You are on page 1of 12

International Journal of Heat and Mass Transfer 147 (2020) 119015

Contents lists available at ScienceDirect

International Journal of Heat and Mass Transfer


journal homepage: www.elsevier.com/locate/hmt

Effects of surface inclination and type of surface roughness on the


nucleate boiling heat transfer performance of HFE-7200 dielectric fluid
Ngoctan Tran a,∗, Uzair Sajjad a, Rick Lin b, Chi-Chuan Wang a,∗
a
Department of Mechanical Engineering, National Chiao Tung University, 1001 University Road, Hsinchu 300, Taiwan
b
Delta electronics, Taoyuan 300, Taiwan

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

Article history: In the present work, the effects of surface roughness (including roughness from machined roughness and
Received 18 August 2019 sandblasted roughness), surface inclinations, and heat fluxes on the nucleate pool boiling heat transfer
Revised 3 November 2019
of HFE-7200 dielectric liquid at a working pressure of 1 atm. are examined in detail. Five aluminum
Accepted 6 November 2019
surfaces, with different roughness from 0.45 μm to 9 μm, are employed as test samples. The nucleate
Available online 15 November 2019
boiling phenomenon is examined with the wall-superheat temperature ranging from 5 °C to 30 °C, and
Keywords: surface inclinations of 0°, 90°, and 180°, respectively. SEM images and nucleate boiling bubbles captured
Nucleate pool boiling by a high-speed camera are utilized for further analyses. For all cases in this study, it is found that the
Dielectric fluid nucleate boiling heat transfer of the sandblasted rough surfaces is superior to those of the polished and
Surface roughness machined rough surfaces. For instance, at a wall-superheat temperature of 30 °C and inclination angle of
Inclinations 0°, the heat flux is augmented up to 69.7%, and 139.3% when compared to the polished surface for the
sandblasted surfaces with the roughness of 6.2 μm, and 9 μm, respectively. In addition, novel correlations
are also proposed for predicting the nucleate pool boiling heat transfer coefficient of HEF-7200 dielectric
fluid with different surface roughness, inclinations and heat fluxes.
© 2019 Elsevier Ltd. All rights reserved.

1. Introduction coatings [13,14]. The techniques employed for the surface modi-
fication entail various machining processes [15–17], sintering [18],
The utilization of high heat flux surfaces is quite common in a and electrodeposition [19,20], and chemical vapor deposition [21],
lot of industries, e.g., HVAC (heating, ventilation, and air condition- etc. Various refrigerants have been tested and reported includ-
ing), microelectronic chip cooling, nuclear power plants, boilers, ing ethanol [8], n-pentane [11], R-1234ze and R-134a [22], water
etc. [1–5]. Cooling is a major concern in these applications for en- [7,8,10–13, 15,16,19], and dielectric fluids [8,17,23].
suring safe operation. Appreciable measures to tailor the high-flux For concerns in direct contact of electronic heat devices with
applications had been made available, including jet impingement the working fluids, dielectric fluids like hydrofluoroethers (HFE-
cooling, single-phase cooling, spray cooling, cooling with phase 70 0 0, HFE-710 0, HFE-720 0, and HFE-730 0) and fluorocarbons (PF-
change materials, heat pipes, pool boiling, convective evaporation 5060, FC-87, and FC-72) are the best choices because of their high
and the like [6]. Among these cooling schemes, pool boiling is a dielectric strength, better chemical compatibility, and electrical in-
promising method because it can well tackle the issue of high-flux ertness. Besides, these working fluids can well maintain the junc-
heat dissipation with a very low wall superheat. tion temperature for electronic devices below the threshold value
The proficiency of the pool boiling process mainly depends on (typically in the range 80 °C to 130 °C). A short summary of the lit-
the interaction between liquid (wettability) and surface (rough- erature regarding the modified surfaces for dielectric fluids subject
ness). In line with this, significant efforts for surface modifica- to nucleate boiling performance is listed in Table 1.
tion to augment pool boiling were depicted in the literature. Based on the aforementioned reviewed literature, it is found
These surface modifications include micro-pillars [7], micro pin- that the two-phase immersion cooling using dielectric fluid is a
fins [8,9], microchannels [10], metal foams [11,12], and porous promising method for high-flux electronic devices. Among these
dielectric fluids, HFE-7200 not only contains zero ozone depletion
potential and low global warming potential but also has chemi-

