approximately 40% of the
upward-facing orientation. In addition, the hMNB of 1.72 W/cm2·K was also recorded for the
rougher surface (Ra = 1.44 µm) whereas hMNB of the polished surface was only 0.67 W/cm2·K,
which represents more than 150% enhancement due to the increased surface roughness.
Visualization of the bubble dynamics on these surfaces by Suszko and El-Genk [87] showed
that the smooth copper surface has higher bubble growth rate, departure diameter and
detachment frequency than the rougher copper surfaces and, for all surfaces, the bubble growth
rates were proportional to the square root of time indicating that the bubble growth was
thermally driven. In addition, based on the results obtained, it was also concluded that the
enhancement in heat transfer with increasing surfaces roughness was primarily due to the
increase in surface active nucleation site density.
In the past two decades, MEMS/NEMS processes such as depositions, photolithography and
etching that were commonly used in semiconductor device fabrication have also been explored
to produce enhanced surfaces for promoting boiling. Highly-ordered micro/nano-size surface
features could be produced with these processes. For example, arrays of microcavities with
13
Pre-publication version
K.C. Leong et al. Applied Thermal Engineering 112 (2017) 999–1019
http://dx.doi.org/10.1016/j.applthermaleng.2016.10.138
mouth diameters of 50 µm to 200 µm have been fabricated by Yu et al. [88]. Honda et al. [89],
on the other hand, produced micro-pin-fin arrays of 50 µm fin thickness and fin heights of 60
µm. Wei et al. [90], subsequently, extended the studies to include surfaces of five different pin
fin geometries. Pool boiling results in FC-72 showed that the micro-pin-fin surfaces displayed
distinctively steep boiling curves with very small change in wall superheat with increasing heat
flux. Wei et al. [90] suggested that the steep boiling curve was due to the highly-structured and
uniformly-distributed micro-pin-fins which resulted in the activation of the nucleation sites at
almost the same wall superheat. More recently, micro-pin-finned surfaces were investigated by
Xue et al. [91] under the effects of microgravity. In comparison with the boiling curves of the
micro-pin-finned surfaces under normal gravity, microgravity showed insignificant effect on
boiling heat transfer. Xue et al. [91] suggested that the independence of gravity was due to the
micro-pin-fin structures which allowed sufficient supply of fluid to the heated surface by
capillary driving force. Subsequently, a new model for predicting the bubble departure radius
from the micro-pin-finned surface under microgravity was developed by Zhang et al. [92].
Graphite and metallic foams are the two types of open cell foams that have been investigated
for boiling heat transfer applications. Graphite foams can be produced from carbon-rich
precursors (also known as pitch) where the final forms are obtained by processes such as
blowing, carbonization and graphitization under inert and high temperature environment [93].
On the other hand, metallic foams can be fabricated using infiltration casting technique which
is typically achieved by solidification of superheated liquid in high pressure and vacuum
environment [94].
Experimental studies on pool boiling of metallic foams of different pore densities were
conducted by several researchers [95, 96]. However, their investigations were performed in
deionized water and surfactant solutions. On the other hand, detailed studies on the pool boiling
performance of porous graphite foams in dielectric fluids have been reported by Leong et al.
[97], Jin et al. [98], Pranoto et al. [99], El-Genk and Parker [100] and El-Genk [101]. As shown
in Fig. 11 (a), graphite foams have spherical and interconnected pores which are favorable for
bubble generation. In addition, the large surface-to-volume ratio and high bulk thermal
conductivity (150 W/m·K) [93] also made graphite foams a promising material for enhanced
boiling heat transfer applications. Leong et al. [97] investigated cubic graphite foams
(“Pocofoam” and “Kfoam”) of different bulk thermal conductivities in FC-72. Their results
indicated that the boiling thermal resistances (Rb) of graphite foams were approximately 2 times
lower than a copper block even though they have lower bulk thermal conductivity. “Kfoam”
exhibited slightly better boiling performance than “Pocofoam”. Jin et al. [98] studied the effects
of graphite foams in FC-72 and HFE-7000 and determined that lower Rb was achieved with
HFE-7000 when compared with FC-72. Pranoto et al. [99] subsequently extended the studies
to include finned porous graphite as shown in Fig. 11 (b). The finned porous graphite, however,
exhibited lower average heat transfer coefficients than the block graphite. Based on the high
quality images of the boiling phenomenon, it was evident that larger bubble density was
generated from the block graphite as compared to the finned graphite, suggesting that the
internal pores of the graphite foam contributed significantly to the enhanced heat transfer
performance. On the other hand, El-Genk and Parker [100] conducted detailed investigations
on boiling performance of porous graphite in FC-72 and HFE-7100 under saturated and
subcooled conditions and over a range of surface orientations. It was shown that up to 52.7%
higher boiling heat transfer coefficient was achieved with porous graphite in FC-72 as
compared to HFE-7100. In addition, the surface orientation had negligible effect on the
14
Pre-publication version
K.C. Leong et al. Applied Thermal Engineering 112 (2017) 999–1019
http://dx.doi.org/10.1016/j.applthermaleng.2016.10.138
maximum heat transfer coefficient (ℎ𝐵𝐵∗ ) of the porous graphite although ℎ𝐵𝐵∗ was seen to increase
linearly with increasing subcooling.
Selective Laser Melting (SLM) is an additive manufacturing (AM) technique which utilizes a
high-power laser source to melt and fuse the base metal powder layer-by-layer to form a three-
dimensional geometry. This technique enables customized parts and highly complex designs
to be fabricated based on the input from computer aided design (CAD) software. Due to these
advantages, there has been an increasing use of SLM for producing functional heat transfer
devices [102-104]. However, publications on enhancing nucleate boiling heat transfer by SLM
are scarce. Ho et al. [105] explored the possibility of improving nucleate pool heat transfer of
FC-72 by fabricating microstructured surfaces of different designs. Micro-cavity, micro-fin and
micro-channel surfaces were produced from an aluminum alloy, AlSi10Mg, and their heat
transfer performances were experimentally investigated in a water-cooled thermosyphon. The
enhanced surfaces exhibit 63.5% increase in heat transfer coefficient as compared to a plain
AlSi10Mg surface. In a separate investigation, Ho et al. [106] further evaluated the nucleate
boiling characteristics of a plain AlSi10Mg surface fabricated by SLM against that of a
commercially casted plain Al-6061 surface. The results indicated that the plain AlSi10Mg
improved the heat transfer coefficient by 40.6% and CHF by 31% as compared to the plain Al-
6061 surface. The enhanced thermal performance of the plain AlSi10Mg surface was attributed
to the inherent microscale cavities and grooves that were formed as a result of the laser melting
process. In addition, surfaces with arrays of micro-cavity and micro-fins of different
configuration were also produced where further enhancements in the average heat transfer
coefficient and CHF of 70% and 76% as compared to plain Al-6061 were recorded. While the
graphite foams, as discussed above, consist of non-uniformly distributed pores of different pore
sizes, the possibility of fabricating structured porous substrates for enhancing pool boiling
using SLM was recently demonstrated by Wong and Leong [107]. The porous structures were
engineered by replicating arrays of Octet-truss unit cells to form specimens of heights ranging
from 2.5 mm to 10 mm as shown in Fig. 12. The structures were tested in saturated FC-72 and
improvement in heat transfer coefficients and CHF of up to 91.7% and more than 100% were
obtained respectively for the porous structures as compared to plain Al-6061.
Selected publications on pool boiling with intrinsic surfaces reviewed above are summarized
in Table 3 and comparisons of the boiling curves of the better performing surfaces from
selected publications are reproduced in Fig. 14.
The fabrication of hybrid surfaces usually involves two or more of the processes presented in
Sections 2.2.1 and 2.2.2. For instance, McHale et al. [108] applied a particle sintering process
followed by microwave-plasma CVD to produce enhanced surfaces with CNT arrays grown
over sintered copper particles. The sintered copper particles were coated over the surface by
the sintering process whereas CVD was employed to grow the CNTs. Their experiments in
HFE-7300 showed that the hybrid surface resulted in the highest boiling heat transfer
performance as compared to other enhanced surfaces. In addition, it is also interesting to note
that the enhancement in CHF of the hybrid surface as compared to plain copper was also higher
than the combined CHF enhancements of the enhanced surfaces produced by individual
technique.
15
Pre-publication version
K.C. Leong et al. Applied Thermal Engineering 112 (2017) 999–1019
http://dx.doi.org/10.1016/j.applthermaleng.2016.10.138
Im et al. [109], on the other hand, fabricated microgrooved surfaces deposited with flower-like
copper oxide nanostructures. The microgrooves were created by cutting the surface using an
automatic die saw whereas the flower-like nanostructures were deposited by immersing the
surface into NaOH solution with (NH4)2S2O8 oxidant. Similar to McHale et al. [108], higher
boiling heat transfer coefficient and CHF were achieved with the hybrid surface as compared
to the enhanced surfaces fabricated by individual process. In addition, capillary wicking
induced by the large surface-to-volume ratio was suggested as the primary mechanism for the
enhancements observed.
Recently, hybrid surfaces were also developed by Patil and Kandlikar [110]. The surfaces
consisted of micro-size open channels with microporous coating on the top of the channel fins.
The open channels were fabricated by CNC machining whereas a two-stage electrodeposition
technique was employed to achieve the preferential microporous coating. As explained by Patil
and Kandlikar [110], and as illustrated in Fig. 13, this surface design allowed bubble nucleation
to take place on the porous coating at the fin top whereas liquid circulation was generated in
the microchannel. Hence, the additional nucleation sites on the porous coating and the current
of fluid flow in the microchannels provided cohesive mechanisms for enhancing the pool
boiling performance. Following Patil and Kandlikar [110], Jaikumar and Kandlikar [111]
experimentally investigated similar hybrid surfaces with a fin top microporous coating but with
varying microchannel sizes in FC-87 dielectric fluid. In all, the hybrid surface recorded 270%
enhancement in CHF as compared to the plain surface and a maximum heat transfer coefficient
of 20 kW/m2·K was achieved.
16
Pre-publication version
K.C. Leong et al. Applied Thermal Engineering 112 (2017) 999–1019
http://dx.doi.org/10.1016/j.applthermaleng.2016.10.138
3. Flow boiling
3.1 Overview
Flow boiling occurs under the presence of a pump to introduce convective heat transfer and to
allow orientation flexibility. Due to fluid convection, nucleate flow boiling heat transfer can
exceed that of pool boiling. However, in comparison with pool boiling, the flow boiling process
is additionally affected by factors such as fluid mass flux and vapor quality. In addition, the
presence of a pump also creates additional system complexity and pressure drop is also an
important consideration which affects the cost and size of the pumping system.
Boiling-dominated and convection-dominated modes are the two modes of heat transfer
commonly observed in flow boiling processes. In the boiling-dominated mode, the heat
transfer coefficient varies significantly as heat flux changes whereas in the convection-
dominated mode, mass flux and vapor quality have more significant effects on the heat
transfer coefficient.
Flow boiling can also be divided into different regimes, as shown in Fig. 15, namely bubbly
flow, plug flow, slug flow, wavy flow and annular flow. Many researchers have presented
correlations to predict flow boiling heat transfer performance and pressure drop [112-117].
Kandlikar [116] had compiled correlations for saturated two-phase flow boiling heat transfer
inside horizontal and vertical tubes using a wide range of experimental data and 10 different
fluids. Thome et al. [117] presented a comprehensive review of flow boiling heat transfer, two-
phase pressure drops and flow patterns of ammonia and hydrocarbons applied in air-
conditioning, refrigeration and heat pump systems.
Flow boiling heat transfer can be enhanced in channels with small dimensions due to the
increase of surface-to-volume ratio. The mini/microchannel can be fabricated by etching or
mechanical cutting methods. Thome [118] presented a review of flow boiling mechanisms and
models in mini and microchannels. A comprehensive review on the history, advances and
challenges of flow boiling heat transfer in microchannel is covered by Kandlikar [119]. Bertsch
et al. [120] compiled predictions from 25 heat transfer correlations using a wide range of
experiments from 10 independent sources. The hydraulic diameter of the tubes and channels
considered in all the experiments was 2 mm or smaller.
17
Pre-publication version
K.C. Leong et al. Applied Thermal Engineering 112 (2017) 999–1019
http://dx.doi.org/10.1016/j.applthermaleng.2016.10.138
transition. The Froude number (Fr) was used to account for the inertia and gravity forces. The
liquid and vapor phase viscosity and density, Reynolds number (Re) and Weber number (We)
were used to account for the viscous, surface tension and shear forces. The definitions of the
Froude number, Reynolds number and Weber number are given in Eqs. (12), (13) and (14),
respectively. The introduction of these dimensionless numbers helps to better account for the
transition from surface tension-dominated isolated bubble flow to shear-dominated annular
flow. More recently, Tibirica and Ribatski [124] have completed a review of the
characterization of the transition between flow boiling under macroscale and microscale
conditions.
1 σ
Co = (11)
𝐷𝐷 �𝑔𝑔(ρ𝑙𝑙 −ρ𝑔𝑔 )
𝐺𝐺 2
Fr = ρ2𝑔𝑔𝑔𝑔 (12)
𝑙𝑙
𝐺𝐺𝐺𝐺
Re = (13)
μ
𝐺𝐺 2 𝐷𝐷
We = (14)
σρ
Harirchian and Garimella [125] proposed a new transition criterion using the convective
confinement number, which predicts the conditions under which microscale confinement
effects are exhibited in flow boiling. The Boiling number (Bl), which is the nondimensional
form of the heat flux, and Bond number (Bo), which is the ratio of buoyancy to surface tension
forces, are introduced and shown in Eqs. (15) and (16). The convective confinement number is
defined as the product of Reynolds number (Re) and the square root of Bo. It is shown that for
flow boiling of FC-77, physical confinement in the microchannels exists for convective
confinement numbers smaller than 160. Under this condition, thin film evaporation contributes
to heat transfer in addition to nucleate boiling, and results in larger values of heat transfer
coefficient compared to those cases in which no confinement is observed and nucleate boiling
is dominant. A comprehensive flow regime is computed for FC-77 as shown in Fig 16.
𝑞𝑞 "
Bl = Gℎ (15)
𝑓𝑓𝑓𝑓
Kandlikar [126] used a scaling analysis to identify the relative effects of different forces on the
flow boiling process at the microscale level. The different forces considered for scaling are
inertia, surface tension, shear, gravity and evaporation momentum forces. Flow pattern
transitions and stability for flow boiling of water and FC-77 were evaluated. It was found that
surface tension and evaporation momentum forces are important at the microscale level. The
elongated bubble/slug flow pattern is presented as a limiting factor in heat transfer of a
microchannel, as shown in Fig. 17. This flow pattern is similar to the bubble growth and
departure phases in pool boiling. Transient conduction/microconvection was proposed to be
the dominant heat transfer mechanism associated with this flow pattern.
18
Pre-publication version
K.C. Leong et al. Applied Thermal Engineering 112 (2017) 999–1019
http://dx.doi.org/10.1016/j.applthermaleng.2016.10.138
Wang et al. [127] experimentally studied flow boiling of FC-72 and ethanol in microchannels
of high aspect ratios. The hydraulic diameters (dh) of the microchannels considered were 571
μm, 762 μm and 1454 μm. Mass fluxes of 11.2 kg/m2·s, 22.4 kg/m2·s and 44.8 kg/m2·s were
investigated. High speed visualization was employed to capture the flow patterns. Periodic
recoiling and re-wetting were captured for both liquids as shown in Fig 18.
Flow instabilities have been observed by researchers during flow boiling in microchannel.
Hetsroni et al. [128] performed flow visualization on boiling in microchannel and observed
explosive boiling with periodic wetting and dry-outs. Pressure drop oscillations were also
observed and they tend to increase with increasing vapor quality. The study also showed the
strong dependence of the heat transfer coefficient on the vapor quality. Wang et al. [129]
performed flow boiling using water in microchannel and attributed the instability to periodic
bubble expansion in both upstream and downstream directions, causing increase of wall
temperature and pressure drop. Bogojevic et al. [130] further identified two types of flow
instabilities, namely high amplitude/low frequency oscillations and low amplitude/high
frequency oscillations. In order to reduce the effect of flow instabilities, Kosar et al. [131]
introduced flow restrictors at the microchannel inlet. Kandlikar et al. [132] introduced artificial
nucleation sites and pressure drop elements. The combination of fabricated nucleation sites in
conjunction with the 4% area pressure drop elements completely eliminated the instabilities
associated with the reverse flow although this was at the expense of higher pressure drop. In
order to better understand the pressure drop issues, Kim and Mudawar [133] has compiled a
comprehensive review on the prediction of two-phase frictional pressure drop for
mini/microchannel saturated flow boiling. A wide range of working fluids, hydraulic
diameters, mass fluxes and reduced pressures were considered.
Harirchian and Garimella [134] investigated the effect of two-phase heat transfer coefficient
and pressure drop on different sizes of microchannels. FC-77 was used as the coolant fluid.
Seven different parallel microchannels with fixed depth 400 µm and widths between 100 and
5850 µm were fabricated using a dicing saw on a silicon chip. It was observed that the heat
transfer coefficient was independent of microchannel width above 400 µm. For a microchannel
of 100 µm width, occurrence of annular flow at a lower heat flux range resulted in high heat
transfer coefficient, but the performance became similar to microchannels of other widths at
high heat flux. The pressure drop of two-phase flow was observed to increase with decreasing
microchannel width at a fixed heat flux. The optimum width was determined to be 400 µm for
high transfer and pumping power. Harirchian and Garimella [135] further extended their work
by performing experiments with five additional microchannel substrates with channel depths
of 100 and 250 μm and widths ranging from 100 to 1000 μm. Flow visualizations and heat
transfer results showed that the cross-sectional area of the microchannel is important instead
of solely relying on width, depth or aspect ratio. It was observed that vapor confinement
occurred readily in microchannel of cross-sectional area smaller than 0.089 mm2. Under this
threshold, contributions of nucleate boiling and liquid film evaporation were proposed to
enhance heat transfer.
19
Pre-publication version
K.C. Leong et al. Applied Thermal Engineering 112 (2017) 999–1019
http://dx.doi.org/10.1016/j.applthermaleng.2016.10.138
Ma et al. [136] fabricated micro-pin-fins on chip surfaces by using a dry etching technique to
enhance the boiling heat transfer of FC-72. Different fluid velocities, liquid subcooling and
chip configuration were investigated in their study. It was found that all the micro-pin-finned
surfaces showed considerable heat transfer enhancement and increased CHF compared to the
smooth surface. They also stated that the CHF values for all surfaces increased with fluid
velocity and subcooling. The heat transfer could also be increased by increasing fin height. Wei
et al. [137] conducted flow boiling on a silicon chip with micro-pin-fins fabricated using dry
etching technique. Sub-cooled flow boiling with different flow velocities of FC-72 were
performed on the silicon chip. It shows that the boiling heat transfer can be enhanced by
increasing the total surface area. The evaporation of superheated liquid within the confined
gaps between fins and micro-convection caused by thermocapillary force due to the suction of
a bubble hovering on the top of micro-pin-fins were proposed mechanisms which enhanced
heat transfer. Guo et al. [138] further conducted sub-cooled flow boiling with impingement of
FC-72 simultaneously on the micro-pin-finned silicon chip. They showed that the boiling heat
transfer can be enhanced by increasing the total surface area and the jet impingement velocity.