Corresponding authors. cal and thermal stability, non-flammability, and low toxicity [24].
E-mail addresses: ngoctantran73@gmail.com (N. Tran), energyengineer01@ Moreover, the studies on nucleate pool boiling of HFE-7200 dielec-
gmail.com (U. Sajjad), RICK.WZ.LIN@deltaww.com (R. Lin), ccwang@nctu.edu.tw (C.-
C. Wang).
tric fluid are limited. Hence, further investigation of HFE-7200 di-

https://doi.org/10.1016/j.ijheatmasstransfer.2019.119015
0017-9310/© 2019 Elsevier Ltd. All rights reserved.
2 N. Tran, U. Sajjad and R. Lin et al. / International Journal of Heat and Mass Transfer 147 (2020) 119015

present experimental results, correlations for predicting the nucle-


Nomenclature ate pool boiling HTCs for different surface roughness and inclina-
tions are developed and proposed in Section 3.4. For consistency of
A heat transfer area, (m2 ) presentation, the five test samples, with the surface roughness of
Tsat wall superheat, (°C) 0.45 μm, 0.85 μm, 5 μm, 6.2 μm, and 9 μm, are named as test sam-
h heat transfer coefficient, (W/m2 K) ples 1, 2, 3, 4, and 5, respectively. Because the surface roughness of
k thermal conductivity, (W/mK) the test sample 1 is too small and polished, thereby it is defined
M the molecular mass, (kg/kmol) as a smooth surface and adopted as a baseline.
PR reduced pressure, (-)
Q heat transfer rate, (W) 2.2. Experimental setup, test sample, and testing procedure
q heat flux, (W/m2 )
Ra surface roughness, (m) In this study, the nucleate pool boiling heat transfer of the HFE-
Tsat fluid saturation temperature, (°C) 7200 dielectric fluid is examined on aluminum test samples with
Tw wall temperature, (°C) a square area of 40 × 40 mm2 as shown in Fig. 1. Fig. 1(a) presents
a 3D drawing of the test samples and Fig. 1(b) presents the di-
Subscripts
mensions of the test samples. The surface roughness of the test
sat saturation
samples is measured by using a portable surface roughness tester
w wall
(model of SJ-210, Mitutoyo). The surface roughness is obtained by
averaging ten times reading of roughness value. In addition, for
electric fluid is essential as far as high-performance thermal man- more information about the test samples, the surface roughness of
agement of electronic devices is concerned. the test samples is also captured by using a scanning electronic
Though dielectric fluids show great potential in immersion microscope (SEM). SEM images of the five test samples are pre-
cooling, there are some drawbacks associated with this dielectric sented in Fig. 1(c)–(g) for surface roughness of 0.45 μm, 0.85 μm,
fluid such as low surface tension which makes it highly wetting. 5 μm, 6.2 μm, and 9 μm, respectively. A heating block made of cop-
Due to its highly wetting nature, it results in high superheat incip- per is schematically shown in Fig. 2(a). Fig. 2(b) presents the di-
ience compared to the non-wetting, low-wetting, and other con- mensions of the heating block and the locations of the correspond-
ventional refrigerants like water. Besides, the thermophysical prop- ing thermocouples on the heating block. Aluminum samples are
erties of HFE-7200 are inferior to typical refrigerants or water. In attached to the copper heating block. Low melting temperature al-
this regard, many more studies of the surface modifications are es- loy (LMTA) is used to reduce the thermal contact resistance be-
sential for further augment pool boiling heat transfer of dielectric tween the sample and the heating block. Besides, the pressure
fluids. is created between the test sample and copper block in order to
For elaborating the effects of surface modification on the nu- further reduce the thermal contact resistance. The contact resis-
cleate boiling performance for dielectric fluid. The aim of this tance filled with the LMTA is about 0.03 °C cm2 W−1 as reported
work is to experimentally investigate the effects of type of surface in our recent study [25]; therefore, the contact resistance insignifi-
roughness (machined, polished, and sandblasted), heat flux, and cantly influences the present measurement. Four cartridge heaters
surface inclination on the saturated nucleate pool boiling of the (with a diameter of 10 mm and length of 50 mm) with a maxi-
well-wetting dielectric liquid, HFE-7200. The corresponding perfor- mum heat generation of 200 W for each heater are employed as
mance is explained with the associated visualization of bubble dy- a heat source. The cartridge heaters are inserted into four holes
namics. Tests are conducted at atmospheric pressure. This investi- at the bottom of the copper block as shown in Fig. 2(a). To re-
gation manifests its individuality in several aspects. Firstly, there duce the heat loss between the heating block and the working
are very limited studies on pool boiling of the well-wetting di- fluid as well as to support the heating block, a liquid-isolated cover
electric fluid, HFE-7200 (with very low global warming potential) made of Polyetheretherketone (PEEK) is designed and a tiny gap is
in the literature. Besides, different fabrication methods (polishing, also designed between the heating block and the PEEK cover to
machining, and sandblasting) are employed to fabricate rough sur- increase the thermal resistance between the heat source and the
faces for a range of roughness and their pool boiling performance, working fluid aiming at minimizing the heat loss as presented in
which is reported and discussed. Moreover, the pool boiling test Fig. 2(c) and (d). To measure the heat flux from the heat source to
results of the surfaces are utilized to derive new correlations for the examining surface, three sheathed thermocouples, T7, T8, and
prediction of nucleate boiling heat transfer coefficient for different T9 are designed at a centerline of the copper block as shown in
types of surface roughness (for a range of roughness), heat flux, Fig. 2(b). To measure the heat loss from the upper part of the cop-
and surface inclination. per block (after the three measured points), five sheathed thermo-
couples, entitled T2–T6, are installed at five designed locations on
2. Methodology a plane near the top of the heating block as depicted in Fig. 2(b).
To estimate the temperature of the effective surface, a sheathed
2.1. Work description thermocouple, T1, is located at the centerline of the test sample as
shown in Fig. 1(b). A schematic diagram of the experimental sys-
The aim of this investigation is to analyze the effects of sur- tem is presented in Fig. 3. A circular test chamber of total capacity
face roughness, surface inclination angle and heat flux on the ther- 45 liters (approximate 50 0 mm height and 30 0 mm diameter along
mal performance of the HFE-7200 dielectric fluid subject to nucle- with extended volumes of observation windows) made of stainless
ate pool boiling. Therefore, the wall-superheat temperature ranges steel 304 with three glass windows (two on the sidewall and third
from 5 °C to 30 °C is examined in detail. For investigating the afore- at the top of the chamber) for lighting and video recording is used
mentioned effects, five aluminum surfaces, with the surface rough- for experimentation. An amount of approximately 34 liters of the
ness of 0.45 μm, 0.85 μm, 5 μm, 6.2 μm, and 9 μm, are employed working fluid (HFE-7200 dielectric fluid) is charged into the test
as test samples and the test results are elaborated in Section 3.1. chamber. Non-condensable gases are eliminated by the degassing
The heat flux ranges from 9 kW/m2 to 90 kW/m2 and relevant re- device method. The saturated state of the working fluid is con-
sults are reported in Section 3.2. Three different surface inclina- trolled by a preheater and a condenser. Temperatures and absolute
tions, 0°, 90°, and 180°, are examined in Section 3.3. Based on the pressure of the experimental system are recorded by a data acqui-
N. Tran, U. Sajjad and R. Lin et al. / International Journal of Heat and Mass Transfer 147 (2020) 119015 3