It was observed that a large jet impingement velocity caused more turbulence and micro-
convection, which helped to improve heat transfer at high heat flux and to delay the occurrence
of CHF. Krishnamurthy and Peles [139] conducted subcooled and saturated flow boiling
experiments of HFE-7000 across a single row of inline micro pin fins entrenched in a
microchannel. Heat transfer coefficient enhancement was observed for the microchannel with
fins as compared to a plain microchannel. The enhancement was higher with increasing
Reynolds number and it was attributed to the wake interaction between the pin fins and better
fluid mixing.
Rainey et al. [140] studied flow boiling from microporous coated surfaces in sub-cooled FC-
72. Their experimental results showed that microporous surfaces outperformed plain surfaces
and enhanced the CHF. Ammerman and You [141] performed flow boiling using FC-87 on
microporous coated surfaces. It was observed that the presence of coating resulted in boiling
incipience at lower wall superheats, increased heat transfer coefficients and delay of CHF. The
enhancement was attributed to the increased nucleation sites and bubble departure frequency.
As compared to a plain surface, the coated surfaces were able to increase CHF between 14%
and 36%. Sun et al. [142] coated copper particles of 20, 50 and 120 µm onto horizontal
minichannels, with hydraulic diameters ranging from 0.49 mm to 1.26 mm. The coatings were
produced by sintering spherical copper particles on the bottom surface of the channel. Heat
transfer coefficients of between 7 and 10 times were achieved as compared to an uncoated
surface. The confinement effect was also studied in their experiments. It was observed that the
decrease in the channel size and liquid mass flux strengthened the confinement effect, resulting
in the reduction in CHF values. Bai et al. [143] investigated flow boiling in parallel
microchannels with metallic porous coatings. Anhydrous ethanol was used as the fluid and
microchannels of hydraulic diameter of 540 µm were used. A solid-state sintering method was
used to coat copper particles of 30, 55 and 95 µm in diameter. The surfaces with porous
coatings experienced higher pressure drop as compared to the plain surface. This was attributed
to the increased shear stress due to the higher surface roughness and induced turbulence caused
by the porous coating. Heat transfer was enhanced by up to 90% for surfaces with porous
coating particularly at low vapor quality regime although the enhancement diminished greatly
when the vapor quality was greater than 0.25.
20
Pre-publication version
K.C. Leong et al. Applied Thermal Engineering 112 (2017) 999–1019
http://dx.doi.org/10.1016/j.applthermaleng.2016.10.138
Porous foams with large surface-area-to-volume ratio, low density, and high bulk thermal
conductivity can be used as boiling evaporators for high heat flux device systems. Fluid flow
mixing is enhanced due to the increased tortuosity of porous foams. Kim et al. [144] performed
flow boiling experiments of FC-72 in a foam-filled channel using copper foams of 10 and 20
pores per inch (PPI). It was reported that the increase of mass flux from 20 kg/m2·s to 48
kg/m2·s enhanced the heat transfer rate by as much as 15%. Further increase of the mass flux
up to 72 kg/m2·s did not enhance the cooling capability. The foam with a larger pore size was
shown to have a better heat transfer characteristic as the vapor could escape more easily at high
heat flux. Their experimental results showed that the high porosity and large pore size foam,
i.e., the 95% and 10 PPI copper foam, gave the best result, achieving a heat transfer coefficient
of 10 kW/m2·K. Lu and Zhao [145] performed numerical and experimental analyses of flow
boiling in horizontal metal foam tubes. It was found that the decrease in pore size with increase
of pore density (PPI) resulted in significant improvement to the overall heat transfer
performance due to the increase of surface area and strong flow mixing.
Graphite porous foams with high thermal conductivity were investigated by Pranoto and Leong
[146] for flow boiling using FC-72. Two different graphite foams of 61% and 72% porosity
were used. Evaporator gaps of 6, 4, and 2 mm were tested with coolant mass fluxes of 50, 100,
and 150 kg/m2⋅s. The experimental results show that the evaporator gap, coolant mass flux, and
foam properties have effects on the flow boiling characteristics and performance. It was found
that the use of foams of 61% and 72% porosity had enhanced the boiling heat transfer
coefficients by up to 2.5 and 1.9 times, respectively as compared to those of a smooth surface.
The better enhancement of the foam with 61% porosity is attributed to more active nucleation
sites due to the higher surface-area-to-volume ratio. High speed visualization had confirmed
the higher bubble departure frequency from the foam of 61% porosity.
Modeling of two-phase flow in porous foams is difficult due to the complexity of geometry
and fluid-vapor motions. Li and Leong [147] numerically and experimentally investigated flow
boiling of water and FC-72 in aluminum foams. The heat transfer process prior to the onset of
nucleate boiling and the hysteresis effect were investigated. It was found experimentally that
hysteresis occurred for water but was absent for FC-72. Numerical simulations were performed
for both single- and two-phase heat transfer using the finite volume method. Reasonable
agreements were obtained between the numerical and experimental results.
An experimental study of flow boiling heat transfer inside a channel filled with metallic foam
was performed by Madani et al. [148]. A metallic copper foam of 36 PPI and 97% porosity and
n-pentane were used in this study. The coolant mass flux and the heat flux were varied from 10
to 100 kg/m2·s and 0 to 25 W/cm2, respectively. Their experimental results showed that the
heat transfer coefficient was enhanced by a factor from 2 to 4 as compared to the plain tube
heat transfer coefficient calculated from the Gungor-Winterton correlation [149].
Selected publications on flow boiling with enhanced surfaces reviewed above are summarized
in Table 4.
4. Concluding remarks
This paper critically reviews recent publications of pool and flow boiling of dielectric fluids
over enhanced surfaces. The various models developed to characterize the nucleate boiling
21
Pre-publication version
K.C. Leong et al. Applied Thermal Engineering 112 (2017) 999–1019
http://dx.doi.org/10.1016/j.applthermaleng.2016.10.138
curves and to predict critical heat fluxes of plain and enhanced surfaces and fundamental
studies performed are analyzed and some of their benefits and shortcomings are highlighted.
The numerous surface fabrication/modification techniques used in enhancing boiling heat
transfer have achieved significant success over the last few decades. These
fabrication/modification techniques are summarized in this paper and attempts are also made
to categorize numerous enhanced surfaces reported in the literature. In addition, the heat
transfer performances of these enhanced surfaces are also evaluated and their associated
thermal transport mechanisms are elucidated.
Based on the above review conducted, the following areas which require further investigations
are identified:
(1) While it has been shown through bubble visualization and temperature measurements that
transient conduction and/or microconvection are the dominant mechanisms in heat transfer
of isolated bubbles on plain surfaces [34-38], the effects of enhanced surface features on
the bubble dynamics and the associated heat transfer mechanisms are still not well
established. Parametric studies of enhanced surfaces such as micro-fin, micro-cavity or
porous surfaces with different heights, diameters and pitches using high speed and high
resolution visualization are therefore essential to achieve fundamental understanding of the
bubble behavior and the mechanisms involved.
(2) In order to achieve fundamental understanding of the bubble behavior and the thermal
transport mechanisms involved, appropriate surface modification/fabrication techniques
have to be chosen. These modification/fabrication techniques should allow dimensions of
the surface features (in the micro/nano-scale range) to be accurately manufactured. In this
aspect, techniques such as MEMS/NEMS related processes and SLM have significant
advantages over other techniques.
(3) The high nucleation site density of enhanced surfaces is a reason widely suggested for the
observed improvements in boiling performance as compared to plain surfaces. However,
the heat transfer coefficient and CHF are also strongly influenced by surface wettability.
Despite the highly wetting nature of dielectric fluids, previous investigations suggest that
significantly different heat transfer characteristics can also be achieved with enhanced
surfaces of different wettabilities [65, 71]. There are, however, limited published studies in
this area with dielectric fluids and more comparative studies are needed.
(4) Empirical and semi-empirical correlations for predicting the heat transfer and pressure drop
performances of enhanced surfaces during boiling have achieved relative success. On the
other hand, mechanism-based pool and flow boiling models for enhanced surfaces are
limited. A fundamental understanding of the boiling mechanisms as mentioned in (1)-(3)
is, therefore, essential for the development of accurate mechanism-based models.
22
Pre-publication version
K.C. Leong et al. Applied Thermal Engineering 112 (2017) 999–1019
http://dx.doi.org/10.1016/j.applthermaleng.2016.10.138
(5) In order for the enhanced surfaces to be implemented in engineering systems, the reliability
of the surfaces has to be evaluated. The surfaces should be subjected to systematic testing
under extended durations so as to determine the change in boiling characteristics of these
surfaces over time.
Acknowledgements
The authors would like to acknowledge the financial support under Nanyang Technological
University Singapore’s Academic Research Fund (AcRF) Tier 1 Grant No. RG 119/14 for
recent studies on pool and flow boiling of surfaces fabricated by selective laser melting (SLM)
reported in this review. Funding for the SLM facility by National Research Foundation,
Singapore is gratefully acknowledged.
23
Pre-publication version
K.C. Leong et al. Applied Thermal Engineering 112 (2017) 999–1019
http://dx.doi.org/10.1016/j.applthermaleng.2016.10.138
References
[1] Bar-Cohen, A., Holloway, C.A., Thermal science and engineering from macro to nano in 200
years, Proceedings of the 15th International Heat Transfer Conference, IHTC-15, August 10-15,
Kyoto, Japan (2014).
[2] Mudawar, I., Assessment of high-heat-flux thermal management schemes, IEEE Transactions on
Components and Packaging Technologies, Vol. 24, No. 2 (2001).