Table 1
A summary of pool boiling studies of modified surfaces and dielectric fluids.

Fabrication Surface Test Dielectric Comparison with


Reference
methods characteristics conditions fluids plain surface

El-Genk and Suszko [36], Mechanically Roughness value: 0.039 to 1.79 μm Saturated pool boiling PF-5060 150% enhancement in HTC
Suszko and El-Genk roughened at 1 atm (1.72 W/cm2 K), and 37%
[37] enhancement in CHF (21.5 W/cm2 )
Yu et al. [38] MEMS/NEMS Cylindrical cavity arrays Saturated pool boiling FC-72 HTC value 11,000 W/m2 K CHF value
at 1 atm 30,000 W/m2 (2.5 times)
El-Genk and Parker [23] Open cell foams Pores and re-entrant cavities Saturated and HFE-7100 CHF is enhanced up to 60%.
subcooled pool
boiling at 1 atm
Jones et al. [39] Electrical discharge Roughness value: 1.08 to 10.00 μm Saturated pool boiling FC-77 HTC enhancement is up to 120%,
machining at 1 atm while CHF is enhanced up to 40%.
Sahu et al. [40] Spray coating Copper nanofiber of 15–20 μm Saturated pool boiling HFE-7300 CHF is enhanced up to 33%.
at 1 atm
El-Genk and Ali [14] Electrochemical Thickness ranging from 95–220 μm Saturated pool boiling PF-5060 HTC increased up to 17 times, while
deposition at 1 atm CHF up to 70%
Kim et al. [41] Particle sintering Particles ranging from 20 nm to Pool boiling at 1 atm FC-72 76% enhancement in HTC
9 mm placed on polished surface

sition recorder (MX100, Yokogawa) connected to a computer for The actual amount of heat flux through the examined surface is
further analysis. The inclination of the heater surface, as defined in estimated as follows:
Fig. 3, is controlled by a rotational controller. A high-speed cam-
era is placed in front of the side window to capture the boiling Qactual
qactual = (5)
process. A

2.3. Data reduction and uncertainty analysis where A is the cross-sectional area of the examined surface. As the
actual heat transfer area is influenced by the surface roughness;
The cross-section of the effective surface of the test sample is therefore, the cross-sectional area is used in this case. To obtain
designed the same as that of the cross-section of the upper part of the surface temperature, Tw , of the test sample, a sheathed ther-
the heating block, and four sides of the heating block are directly mocouple (T1) is installed at the center of the test sample as de-
contacted with unmoving air within the air gap. The thermal con- picted in Fig. 1(b). Tw is calculated by the following equation:
ductivity of air is extremely small compared to that of the cop-
per. In addition, the experiments are carried out for investigating
qactual L1
Tw = T1 − (6)
the nucleate boiling, thereby the generated heat from the heaters k
would be dissipated by the examining surface. It is assumed that
where the distance between the thermocouple T1 and the examin-
the heat transferred through any point on a cross-section on the
ing surface, temperature obtained by the thermocouple T1, thermal
upper part of the heated block is equal; hence one-dimensional
conductivity of the test sample, and the actual heat flux transferred
conduction law is employed and the heat flux through the cross-
though the examining surface, are denoted by L1 , T1 , k and qactual ,
section of the heating block is estimated as follows:
respectively. The heat transfer coefficient of the examining surface,
dT h, is calculated as follows:
qideal = −kcu (1)
dx
The heat transfer rate through the copper block is calculated as Qideal − Qloss Qactual qactual
h= = = (7)
follows: A(Tsat ) A(Tw − Tsat ) Tw − Tsat
dT
Qideal = −kcu A (2) where Tsat is the saturation temperature of the dielectric fluid, Ais
dx the cross-sectional area of the examining surface. For all the cases
where the temperature gradient, dTdx
, is calculated by Taylor’s series in this study, the investigation is carried out under the saturated
[26,27] for a three-point backward-difference approximation as fol- nucleate pool boiling regime of the HFE-7200 dielectric fluid at the
lows: atmospheric pressure. The saturation temperature of the working
dT 3T7 − 4T8 + T9 fluid is obtained with respect to the measured pressure within the
= (3) testing chamber and is rechecked by the refrigerant measured tem-
dx 2 x
perature. From Zhu et al. [27], the relative uncertainty in the heat
where T7, T8, and T9 are the temperatures measured by thermo- transfer coefficient is estimated as follows:
couples T7, T8, and T9, respectively, which are installed at three
locations as presented in Fig. 2(b), and xis the distance between  2  2  2  12
two neighboring thermocouples T7 and T8 or T8 and T9 as also h q Tw Tsat
= + + (8)
presented in Fig. 2(b). h q Tw − Tsat Tw − Tsat
To measure the heat loss, five thermocouples, including T2, T3,
T4, T5, and T6 are assembled on a plane, which is perpendicular
The accuracies of the temperature measurement and the tol-
to the centerline and near to the top surface of the heating block.
erance of the manufacturing process are ± 0.1 °C and ± 5%, re-
The locations of the thermocouples are presented in Fig. 2(a) and
spectively. The maximum relative uncertainty of the present study
(b). The actual heat transfer rate through the examined surface is
is estimated to be 19.2% at a heat flux of 3.86 kW/m2 of the test
calculated by the following equation:
sample 1 with a surface roughness of 0.45 μm and the inclination
Qactual = Qideal − Qloss (4) angle of 180°.
4 N. Tran, U. Sajjad and R. Lin et al. / International Journal of Heat and Mass Transfer 147 (2020) 119015

Fig. 1. Dimensions of the test sample and the heating block.


N. Tran, U. Sajjad and R. Lin et al. / International Journal of Heat and Mass Transfer 147 (2020) 119015 5

Fig. 2. Drawings of the heated block and PEEK cover.

3. Results and discussion the sandblasted rough surfaces. The nucleate pool boiling phe-
nomenon of the HFE-7200 dielectric fluid is examined at atmo-
3.1. Effect of surface roughness for various surface machining process spheric pressure and boiling temperature of 76 °C with the wall
on the nucleate pool boiling superheat ranging from 5 °C to 30 °C. The thermophysical proper-
ties of the HFE-7200 dielectric fluid at the atmospheric condition
In this section, nucleate boiling characteristics of the HFE-7200 are listed in Table 2. The experimental results, in terms of HTC,
dielectric fluid on effective surfaces are examined in detail. Five wall superheat, and boiling curve, are presented in Fig. 4. Fig. 4(a)
aluminum surfaces are employed as test samples, in which one and (b) presents test results of the HTC and wall superheat ver-
polished surface with a roughness of 0.45 μm is regarded as a sus the surface roughness for three different surface inclination an-
smooth surface and the baseline; two machined rough surfaces gles of 0°, 90°, and 180° at two fixed heat fluxes of 14 kW/m2 and
with the surface roughness of 0.85 μm, and 5 μm, are employed 40 kW/m2 , respectively. Experiments were repeated at least three
as machined rough surfaces; and two sandblasted rough surfaces times in order to ensure repeatability of the measurements, and it
with the surface roughness of 6.2 μm and 9 μm are examined as is found that consistent results prevail from all the experiments.
6 N. Tran, U. Sajjad and R. Lin et al. / International Journal of Heat and Mass Transfer 147 (2020) 119015

Fig. 3. A schematic diagram of the experimental setup.