[3] Mudawar, I., Bharathan, D., Kelly, K., Narumanchi, S., Two-phase spray cooling of hybrid vehicle
electronics, Thermal and Thermomechanical Phenomena in Electronic Systems, ITHERM, 11th
Intersociety Conference, May 28-31, Orlando, USA (2008).
[4] Bar-Cohen, A., Thermal management of electronic components with dielectric liquid, JSME
International Journal Series B, Vol. 36, no. 1, pp. 1-24 (1993).
[5] van Gils, R.W., Danilov, D., Notten, P.H.L., Speetjens, M.F.M., Nijmeijer, H., Battery thermal
management by boiling heat-transfer, Energy Conversion and Management, Vol. 79, pp. 9-17
(2014).
[6] Bar-Cohen, A., Arik, M., Ohadi, M., Direct liquid cooling of high flux micro and nano electronic
components, Proceedings of the IEEE, Vol. 94, No. 8, pp. 1549-1570 (2006).
[7] Kakaç S., Yurucu H., Hijikata K.A., Cooling of Electronic Systems, Kluwer Academic Publishers
(1994).
[8] Ciloglu, D., Bolukbasi, A., A comprehensive review on pool boiling of nanofluids, Applied
Thermal Engineering, Vol. 84, pp. 45-63 (2015).
[9] Fang, X., Wang, R., Chen, W., Zhang, H., Ma, C., A review of flow boiling heat transfer of
nanofluids, Applied Thermal Engineering, Vol. 91, pp. 1003-1017 (2015).
[10] Patil, C.M., Kandlikar, S.G., Review of the manufacturing techniques for porous surfaces used in
enhanced pool boiling, Heat Transfer Engineering, Vol. 35, pp. 887-902 (2014).
[11] Shojaeian, M., Kosar, A., Pool boiling and flow boiling on micro- and nanostructured surfaces,
Experimental Thermal and Fluid Science, Vol. 63, pp. 45-73 (2015).
[12] Kim, D.E., Yu, D.I., Jerng, D.W., Kim, M.H., Ahn, H.S., Review of boiling heat transfer
enhancement on micro/nanostructured surfaces, Experimental Thermal and Fluid Science, Vol. 66,
pp. 173-196 (2015).
[13] 3MTM FluorinertTM Electronic and NovecTM Engineering Liquids product information,
www.3m.com/fluids, 3M Corporation.
[14] Nukiyama, S., The maximum and minimum values of the heat Q transmitted from metal to boiling
water under atmospheric pressure, International Journal of Heat and Mass Transfer, Vol. 9, pp.
1419-1433 (1966).
[15] Rohsenow, W.M., A method of correlating heat transfer data for surface boiling of liquids,
Transactions of ASME, Vol. 74, pp. 969-967 (1952).
[16] Mikic, B.B., Rohsenow, M.W., A new correlation of pool-boiling data including the effect of
heating surface characteristics, Journal of Heat Transfer, Vol. 9, pp. 245-250 (1969).
[17] Cooper, M.G., Lloyd, A.J.P, The microlayer in nucleate boiling, International Journal of Heat and
Mass Transfer, Vol. 12, pp. 895-913 (1969).
[18] Snyder, N.R., Edwards, D.K., Summary of conference on bubble dynamics and boiling heat
transfer, Jet Propulsion Laboratory Memo, 20, pp. 20-137 (1956).
[19] Moore, F.D., Mesler, R.B., The measurement of rapid surface temperature fluctuations during
nucleate boiling of water, AIChE, Vol. 7, pp. 620-624 (1961).
[20] Judd, R.L., Hwang, K.S., A comprehensive model for nucleate boiling heat transfer, including
microlayer evaporation, Journal of Heat Transfer, Vol. 98, pp. 623-629 (1976).
[21] Benjamin, R.J., Balakrishnan, A.R., Nucleate pool boiling heat transfer of pure liquids at low to
moderate heat fluxes, International Journal of Heat and Mass Transfer, Vol. 39, pp. 2495-2504
(1996).
[22] Wayner, P.C., Kao, Y.K., LaCroix, L.V., The interline heat-transfer coefficient of an evaporating
wetting film, International Journal of Heat and Mass Transfer, Vol. 19, pp. 487-492 (1976).
[23] Stephan, P., Hammer, J., A new model for nucleate boiling heat transfer, Wärme- and
Stoffübertragung, Vol. 30, pp. 119-125 (1994).
24
Pre-publication version
K.C. Leong et al. Applied Thermal Engineering 112 (2017) 999–1019
http://dx.doi.org/10.1016/j.applthermaleng.2016.10.138
[24] Kern, J., Stephan, P., Theoretical model for nucleate boiling heat and mass transfer of binary
mixtures, Journal of Heat Transfer, Vol. 125, pp. 1106-1115 (2003).
[25] Zhao, Y.-H., Masuoka, T., Tsuruta, T., Unified theoretical prediction of fully developed nucleate
boiling and critical heat flux based on a dynamic microlayer model, International Journal of Heat
and Mass Transfer, Vol. 45, pp. 3189-3197 (2002).
[26] Das, A.K., Das, P.K., Saha, P., Heat transfer during pool boiling based on evaporation from micro
and macrolayer, International Journal of Heat and Mass Transfer, Vol. 49, pp. 3487-3499 (2006).
[27] Li Y.-Y., Liu, Z.-H., Wang, G.-S., A predictive model of nucleate pool boiling on heated
hydrophilic surfaces, International Journal of Heat and Mass Transfer, Vol. 65, pp. 789-797
(2013).
[28] Son, G., Dhir, V.K., Ramanujapu, N., Dynamics and heat transfer associated with a single bubble
during nucleate boiling on a horizontal surface, Journal of Heat Transfer, Vol. 121, pp. 623-631
(1999).
[29] Son, G., Ramanujapu, N., Dhir, V.K., Numerical simulation of bubble merger process on a single
nucleation site during pool nucleate boiling, Journal of Heat Transfer, Vol. 124, pp. 51-62 (2002).
[30] Kunkelmann, C., Stephan, P., Numerical simulation of the transient heat transfer during nucleate
boiling of refrigerant HFE-7100, International Journal of Refrigeration, Vol. 33, pp.1221-1228
(2010).
[31] Jia, H.W., Zhang, P., Fu, X., Jiang, S.C., A numerical investigation of nucleate boiling at a constant
surface temperature, Applied Thermal Engineering, Vol. 88, pp. 248-257 (2015).
[32] Gong, S., Cheng, P., Lattice Boltzmann simulation of periodic bubble nucleation, growth and
departure from a heated surface in pool boiling, International Journal of Heat and Mass Transfer,
Vol. 64, pp. 122-132 (2013).
[33] Gong, S., Cheng, P., Numerical simulation of pool boiling heat transfer on smooth surfaces with
mixed wettability by lattice Boltzmann method, International Journal of Heat and Mass Transfer,
Vol. 80, pp. 206-216 (2015).
[34] Demiray, F., Kim, J., Microscale heat transfer measurements during pool boiling of FC-72: effect
of subcooling, International Journal of Heat and Mass Transfer, Vol. 47, pp. 3257-3268 (2004).
[35] Myers, J.G., Yerramilli, V.K., Hussey, S.W., Yee, G.F., Kim, J., Time and space resolved wall
temperature and heat flux measurements during nucleate boiling with constant heat flux boundary
conditions, International Journal of Heat and Mass Transfer, Vol. 48, pp. 2429-2442 (2005).
[36] Wagner, E., Stephan, P., High-resolution measurements at nucleate boiling of pure FC-84 and FC-
3284 and its binary mixtures, Journal of Heat Transfer, Vol. 131, pp. 121008-1 - 121008-12
(2009).
[37] Moghaddam, S., Kiger, K., Physical mechanisms of heat transfer during single bubble nucleate
boiling of FC-72 under saturation conditions-I. Experiment investigation, International Journal of
Heat and Mass Transfer, Vol. 52, pp. 1284-1294 (2009).
[38] Moghaddam, S., Kiger, K., Physical mechanisms of heat transfer during single bubble nucleate
boiling of FC-72 under saturation conditions-II. Theoretical analysis, International Journal of Heat
and Mass Transfer, Vol. 52, pp. 1295-1303 (2009).
[39] Chen, T., Chung, J.N., Coalescence of bubbles in nucleate boiling on microheaters, International
Journal of Heat and Mass Transfer, Vol. 45, pp. 2329-2341 (2002).
[40] Seidel, S., Cioulachtjian, S., Bonjour, J., Experimental analysis of bubble growth, departure and
interactions during pool boiling on artificial nucleation sites, Experimental Thermal and Fluid
Science, Vol. 32, pp. 1504-1511 (2008).
[41] Bi, J., Lin, X., Chistopher, D.M., Effects of bubble coalescence dynamics on heat flux distributions
under bubbles, AIChE, Vol. 59, pp. 1735-1745 (2013).
[42] Ramaswamy, C., Joshi, Y., Nakayama, W., Johnson, W.B., High-speed visualization of boiling
from an enhanced surface, International Journal of Heat and Mass Transfer, Vol. 45, pp. 4761-
4771 (2002).
[43] Ghiu, C-D., Joshi, Y.K., Visualization study of pool boiling from thin confined enhanced
structures, International Journal of Heat and Mass Transfer, Vol. 45, pp. 4287-4299 (2005).
[44] Teodori, E., Moita, A.S., Moreira, A.L.N., Study of the combined effects of liquid properties and
surface micropatterning on pool boiling heat transfer, Proceedings of the 15th International Heat
Transfer Conference, IHTC-15, August 10-15, Kyoto, Japan (2014).
25
Pre-publication version
K.C. Leong et al. Applied Thermal Engineering 112 (2017) 999–1019
http://dx.doi.org/10.1016/j.applthermaleng.2016.10.138
[45] Moita, A.S., Teodori, E., Moreria, A.L.N., Influence of surface topography in the boiling
mechanisms, International Journal of Heat and Fluid Flow, Vol. 52, pp. 50-63 (2015).
[46] Zuber, N., Hydrodynamic Aspects of Boiling Heat Transfer, AEC Report No. AECU-4439 (1959).
[47] Gaertner, R.F., Photographic study of nucleate pool boiling on a horizontal surface, Journal of
Heat Transfer, Vol. 87, pp. 17-29 (1965).
[48] Haramura,Y., Katto, Y., A new hydrodynamic model of critical heat flux applicable widely to both
pool and forced convection boiling on submerged bodies in saturated liquids, International Journal
of Heat and Mass Transfer, Vol. 3, pp. 389-399 (1983).
[49] El-Genk, M., Parker, J.L., Nucleate boiling of FC-72 and HFE-7100 on porous graphite at different
orientations and liquid subcooling, Energy Conversion and Management, Vol. 49, pp. 733-750
(2008).
[50] Ho, J.Y., Leong, K.C., Yang, C., Saturated pool boiling from carbon nanotube surfaces at different
orientations, International Journal of Heat and Mass Transfer, Vol. 79, pp. 893-904 (2014).
[51] Arik, M., Bar-Cohen, A., Effusivity-based correlation of surface property effects in pool boiling
CHF of dielectric liquids, International Journal of Heat and Mass Transfer, Vol. 46, pp. 3755-
3764 (2003).
[52] Kandlikar, S.G., A theoretical model to predict pool boiling CHF incorporating effects of contact
angle and orientation, Journal of Heat Transfer, Vol. 123, pp. 1071-1079 (2001).
[53] Kim, H.D., Kim, M.H., Effect of nanoparticle deposition on capillary wicking that influences the
critical heat flux in nanofluids, Applied Physics Letters, Vol. 91, pp. 014104 (2007).
[54] Ahn, H.S, Lee, C., Kim, J., Kim, M.H., The effect of capillary wicking action of micro/ nano
structures on pool boiling critical heat flux, International Journal of Heat and Mass Transfer, Vol.
55, pp. 89-92 (2012).
[55] Chu, K-H., Enright, R., Wang, E.N., Structured surfaces for enhanced pool boiling heat transfer,
Applied Physics Letters, Vol. 100, pp. 241603 (2012).
[56] Wenzel, R.N., Resistance of solid surface to wetting by water, Industrial and Engineering
Chemistry, Vol. 28, pp. 988-994 (1936).
[57] Zou, A., Maroo, S.C., Critical height of micro/nano structures for pool boiling heat transfer
enhancement, Applied Physics Letters, Vol. 103, pp. 221602 (2013).
[58] Rahman, M.M., Ölçeroğlu,, E., McCarthy, M., Role of wickability on the critical heat flux of
structured superhydrophilic surface, Langmuir, Vol. 30, pp. 11225-11234 (2014).
[59] Quan, X., Dong, L., Cheng, P., A CHF model for saturated pool boiling on a heated surface with
micro/nano-scale structures, International Journal of Heat and Mass Transfer, Vol. 76, pp. 452-
458 (2014).
[60] Rainey, K.N., You, S.M., Effect of heater size and orientation on pool boiling heat transfer from
microporous coated surfaces, International Journal of Heat and Mass Transfer, Vol. 44, pp. 2589-
2599 (2001).
[61] Chang, J.Y., You, S.M., Boiling heat transfer phenomena from micro-porous and porous surfaces
in saturated FC-72, International Journal of Heat and Mass Transfer, Vol. 40, pp. 4437-4447
(1997).
[62] Kim, J.H., Rainey, K.N., You, S.M., Pak, J.Y., Mechanism of nucleate boiling heat transfer
enhancement from microscopic surfaces in saturated FC-72, Journal of Heat Transfer, Vol. 124,
pp. 500-506 (2002).
[63] Arik, M., Bar-Cohen, A., You, S.M., Enhancement of pool boiling critical heat flux in dielectric
liqiuds by microscopic coatings, International Journal of Heat and Mass Transfer, Vol. 50, pp.
997-1009 (2007).
[64] Takata, Y., Hidaka, S., Cao, J.M., Nakamura, T., Yamamoto, H., Masuda, M., Ito, T., Effect of
surface wettability on boiling and evaporation, Energy, Vol. 30, pp. 209-220 (2005).
[65] Wu, W., Bostanci, H., Chow, L.C., Hong, Y., Su, M., Kizito, J.P., Nucleate boiling heat transfer
enhancement for water and FC-72 on titanium oxide and silicon oxide surfaces, International
Journal of Heat and Mass Transfer, Vol. 53, pp. 1773-1777 (2010).
[66] Dewangan, A.K., Kumar, A., Kumar, R., Nucleate boiling of pure and quasi-azeotropic refrigerants
from copper coated surfaces, Applied Thermal Engineering, Vol. 94, pp. 395-403 (2016).
[67] Pialago, E.J.T., Kwon, O.K., Park, C.W., Nucleate boiling heat transfer of R134a on cold sprayed
CNT-Cu composite coatings, Applied Thermal Engineering, Vol. 56, pp. 112-119 (2013).
26
Pre-publication version
K.C. Leong et al. Applied Thermal Engineering 112 (2017) 999–1019
http://dx.doi.org/10.1016/j.applthermaleng.2016.10.138
[68] Pialago, E.J.T., Kwon, O.K., Jin, J.S., Park, C.W., Nucleate pool boiling of R134a on cold sprayed
Cu-CNT-SiC and Cu-CNT-AIN composite coatings, Applied Thermal Engineering, Vol. 103, pp.
648-694 (2016).
[69] Sahu, R.P., Sinha-Ray, S., Sinha-Ray, S., Yarin, A.L., Pool boiling on nano-tectured surfaces
comprised of electrically-assisted supersonically soluation-blown, copper-plated nanofibers:
Experiments and theory, International Journal of Heat and Mass Transfer, Vol. 87, pp. 521-535
(2015).
[70] Sahu, R.P., Sinha-Ray, S., Sinha-Ray, S., Yarin, A.L., Pool boiling of Novec 7300 and self-
rewetting fluids on electrically-assisted supersonically soluation-blown, copper-plated nanofibers,
Intenrational Journal of Heat and Mass Transfer, Vol. 95, pp. 83-93 (2016).
[71] An, S., Kim, D-Y., Lee, J-G., Jo, H.S., Kim, M.W., Al-Deyab, S.S., Choi, J., Yoon S.S.,
Supersonically sprayed reduced graphene oxide film to enhance critical heat flux in pool boiling,
International Journal of Heat and Mass Transfer, Vol. 98, pp. 124-130 (2016).
[72] Kim, T.Y., Weibel, J.A., Garimella, S.V., A free-particles-based technique for boiling heat transfer
enhancement in a wetting liquid, International Journal of Heat and Mass Transfer,Vol. 71, pp.
808-817 (2014).
[73] Sarangi, S., Weibel, J.A., Garimella, S.V., Effect of particle size on surface-coating enhancement
of pool boiling heat transfer, International Journal of Heat and Mass Transfer, Vol. 81, pp. 103-
113 (2015).
[74] Thiagarajan, S.J., Yang, R., King, C., Narumanchi, S., Bubble dynamics and nucleate pool boiling
heat transfer on microporous copper surfaces, International Journal of Heat and Mass Transfer,
Vol. 89, pp. 1297-1315 (2015).
[75] El-Genk, M.S., Ali, A.F., Enhanced nucleate boiling on copper micro-porous surfaces,
International Journal of Multiphase Flow, Vol. 36, pp. 780-792 (2010).
[76] El-Genk, M.S., Ali, A.F., Enhancement of saturation boiling of PF-5060 on microporous copper
dendrite surfaces, Journal of Heat Transfer, Vol. 132, pp. 071501-1 - 071501-9 (2010).
[77] Ali, A.F., El-Genk, M.S., Spreaders of immersion nucleate boiling cooling of a computer chip with
a central hot spot, Energy Conversion and Management, Vol. 53, pp. 259-267 (2012).
[78] Ali, A.F., El-Genk, M.S., Effect of inclination on saturation boiling of PF-5060 dielectric liquid
on 80- and 137-µ thick copper micro-porous surfaces, International Jouranl of Thermal Sciences,
Vol. 53, pp. 42-48 (2012).
[79] Furberg, R., Palm, B., Boiling heat transfer on a dendritic and micro-porous surface in R134a and
FC-72, Applied Thermal Engineering, Vol. 31, pp. 3595-3603 (2011).
[80] Ujereh, S., Fisher, T., Mudawar, I., Effects of carbon nanotubes arrays on nucleate pool boiling,
International Journal of Heat and Mass Transfer, Vol. 50, pp. 4023-4038 (2007).
[81] Sathyamurthi, V., Ahn, H-S., Banerjee, D., Lau, S.C., Subcooled pool boiling experiments on
horizontal heaters coated with carbon nanotubes, Journal of Heat Transfer, Vol. 131, pp. 071501-
1 - 071501-10 (2009).
[82] Guglielmini, G., Misale, M., Schenone, C., Boiling of saturated FC-72 on square pin fin arrays,
International Journal of Thermal Sciences, Vol. 41, pp. 599-608 (2002).
[83] Yu, C.K., Lu, D.C., Pool boiling heat transfer on horizontal rectangular fin array in saturated FC-
72, International Journal of Heat and Mass Transfer, Vol. 50, pp. 3624-3637 (2007).
[84] Jones, B.J., McHale, J.P., Garimella, S.V., The influence of surface roughness on nucleate pool
boiling heat transfer, Journal of Heat Transfer, Vol 131, pp. 121009-1 - 121009-14 (2009).