Table 2 4, and 5 respectively, compared to the smooth surface. Fig. 4(e)–


Thermophysical properties of the HFE-7200 di-
(g) present the photos of nucleate boiling captured by the high-
electric fluid (at 1 atm.).
speed camera at a heat flux value of 9.16 kW/m2 for the test sam-
Properties HFE-7200 ples 1 (0.45 μm), 3 (5 μm) and 4 (6.2 μm), respectively. It is noted
Formulation C4 F9 OC2 H5 that the bubble density subject to high heat fluxes is very difficult
Molecular Wt. 264 to capture quantitatively; therefore, the photos of boiling phenom-
Boiling point (°C) 76 ena at a lower heat flux of 9.16 kW/m2 are used to elaborate the
Freeze point (°C) −138
test results in subsequent sections. The photos show that the bub-
Liquid density (g/ml) 1.43
Surface tension (dynes/cm) 13.6 ble density is increased with the surface roughness. With a higher
Vapor pressure (mmHg) 109 surface roughness, the heating surface provides many more cavities
Viscosity (cps) 0.61 leading to that more bubbles are generated and departed as seen
Specific heat (cal/g°C) 0.29
in the photos. In addition, the bubble density also rises apprecia-
bly against the heat flux under the nucleate pool boiling regime
[29]. As also presented by SEM images in Fig. 1(c)–(g), to create
Besides, no hysteresis was found. The results show that the HTCs a smooth surface, the original aluminum surface is polished by us-
increase with the increase in the surface roughness, and the in- ing sandpapers. The smoothness of the surface depends on the size
creasing rate of the sandblasted rough surfaces surpasses that of of the sandpaper and the polishing time. To create a machined
the machined rough surfaces. The required wall superheat tem- rough surface, the original surface is dug by blades installed on
perature is decreased with the increase in surface roughness. This a machine following a designed direction; therefore, parallel small
is because the roughness can promote the activation process of channels are created on the surface. The shape, depth, and width
the bubble nucleation sites and the bubble departing frequency, of the channels also depend on the used blades and the applied
which in turn promotes the nucleate boiling heat transfer [28–32]. force. To create a sandblasted rough surface, sands are injected
The results also show that the wall superheat of the sandblasted directly on the original aluminum surface; therefore, the cavities
rough surfaces decreases faster than that of the machined rough are created on the surfaces. The size and depth of the cavities de-
surfaces, while the wall superheat of the smooth surface is the pend on the size of the impingement sands and the impingement
highest. time. By using the aforementioned methods, the smooth, machined
Fig. 4(c) and (d) presents boiling curves of the five surfaces for rough, and sandblasted rough surfaces are created as presented
two different surface inclination angles of 0° and 90°, respectively. in Fig. 1(c), 1(d,e), and 1(g-f), respectively. The images show that
The results show that the heat flux (or in terms of wall superheat) the cavities in the polished surface (Fig. 1(c)) are very shallow.
controls the number of active nucleation sites, which are in line The cavities of the sandblasted surfaces are not only deeper but
with the prior report [28], meaning increased wall superheat pro- also much more than those of the machined rough surfaces. The
motes the activation of nucleation sites and decreasing the wall su- more and deeper the cavities on a surface area are, the larger the
perheat results in reducing activation of nucleation sites. It is found contacting area between the surface and the working fluid is ob-
that at the same wall superheat, the heat flux increases with the tained. The larger contacting porous area results in the more bub-
rise of the surface roughness, and this effect is more pronounced ble nucleations on the surface. In addition, more nucleating sites
for the surface with higher roughness. Similarly but with different can be activated, and the more the bubbles are generated [33,34].
rising trend, the sandblasted rough surfaces offers a higher heat These reasons explain the increased number of bubbles with the
transfer performance than that of the machined rough surfaces, increase in the surface roughness as presented in Fig. 4(e)–(g). Fig.
while the smooth surface exhibits the poorest heat transfer perfor- 4(e) and (f) and the foregoing reasons explain the rise in HTC,
mance among all the test samples. Specifically, at a wall superheat lower wall superheat, and the boiling curves depending on the sur-
of 30 °C and inclination of 0°, the corresponding heat flux is aug- face roughness and roughness machining methods as presented in
mented up to 6.2%, 46%, 69.7%, and 139.3% for test samples 2, 3, Fig. 4(a)–(d).
N. Tran, U. Sajjad and R. Lin et al. / International Journal of Heat and Mass Transfer 147 (2020) 119015 7

Fig. 4. Heat transfer coefficient and temperature differences of different roughnesses.

3.2. Effect of inclination angle on the nucleate boiling HTC of the surface with an inclination of 90° is slightly higher
than that with the inclination of 0°. At the same inclination an-
In this section, the nucleate boiling phenomena of the HFE- gles of 0°, or 90°, the HTC of the sandblasted rough surface is
7200 dielectric fluid on the three test samples, with the surface much higher than those of the smooth and machined rough sur-
roughness of 0.45 μm, 5 μm, and 9 μm representing for smooth, face. For instance, at 40 kW/m2 heat flux and 0° inclination, the
machined rough and sandblasted surfaces, respectively, are exam- HTCs are improved up to 20% and 73.7% compared to the base-
ined in detail at three inclination angles of 0°, 90°, and 180°. The line for the machined and sandblasted rough surfaces, respectively.
experimental conditions presented in Section 3.1 are employed. Also for the same heat flux (40 kW/m2 ), the HTC of the sandblasted
Fig. 5(a) presents the HTCs versus the heat flux of the test samples surface with the inclination of 90° is 6.7% higher than that of 0°.
for three different inclination angles of 0°, 90°, and 180°, respec- With the inclination of 180° and at a heat flux to be smaller than
tively. The results show that the HTC is increased with the heat 25 kW/m2 , the HTCs of the machined and sandblasted rough sur-
flux and surface roughness. For the same surface roughness, the faces are slightly higher than that with the inclination of 0°; how-
8 N. Tran, U. Sajjad and R. Lin et al. / International Journal of Heat and Mass Transfer 147 (2020) 119015

Fig. 5. The experimental results of nucleate pool boiling for different inclination angles.

ever, that of the smooth surface is lowest when compared to other the roughness of 0.45 μm, 5 μm, and 9 μm subject to three inclina-
inclinations. Fig. 5(b) presents the HTCs versus the wall superheat tions of 0°, 90°, and 180°. The results show that for the inclina-
of the test samples for three different inclination angles of 0°, 90°, tions of 0° and 90°, the heat flux increases with the increase in the
and 180°. The results show that the HTC increases with the rise wall superheat temperature and the increase in surface roughness.
in the wall superheat for all cases. At the same wall superheat, At the inclination of 180°, the heat flux not only increases very
the HTC with an inclination of 90° is higher than that of 0°. The slowly with the rise of the wall superheat but also is limited un-
HTC of the sandblasted rough surface is much higher than that of der 25 kW/m2 . It is interesting to found that at the low wall super-
the smooth surface. For instance, at a wall superheat of 17.65 °C heats, the nucleate boiling HTC of the 90° inclined surface is higher
with the inclination of 0°, the HTCs are enhanced up to 34.7% and than that of the horizontal one (0°) as aforementioned; however,
218% compared to that of the baseline for the machined rough when the wall superheat is increased near to 30 °C, the nucleate
surface (5 μm) and the sandblasted rough surface (9 μm), respec- boiling HTCs of the two inclinations converge together. Fig. 5(d)–(f)
tively. Fig. 5(c) presents the boiling curve of the three surfaces with presents the photos of bubbles captured by the high-speed camera
N. Tran, U. Sajjad and R. Lin et al. / International Journal of Heat and Mass Transfer 147 (2020) 119015 9