[85] McHale, J.P., Garimella, S.V., Bubble nucleation characterisitics in pool boiling of a wetting liquid
on smooth and rough surfaces, International Journal of Multiphase Flow, Vol 36, pp. 249-260
(2009).
[86] El-Genk, M.S., Suszko, A., Saturated nucleate boiling and correlations for PF-5060 dielectric
liquid on inclined rough copper surfaces, Journal of Heat Transfer, Vol 136, pp. 081503-1 -
081503-10 (2014).
[87] Suszko, A., El-Genk, M.S., Saturation boiling of PF-5060 on rough Cu surfaces: Bubbles transient
growth, departure diameter and detachment frequency, International Journal of Heat and Mass
Transfer, Vol 91, pp. 363-373 (2015).
27
Pre-publication version
K.C. Leong et al. Applied Thermal Engineering 112 (2017) 999–1019
http://dx.doi.org/10.1016/j.applthermaleng.2016.10.138
[88] Yu, C.K., Lu, D.C., Cheng, T.C., Pool boiling heat transfer on artificial micro-cavity surfaces in
dielectric fluid FC-72, Journal of Micromechanics and Microengineering, Vol. 16, pp. 2092-2099
(2006).
[89] Honda, H., Takamatsu, H., Wei, J.J., Enhanced boiling of FC-72 on silicon chips with micro-pin-
fins and submicron-scale roughness, Journal of Heat Transfer, Vol. 124, pp. 383-390 (2002).
[90] Wei, J.J., Guo, L.J., Honda, H., Experimental study of boiling phenomena and heat transfer
performances of FC-72 over micro-pin-finned silicon chips, Heat and Mass Transfer, Vol. 41, pp.
744-755 (2005).
[91] Xue, Y-F., Zhao, J-F., Wei, J-J., Zhang, Y-H., Qi, B-J., Experimental study of nucleate pool boiling
of FC-72 on micro-pin-finned surface under microgravity, International Journal of Heat and Mass
Transfer, Vol. 63, pp. 425-433 (2013).
[92] Zhang, Y., Wei, J., Xue, Y., Kong, X., Zhao, J., Bubble dynamics in nucleate pool boiling on
micro-pin-finned surfaces in microgravity, Applied Thermal Engineering, Vol. 70, pp. 172-182
(2014).
[93] Gallego, N.C., Klett, J.W., Carbon foams for thermal management, Carbon, Vol. 41, pp. 1461-
1466 (2003).
[94] Leong, K.C., Jin, L.W., Effect of oscillatory frequency on heat transfer in metal foam heat sinks
of various pore densities, International Journal of Heat and Mass Transfer, Vol. 49, pp. 671-681
(2006).
[95] Xu, Z.G., Zhao, C.Y., Experimental study on pool boiling heat transfer in gradient metal foams,
International Journal of Heat and Mass Transfer, Vol. 85, pp. 824-829 (2015).
[96] Xu, Z.G., Zhao, C.Y., Enhanced boiling heat transfer by gradient porous metals in saturated pure
water and surfactant solutions, Applied Thermal Engineering, Vol. 100, pp. 68-77 (2016).
[97] Leong, K.C., Jin, L.W., Pranoto, I., Li, H.Y., Chai, J.C., Experimental study of enhanced pool
boiling heat transfer using graphite foam inserts, Defect and Diffusion Forum, Vol. 312-315, pp.
352-357 (2011).
[98] Jin, L.W., Leong, K.C., Pranoto, I., Saturated pool boiling heat transfer from highly conductive
graphite foams, Applied Thermal Engineering, Vol. 31, pp. 2685-2693 (2011).
[99] Pranoto, I., Leong, K.C., Jin, L.W., The role of graphite foam pore structure on saturated pool
boiling enhancement, Applied Thermal Engineering, Vol. 42, pp. 163-172 (2012).
[100] El-Genk, M.S., Parker, J.L., Enhanced boiling of HFE-7100 dielectric liquid on porous graphite,
Energy Conversion and Management, Vol. 46, pp. 2455-2481 (2005).
[101] El-Genk, M.S., Immersion cooling nucleate boiling of high power computer chips, Energy
Conversion and Management, Vol. 53, pp. 205-218 (2012).
[102] Wong, M., Owen, I., Sutcliffe, C.J., Puri, A., Convective heat transfer and pressure losses across
novel heat sinks fabricated by selective laser melting, International Journal of Heat and Mass
Transfer, Vol. 52, pp. 281-288 (2009).
[103] Ventola, L., Robotti, F., Dialameh, M., Calignano, F., Manfredi, D., Chiavazo, E., Asinari, P.,
Rough surfaces with enhanced heat transfer of electronic cooling by direct laser sintering,
International Journal of Heat and Mass Transfer, Vol. 75, pp. 58-74 (2014).
[104] Ameli, M., Agnew, B., Leung, P.S., Ng, B., Sutcliffe, C.J., Singh, J., McGlen, R., A novel method
for manufacturing sintered aluminum heat pipes (SAHP), Applied Thermal Engineering, Vol. 52,
pp. 498-504 (2013).
[105] Ho, J.Y., Wong, K.K., Leong, K.C., Yang, C., Enhanced nucleate pool boiling from
microstructured surfaces fabricated by selective laser melting, ASME 2016 5th Micro/Nanoscale
Heat and Mass Transfer International Conference, MNHMT-2016, January 3-6, Biopolis,
Singapore (2016).
[106] Ho, J.Y., Wong, K.K., Leong, K.C., Saturated pool boiling of FC-72 from enhanced surfaces
produced by selective laser melting, International Journal of Heat and Mass Transfer, Vol. 99, pp.
107-121 (2016).
[107] Wong, K.K., Leong, K.C., Pool boiling enhancement of porous structures fabricated by selective
laser melting, Proceedings of the International Symposium of Heat Transfer and Heat Powered
Cycles, IHTS-2016, 27-29 June, Nottingham, United Kingdom (2016).
28
Pre-publication version
K.C. Leong et al. Applied Thermal Engineering 112 (2017) 999–1019
http://dx.doi.org/10.1016/j.applthermaleng.2016.10.138
[108] McHale, J.P., Garimella, S.V., Fisher, T.S., Powell, G.A., Pool boiling performance comparison
of smooth and sintered copper surfaces with and without carbon nanotubes, Nanoscale and
Microscale Thermophysical Engineering, Vol. 15, pp. 133-150 (2011).
[109] Im, Y., Dietz, C., Lee, S.S., Joshi, Y., Flower-like CuO nanostructures for enhanced boiling,
Nanoscale and Microscale Thermophysical Engineering, Vol. 16, pp. 145-153 (2012).
[110] Patil, C.M., Kandlikar, S.G, Pool boiling enhancement through microporous coatings selectively
electrodeposited on fin tops of open microchannels, International Journal of Heat and Mass
Transfer, Vol. 79, pp. 816-828 (2014).
[111] Jaikumar, A., Kandlikar, S.G., Enhanced pool boiling for electronics cooling using porous fin tops
on open microchannels with FC-78, Applied Thermal Engineering, Vol. 91, pp. 426-433 (2015).
[112] Collier, J.G., Thome, J.R., Convective boiling and condensation, Third Edition, Oxford University
Press, Oxford (1994).
[113] Chen, J.C., Correlation for boiling heat transfer to saturated fluids in convective flow, Industrial
& Engineering Chemistry Process Design and Development, Vol. 5, pp. 322-329 (1966).
[114] Mohammed Shah, M., Generalized prediction of heat transfer during subcooled boiling in annuli,
Heat Transfer Engineering, Vol. 4, pp. 24-31 (1983).
[115] Liu, Z., Winterton, R.H.S., A general correlation for saturated and subcooled flow boiling in tubes
and annuli, based on a nucleate pool boiling equation, International Journal of Heat and Mass
Transfer, Vol. 34, pp. 2759-2766 (1991).
[116] Kandlikar, S.G., A general correlation for saturated two-phase flow boiling heat transfer inside
horizontal and vertical tubes, Journal of Heat Transfer, Vol. 112, pp. 219-228 (1990).
[117] Thome, J.R., Cheng, L., Ribatski, G., Vales, L.F., Flow boiling of ammonia and hydrocarbons: a
state-of-the-art review, International Journal of Refrigeration, Vol. 31, pp. 603-620 (2008).
[118] Thome, J.R., Boiling in microchannels: a review of experiment and theory, International Journal
of Heat and Fluid Flow, Vol. 25, pp. 128-139 (2004).
[119] Kandlikar, S.G., History, advances, and challenges in liquid flow and flow boiling heat transfer in
microchannels: a critical review, Journal of Heat Transfer, Vol. 134, pp. 034001-1 - 034001-15
(2012).
[120] Bertsch, S.S., Groll, E.A., Garimella, S.V., Review and comparative analysis of studies on
saturated flow boiling in small channels, Nanoscale and Microscale Thermophysical Engineering,
Vol. 12, pp. 187-227 (2008).
[121] Kandlikar, S.G., Grande, W.J., Evolution of microchannel flow passages-thermohydraulic
performance and fabrication technology, Heat Transfer Engineering, Vol. 24, pp. 3-17 (2003).
[122] Kew, P.A., Cornwell, K., Correlations for the prediction of boiling heat transfer in small-diameter
channels, Applied Thermal Engineering, Vol. 17, pp. 705-715 (1997).
[123] Ong, C.L., Thome, J.R., Macro-to-microchannel transition in two-phase flow: Part 1–Two-phase
flow patterns and film thickness measurements, Experimental Thermal and Fluid Science, Vol. 35,
pp. 37-47 (2011).