for the machined rough surface (5 μm) with a fixed heat flux of curves of the five examined surfaces subject to the three inclina-
9.16 kW/m2 at three inclinations of 0°, 90°, and 180°, respectively. tions. The results show that the heat fluxes are increased with the
The photos show that with a horizontal facing-up surface (inclina- increase in the wall superheat and with the increase in the surface
tion of 0°), a bubble departs the nucleation cavity when the buoy- roughness [34,35]. The heat fluxes are not linearly increased with
ancy force of that bubble overcomes surface tension and the grav- the increase in the surface roughness; yet it depends on the fabri-
ity force applied to it. Yet, with the 90° inclined surface, a bubble cation method. For instance, at a wall superheat of 30 °C, the sur-
at the lower sections does not leave the effective surface when it face roughness increases from 0.85 μm to 5 μm is 4.15 μm (for two
departs the cavity; it slips alongside the effective surface in the up- machined surfaces), from 5 μm (machined) to 6.2 μm (sandblasted)
ward direction. Eventually, the sliding bubbles collide and push the is 1.2 μm and from 6.2 μm to 9 μm (for two sandblasted surfaces)
upper bubbles out of their cavities earlier compared to those with- is 2.8 μm; however, the heat-flux increases for those groups are
out the sliding influence. In essence, the bubbles appeared on the 24.56%, 20.1%, and 42.6%, respectively.
upper sections of the 90° inclined surface are not only the bubbles Fig. 7(a)–(f) presents nucleate pool boiling photos of a test sam-
of the upper sections but also include bubbles from the lower sec- ple with the horizontal surface (0°), and the surface roughness of
tions; therefore, the density of the bubbles of the upper sections is 5 μm for heat fluxes of 9, 15, 25, 34, 39, and 47 kW/m2 , respec-
higher than those of the lower section with the 90° inclined sur- tively. The images show that an increment in the heat flux in-
faces as presented in Fig. 5(e). For the 180° inclination, because the creases the density of bubbles [29]. Moreover, it is found that the
heated surface is facing down; therefore, the bubbles cannot leave bubbles not only go upward but also merge toward the centerline
the surface and go up as other inclinations while the bubbles ac- of the bubble region. This is expected since the buoyancy force cre-
cumulated on the surface to engender a vapor film as presented ates a higher rising bubble velocity that leads to a comparatively
in Fig. 5(f). With the inclination of 0°, the bubbles can immedi- low-pressure zone, thereby bubbles move toward the center zone.
ately leave the surface (a nucleate boiling regime) while the sur- In the nucleate boiling region, the rise of heat flux leads to the
rounding liquid can refill the cavities efficiently to regenerate bub- activation of more nucleate sites and higher departing frequency
bles continuously; therefore, most of the heat, which is transferred as well. Therefore, the density of the bubbles is increased with
from the heat source to the working fluid, is though the latent the increase in heat flux as presented in Fig. 7. The higher sur-
heat. In addition, the higher the wall superheat is, the faster the face roughness could contain more nucleation sites; therefore, the
bubbles generate and detach the surface leading to the more heat HTC increases and the wall superheat temperature is decreased as
to be removed from the surface. The foregoing reasons explain for discussed in Figs. 4 and 5. These explain for the distributions of
the increase of the HTCs with the increase in the wall superheat the HTC and wall superheat temperature for the surfaces as pre-
temperature or the increase in heat flux with the inclination of 0° sented in Fig. 6. The density of the bubbles is much smaller than
as presented in Fig. 5(a)–(c). With the inclination of 90°, at small that of the liquid; therefore, the greater the number of bubbles in
wall superheat, the slipping causes the detachment of bubbles at a vapor-and-liquid mixture is, the smaller the density of that mix-
a higher rate compared to no slipping as aforementioned [33,34]; ture is achieved. This leads to the smaller pressure at the vapor-
however, when the wall superheat reaches near to 30 °C, the up- and-liquid mixed region as compared to that of the liquid region.
ward velocity of the lower-section bubbles is slower than that of This explains the trend of the bubbles that become concentrated
departure rate of the upper-section bubbles; therefore, it could not to the centerline as presented in Fig. 7.
further increase the departure rate of the upper-section bubbles
anymore. This explains for the HTC of the vertical surface (90°) to 3.4. Experimental data verification and correlation development
be slightly higher than that of the horizontal one at a low wall
superheat and converges eventually at high wall superheat as pre- The test results are compared with a well-known correlation
sented in Figs. 4(a), (b) and 5(a)–(c). With the inclination of 180°, for predicting the nucleate pool boiling of the flat plate by Cooper
at a low heat flux, the number of generated bubbles is compara- [35] whose correlation is given in Eq. (9).
tively fewer; therefore, it could not cover the overall surface. As a
h = 90PR (0.12−0.4343lnR p ) (−0.4343 ln PR )
−0.55
consequence, the HTC increases with the increased heat flux un- M−0.5 q0.67 (9)
der 25 kW/m2 depending on the surface roughness as presented where PR , Rp , M, and q represent the reduced pressure, surface
in Fig. 5(a)–(c). When the heat flux is higher than 25 kW/m2 , the roughness, molecular mass, and heat flux respectively. Fig. 8(a)
generated bubbles cover the overall surface. At that moment, the presents a comparison for experimental HTC and HTC based on
wall superheat increases significantly. The vapor blanket deterio- Cooper’s correlation. The comparison shows that a satisfactory
rates the HTC appreciably and the wall superheat of the 180° incli- agreement was found and 71.6% of the current data (from the
nation is raised significantly. It is noted that concerning the safety experiments) against Cooper’s prediction are within ± 40%. The
of the experimental system, the present experiments with the in- difference between the present experimental results and Cooper’s
clination of 180° are only carried out for the heat fluxes under prediction could be due to the differences in the working fluid and
25 kW/m2 . experimental conditions.
On the basis of experimental data, the correlations, which can
3.3. Effect of heat flux on the nucleate boiling well-predict the nucleate pool boiling HTC for the polished, ma-
chined, and sandblasted rough surfaces of the examined range for
In this section, the nucleate pool boiling of the HFE-7200 di- the HFE-7200 dielectric fluid at atmospheric working condition, are
electric fluid on the five test samples with different heat fluxes in proposed, as follows:
the range from 0 to 90 kW/m2 and three surface inclined angles of For smooth and machined rough surfaces,
0°, 90°, and 180° is examined in detail. The experiments are car-
ried out at the same aforementioned conditions. Fig. 6(a), and (b) hmachined−sur f ace
presents the heat transfer coefficient versus the heat flux subject to (0.02242×Ra +0.958 )
= (0.021 + 47.07 ) × Ra 0.01317 × Tsat (10)
the surface inclination of the machined and sandblasted rough sur-
faces, respectively. The results indicate that the HTC increases with For sandblasted rough surfaces,
the rise in the heat flux and this is the reason why the HTCs of
hsandblasted−sur f aces
the sandblasted rough surfaces are much higher than those of the
(0.021607×Ra +0.394 )
smooth or machined rough surfaces. Fig. 6(c) presents the boiling = (0.078 + 155.5 ) × Ra 0.5123 × Tsat (11)
10 N. Tran, U. Sajjad and R. Lin et al. / International Journal of Heat and Mass Transfer 147 (2020) 119015

Fig. 6. Heat transfer coefficients and boiling curves of different heat fluxes.

Fig. 7. Photos of bubble patterns with different heat fluxes.