[124] Tibirica, C.B., Ribatski, G., Flow boiling in micro-scale channels–synthesized literature review,
International Journal of Refrigeration, Vol. 36, pp. 301-324 (2013).
[125] Harirchian, T., Garimella, S.V., A comprehensive flow regime map for microchannel flow boiling
with quantitative transition criteria, International Journal of Heat and Mass Transfer, Vol. 53, pp.
2694-2702 (2010).
[126] Kandlikar, S.G., Scale effects on flow boiling heat transfer in microchannels: A fundamental
perspective, International Journal of Thermal Sciences, Vol. 49, pp. 1073-1085 (2010).
[127] Wang, Y., Sefiane, K., Harmand, S., Flow boiling in high-aspect ratio mini-and micro-channels
with FC-72 and ethanol: Experimental results and heat transfer correlation assessments,
Experimental Thermal and Fluid Science, Vol. 36, pp. 93-106 (2010).
[128] Hetsroni, G., Mosyak, A., Pogrebnyak, E., Segal Z., Explosive boiling of water in parallel micro-
channels, International Journal of Multiphase Flow, Vol. 31, pp. 371–392 (2005).
[129] Wang, G., Cheng, P., Wu, H., Unstable and stable flow boiling in parallel microchannels and in a
single microchannel, International Journal of Heat and Mass Transfer, Vol. 50, pp. 4297–4310
(2007).
29
Pre-publication version
K.C. Leong et al. Applied Thermal Engineering 112 (2017) 999–1019
http://dx.doi.org/10.1016/j.applthermaleng.2016.10.138
[130] Bogojevic, D., Sefiane, K., Walton, A.J., Lin, H., Cummins, G., Two-phase flow instabilities in a
silicon microchannels heat sink, International Journal of Heat and Fluid Flow, Vol. 30, pp. 854–
867 (2009).
[131] Kosar, A., Kuo, C.-J., Peles, Y., Suppression of boiling flow oscillations in parallel microchannels
by inlet restrictors, Journal of Heat Transfer, Vol. 128, p. 251-260 (2006).
[132] Kandlikar, S.G., Kuan, W.K., Willistein, D.A., Borrelli, J., Stabilization of flow boiling in
microchannels using pressure drop elements and fabricated nucleation sites, Journal of Heat
Transfer, Vol. 128, pp. 389–396 (2006).
[133] Kim, S.M., Mudawar, I., Universal approach to predicting two-phase frictional pressure drop for
mini/micro-channel saturated flow boiling, International Journal of Heat and Mass Transfer, Vol.
58, pp. 718-734 (2013).
[134] Harirchian, T., Garimella, S.V., Microchannel size effects on local flow boiling heat transfer to a
dielectric fluid, International Journal of Heat Mass Transfer, Vol. 51, pp. 3724–3735 (2008).
[135] Harirchian, T., Garimella, S.V., The critical role of channel cross-sectional area in microchannel
flow boiling heat transfer, International Journal of Multiphase Flow, Vol. 35, pp. 904-913 (2009).
[136] Ma, A., Wei, J., Yuan, M., Fang, J., Enhanced flow boiling heat transfer of FC-72 on micro-pin-
finned surfaces, International Journal of Heat and Mass Transfer, Vol. 52, pp. 2925-2931 (2009).
[137] Wei, J., Zhao, J., Yuan, M., Xue, Y., Boiling heat transfer enhancement by using micro-pin-finned
surface for electronics cooling, Microgravity Science and Technology, Vol. 21, pp. 159-173
(2009).
[138] Guo, D., Wei, J.J., Zhang, Y.H., Enhanced flow boiling heat transfer with jet impingement on
micro-pin-finned surfaces, Applied Thermal Engineering, Vol. 31, pp. 2042-2051 (2011).
[139] Krishnamurthy, S., Peles, Y., Flow boiling heat transfer on micro pin fins entrenched in a
microchannel, Journal of Heat Transfer, Vol. 132, pp. 041007-1 - 041007-10 (2010).
[140] Rainey, K.N., Li, G., You, S.M., Flow boiling heat transfer from plain and microporous coated
surfaces in subcooled FC-72, Journal of Heat Transfer, Vol. 123, pp. 918-925 (2001).
[141] Ammerman, C.N., You, S.M., Enhancing small-channel convective boiling performance using a
microporous surface coating. Journal of Heat Transfer, Vol. 123, pp. 976-983 (2001).
[142] Sun, Y., Zhang, L., Xu, H., Zhong, X., Flow boiling enhancement of FC-72 from microporous
surfaces in minichannels, Experimental Thermal and Fluid Science, Vol. 35, pp. 1418-1426
(2011).
[143] Bai, P., Tang, T., Tang, B., Enhanced flow boiling in parallel microchannels with metallic porous
coating. Applied Thermal Engineering, Vol. 58, pp. 291-297 (2013).
[144] Kim, D.W., Bar-Cohen, A., and Han, B. Forced convection and flow boiling of a dielectric liquid
in a foam-filled channel, 2008 11th Intersociety Conference on Thermal and Thermomechanical
Phenomena in Electronic Systems, ITHERM 2008, 28-31 May, Orlando, USA (2008).
[145] Lu, W., Zhao, C.Y., Numerical modelling of flow boiling heat transfer in horizontal metal-foam
tubes, Advanced Engineering Materials, Vol. 11, pp. 832-836 (2009).
[146] Pranoto, I., Leong, K.C., An experimental study of flow boiling heat transfer from porous foam
structures in a channel, Applied Thermal Engineering, Vol. 70, pp. 100-114 (2014).
[147] Li, H.Y., Leong, K.C., Experimental and numerical study of single and two-phase flow and heat
transfer in aluminum foams, International Journal of Heat and Mass Transfer, Vol. 54, pp. 4904-
4912 (2011).
[148] Madani, B., Tadrist, L., Topin, F., Experimental analysis of upward flow boiling heat transfer in a
channel provided with copper metallic foam, Applied Thermal Engineering, Vol. 52, pp. 336-344
(2013).
[149] Gungor, K.E., Winterton, R.H.S., Simplified general correlation for saturated flow boiling and
comparisons of correlations with data, Chemical Engineering Research and Design, Vol. 65, pp.
148-165 (1987).
30
Pre-publication version
K.C. Leong et al. Applied Thermal Engineering 112 (2017) 999–1019
http://dx.doi.org/10.1016/j.applthermaleng.2016.10.138
105
Transition
104 Boiling boiling
incipience Minimum
heat flux
103
1 10 100 1000
∆T (ºC)
31
Pre-publication version
K.C. Leong et al. Applied Thermal Engineering 112 (2017) 999–1019
http://dx.doi.org/10.1016/j.applthermaleng.2016.10.138
32
Pre-publication version
K.C. Leong et al. Applied Thermal Engineering 112 (2017) 999–1019
http://dx.doi.org/10.1016/j.applthermaleng.2016.10.138
Fig. 3 Photograph of micro-heaters array with each heater nominally 0.1 mm2 in area [34].
33
Pre-publication version
K.C. Leong et al. Applied Thermal Engineering 112 (2017) 999–1019
http://dx.doi.org/10.1016/j.applthermaleng.2016.10.138
Fig. 4 Bubble ebullition cycle in millisecond time scale and 100 µm length scale [37].
34
Pre-publication version
K.C. Leong et al. Applied Thermal Engineering 112 (2017) 999–1019
http://dx.doi.org/10.1016/j.applthermaleng.2016.10.138
Fig. 5 Wave-front propagation (indicated in vertical white line) during lateral bubble
coalescence [40].
35
Pre-publication version
K.C. Leong et al. Applied Thermal Engineering 112 (2017) 999–1019
http://dx.doi.org/10.1016/j.applthermaleng.2016.10.138
Fig. 6 Microscope images at 10 µm scale (a, b) and SEM images at 1 µm (c, d) of copper-
plated nanofiber [69].
36
Pre-publication version
K.C. Leong et al. Applied Thermal Engineering 112 (2017) 999–1019
http://dx.doi.org/10.1016/j.applthermaleng.2016.10.138
Fig. 7 Photographs of surface produced by particle sintering with particle size of 850-1000
µm (a) top view (b) side view [73].
37
Pre-publication version
K.C. Leong et al. Applied Thermal Engineering 112 (2017) 999–1019
http://dx.doi.org/10.1016/j.applthermaleng.2016.10.138
(a) (b)
Fig. 8 SEM images of microporous surface at (a) 150 µm and (b) 100 µm scales [78].
38
Pre-publication version
K.C. Leong et al. Applied Thermal Engineering 112 (2017) 999–1019
http://dx.doi.org/10.1016/j.applthermaleng.2016.10.138
(a) (b)
Fig. 9 (a) Photographs of fully coated CNT surface produced by CVD (b) SEM image of
vertically-aligned CNT array [50].
39
Pre-publication version
K.C. Leong et al. Applied Thermal Engineering 112 (2017) 999–1019
http://dx.doi.org/10.1016/j.applthermaleng.2016.10.138
40
35
Wu et al. [65]
25
An et al. [71]
q'' (W/cm2)
Fig. 10 Comparison of pool boiling curves of the better performing coated surfaces
reproduced from selected publications.
40
Pre-publication version
K.C. Leong et al. Applied Thermal Engineering 112 (2017) 999–1019
http://dx.doi.org/10.1016/j.applthermaleng.2016.10.138
(a) (b)
Fig. 11 (a) Scanning electron microscope images of graphite foam internal structure [98] (b)
finned porous graphite [99].
41
Pre-publication version
K.C. Leong et al. Applied Thermal Engineering 112 (2017) 999–1019
http://dx.doi.org/10.1016/j.applthermaleng.2016.10.138
Fig. 12 (a) Unit of Octet-truss and (b) porous structures of height range 2.5 mm to 10 mm
fabricated by Wong et al. [107].
42
Pre-publication version
K.C. Leong et al. Applied Thermal Engineering 112 (2017) 999–1019
http://dx.doi.org/10.1016/j.applthermaleng.2016.10.138
Fig. 13 Liquid-vapor pathways on a hybrid surface as proposed by Patil and Kandlikar [110].
43
Pre-publication version
K.C. Leong et al. Applied Thermal Engineering 112 (2017) 999–1019
http://dx.doi.org/10.1016/j.applthermaleng.2016.10.138
50
45
40
35
30
FC-77 as working fluid:
q'' (W/cm2)
10 Ho et al. [106]
HFE-7100 as working fluid:
El-Genk and Parker [100]
5
0
0 10 20 30 40 50 60
∆T (°C)
Fig. 14 Comparison of pool boiling curves of the better performing intrinsic surfaces
reproduced from selected publications.
44
Pre-publication version
K.C. Leong et al. Applied Thermal Engineering 112 (2017) 999–1019
http://dx.doi.org/10.1016/j.applthermaleng.2016.10.138
45
Pre-publication version
K.C. Leong et al. Applied Thermal Engineering 112 (2017) 999–1019
http://dx.doi.org/10.1016/j.applthermaleng.2016.10.138
46
Pre-publication version
K.C. Leong et al. Applied Thermal Engineering 112 (2017) 999–1019
http://dx.doi.org/10.1016/j.applthermaleng.2016.10.138
47
Pre-publication version
K.C. Leong et al. Applied Thermal Engineering 112 (2017) 999–1019
http://dx.doi.org/10.1016/j.applthermaleng.2016.10.138
Fig. 18 Flow regimes of FC-72 boiling in the channel with dh = 1454 µm, (a) slug and
annular flow, camera speed: 1500 fps, (b) wispy-annular flow at G = 22.4 kg/m2·s and q" =
8.13 kW/m2, camera speed: 1000 fps [127].
48
Pre-publication version
K.C. Leong et al. Applied Thermal Engineering 112 (2017) 999–1019
http://dx.doi.org/10.1016/j.applthermaleng.2016.10.138
Table 1 Thermophysical properties of dielectric fluids [13] and water at 1 atm and 25ºC
Liquid specific heat (J/ 1100 1100 1300 1183 1140 4182
kg⋅K)
Liquid thermal conductivity 0.057 0.056 0.075 0.069 0.063 0.61
(W/m⋅K)
Latent heat of vaporization 88 103 142 112 102 2257
(kJ/kg)
Liquid surface tension 10.0 9.0 12.4 13.6 15.0 72.0
(mN/m)
49
Pre-publication version
K.C. Leong et al. Applied Thermal Engineering 112 (2017) 999–1019
http://dx.doi.org/10.1016/j.applthermaleng.2016.10.138
Epoxy binding
Rainey and - Diamond particles of - Pool boiling at FC-72 - More than 300%
You [60] 8-12 µm in size coated atmospheric enhancement in h at
onto copper substrate. pressure 0° surface
- Boiling surface orientation
dimensions of 2 - 64% enhancement
× 2 and 5 × 5 cm in CHF (or
- Boiling surface 21.7W/cm2) at 0°
orientation 0º- surface orientation
180º
Arik et al. -Diamond particles of - Saturated and FC-72 - 60% enhancement
[63] 8-12 µm in size coated 0-72 K in CHF (or 47
onto copper substrate. subcooled pool W/cm2)
- Coating thickness of boiling
50-75 µm. - System
pressure ranged
from 101.3 kPa
to 303.9 kPa
Dipping and dripping process
Spray coating
50
Pre-publication version
K.C. Leong et al. Applied Thermal Engineering 112 (2017) 999–1019
http://dx.doi.org/10.1016/j.applthermaleng.2016.10.138
Kim et al. -Copper free particles - Pool boiling FC-72 - Max. 76.3%
[72] of sizes ranging from at atmospheric enhancement in
20 nm - 9 mm placed pressure average h
on polished copper - Max. 10%
surface. enhancement in
CHF (or 160
kW/m2)
Sarangi et -Copper free and - Pool boiling FC-72 -Free particles: 32%
al. [73] sintered particles of at atmospheric reduction in wall
sizes ranging from 45 - pressure superheat; 44%
1000 µm coated onto higher CHF (or 161
polished copper kW/m2)
surface. -Sintered particles:
95% reduction in
wall superheat; 33%
reduction in CHF
Thiagarajan - Copper-rich - Saturated pool HFE- -Max. 270%
et al. [74] microparticles boiling at 7100 enhancement in h
diameters 5 - 20 µm. atmospheric -Max. 60%
- Coating thickness 100 pressure enhancement in
µm, 360 µm and 700 CHF (or ≈ 45
µm on plain copper W/cm2)
surface.
51
Pre-publication version
K.C. Leong et al. Applied Thermal Engineering 112 (2017) 999–1019
http://dx.doi.org/10.1016/j.applthermaleng.2016.10.138
Electrochemical deposition
52
Pre-publication version
K.C. Leong et al. Applied Thermal Engineering 112 (2017) 999–1019
http://dx.doi.org/10.1016/j.applthermaleng.2016.10.138
53
Pre-publication version
K.C. Leong et al. Applied Thermal Engineering 112 (2017) 999–1019
http://dx.doi.org/10.1016/j.applthermaleng.2016.10.138
Guglielmini et - Array of copper pin fins with square cross- - Saturated pool FC-72 -
al. [82] section. boiling at 0.5, 1.0
- Pin fin width 0.4 to 1.0 mm. and 2.0 bar
- Pin fin length 3 and 6 mm.
- Pin fin spacing 1.2 and 3.0 mm.
Yu and Lu [83] - Array of copper pin fins with square cross- - Saturated pool FC-72 -
section. boiling at h
- Pin fin width 1.0 mm. atmospheric pressure -
- Pin fin length 0.5, 1.0, 2.0 and 4.0 mm. C
- Pin fin spacing 0.5, 1.0 and 2.0 mm.
Jones et al. [84] - Roughened surfaces with averaged surface - Saturated pool FC-77 and -
roughness of 1.08, 2.22, 5.89 and 10.00 µm. boiling at water b
atmospheric pressure -
7
Mechanically roughened surfaces
El-Genk and - Roughened surfaces with average surface - Saturated pool PF-5060 -
Suszko [86] and roughness ranging from 0.039 to 1.79 µm. boiling at m
Suszko and El- atmospheric pressure -
Genk [87] - Boiling surface 2
orientation 0º-180º
54
Pre-publication version
K.C. Leong et al. Applied Thermal Engineering 112 (2017) 999–1019
http://dx.doi.org/10.1016/j.applthermaleng.2016.10.138
Leong et al. [97] - Block and finned “Kfoam” graphite foams - Saturated pool FC-72 and -
and Pranoto et of 78% volume porosity and 0.5 mm pore boiling HFE-7000 t
al. [99] diameter.
- Block and finned “Pocofoam” graphite
foams of 61-75% porosity and 0.31-0.35 mm
pore diameter.
El-Genk and - Porous graphite of 60% volume porosity. - Saturated and 10- HFE-7100 -
Parker [100] - Consist of pore and re-entrant cavities of a 30 K subcooled pool w
few to hundreds of micrometers in size. boiling at -
atmospheric pressure 6
s
Ho et al. [106] - Surfaces with array of microcavities (500 - Pool boiling at FC-72 -
and 700 µm cavity mouth diameter). atmospheric pressure a
- Surfaces with array of microfins (350 and -
500 µm fin diameter). (
55
Pre-publication version
K.C. Leong et al. Applied Thermal Engineering 112 (2017) 999–1019
http://dx.doi.org/10.1016/j.applthermaleng.2016.10.138
Intrinsic surfaces
Rainey et al. - ABM microporous coating of about - Fluid velocities of 0.5, 2 FC-72 -
[140] 0.1–1 mm size cavities that is and 4 m/s -
approximately 50 - 4, 10 and 20 K liquid m
mm thick subcooling s
m
Ammerman and - Square minichannel of 2 mm side - Mass fluxes of 500, 2000 FC-87 -
You [141] length and heated length of 8 cm and 5000 kg/m2·s a
- Microporous coating of diamond - 2 to 21 K liquid -
powder with 8 – 12 µm particle size subcooling C
Sun et al. [142] - Horizontal, rectangular minichannels of - Mass flux of 700 kg/m2·s FC-72 -
0.49, 0.93 and 1.26 mm hydraulic - 20 K liquid subcooling e
diameter t
56
Pre-publication version
K.C. Leong et al. Applied Thermal Engineering 112 (2017) 999–1019
http://dx.doi.org/10.1016/j.applthermaleng.2016.10.138
Bai et al. [143] - Fifteen parallel microchannels of 400 - Mass flux of 182.8 Anhydrous -
mm in width, 900 mm in depth and 32 kg/m2·s ethanol a
mm in length - 58 K liquid subcooling W
- Microporous coating of dendrite copper
powders with diameters of about 30, 55
and 90 mm
Porous foams
Kim et al. [144] - Heated channel of 10 mm width, 7 mm - Mass fluxes of 20, 48 FC-72 and -
height and 37 mm length and 72 kg/m2·s water h
- Copper foams of 10 and 20 PPI, P
porosities of 92% and 95%
Pranoto and - Graphite foams of porosities of 61% - Mass fluxes of 50, 100 FC-72 -
Leong [146] and 72% and 150 kg/m2·s 2
- Foam dimension of 80 mm length, 60
mm width and 6mm height
Madani et al. - Upward flow channel of 50 mm length, - Mass fluxes of 6, 38, n-pentane -
[148] 10 mm width and 100 mm height 57 and 76 kg/m2·s t
- Copper foam of 36 PPI and porosity of q
97%
57