N. Tran, U. Sajjad and R. Lin et al. / International Journal of Heat and Mass Transfer 147 (2020) 119015 11

Fig. 8. A comparison between exp. and pred. results for heat transfer coefficients.

where Ra , Tsat , and  are the surface roughness, wall superheat, 4. Conclusions
and inclination, respectively. A comparison between the experi-
mental and predicted (on the basis of proposed correlation) HTCs In the present study, the influences of surface roughness, sur-
is presented in Fig. 8(b). The comparison shows a decent agree- face inclination, and heat flux on the nucleate pool boiling heat
ment between the experimental and the predicted data. The pre- transfer of HFE-7200 dielectric liquid at the ambient environment
dictive ability of the proposed correlation is that 96% of the test condition (1 atm.) are examined in detail. Five aluminum surfaces
data falls within ±10% predictive span. Fig. 8(c) and (d) presents of three groups of different surface roughness, including (1) pol-
boiling curves, which are predicted by the proposed correlations, ished surface (0.45 μm), (2) machined rough surfaces (0.85 μm and
for the machined and sandblasted rough surfaces with ten dif- 5 μm), and (3) sandblasted rough surfaces (6.2 μm and 9 μm), are
ferent surface roughness at the inclinations of 0° and 90°, re- employed as test samples. The nucleate boiling characteristics are
spectively. The predicted results show that the correlations not examined with the wall superheat ranging from 5 °C to 30 °C, and
only can regenerate the boiling curves of the examined sur- surface inclinations of 0°, 90°, and 180° SEM images and nucleate
faces with the accuracy up to 98% but also can predict the boil- boiling bubbles captured by a high-speed camera are utilized for
ing curves for the new surfaces with different roughness very further analyses. Based on the foregoing discussions, some conclu-
well. sions are encapsulated as follows:
12 N. Tran, U. Sajjad and R. Lin et al. / International Journal of Heat and Mass Transfer 147 (2020) 119015

1. The nucleate boiling heat transfer coefficients are improved [12] Y. Yang, X. Ji, J. Xu, Pool boiling heat transfer on copper foam covers with water
with the increase in the surface roughness; however, the as working fluid, Int. J. Therm. Sci. 49 (2010) 1227–1237.
[13] A. Gheitaghy, H. Saffari, D. Ghasimi, A. Ghasemi, Effect of electrolyte temper-
improvement is not linear with the increase in the surface ature on porous electrodeposited copper for pool boiling enhancement, Appl.
roughness, but it also depends on the method to create the Therm. Eng. 113 (2017) 1097–1106.
roughness. For instance, at 30 °C wall superheat and 0° sur- [14] M.S. El-Genk, A.F. Ali, Enhanced nucleate boiling on copper micro-porous sur-
faces, Int. J. Multiph. Flow 36 (2010) 780–792.
face inclination, the heat flux is improved up to 6.2%, 46%, [15] Y. Tang, J. Zeng, S. Zhang, C. Chen, J. Chen, Effect of structural parameters on
69.7%, and 139.3% when compared to the baseline for the pool boiling heat transfer for porous interconnected microchannel nets, Int. J.
surfaces with the roughness of 0.85 μm, 5 μm, 6.2 μm, and Heat Mass Transf. 93 (2016) 906–917.
[16] C.M. Kruse, T. Anderson, C. Wilson, C. Zuhlke, D. Alexander, G. Gogos, et al., En-
9 μm, respectively. This is because sandblasted rough sur-
hanced pool-boiling heat transfer and critical heat flux on femtosecond laser
faces contain more nucleation sites compared to machined processed stainless steel surfaces, Int. J. Heat Mass Transf. 82 (2015) 109–
and polished surfaces. 116.
[17] J.Y. Ho, K.K. Wong, K.C. Leong, Saturated pool boiling of FC-72 from enhanced
2. At low wall superheats and low heat fluxes, the nucleate boil-
surfaces produced by selective laser melting, Int. J. Heat Mass Transf. 99 (2016)
ing heat transfer performance of a surface with the vertical 107–121.
arrangement (90°) is higher than that of the horizontal in- [18] X. Ji, J. Xu, Z. Zhao, W. Yang, Pool boiling heat transfer on uniform and non-u-
clination (0°). Specifically, at the heat flux of 40 kW/m2 , the niform porous coating surfaces, Exp. Therm. Fluid Sci. 48 (2013) 198–212.
[19] Y.Q. Wang, J.L. Luo, Y. Heng, D.C. Mo, S.S. Lyu, Wettability modification to fur-
heat transfer coefficient of the sandblasted surface (9 μm) ther enhance the pool boiling performance of the micro nano bi-porous copper
with the vertical arrangement (90°) is 6.7% higher than that surface structure, Int. J. Heat Mass Transf. 119 (2018) 333–342.
of the horizontal one (0°). However, no significant differ- [20] P. Xu, Q. Li, Y. Xuan, Enhanced boiling heat transfer on composite porous sur-
face, Int. J. Heat Mass Transf. 80 (2015) 107–114.
ence in HTC is seen amid vertical or horizontal arrangement [21] M.R.S. Naseem Abbas, M. Hussain, S.M.Z. Mehdi, U. Sajjad, Fabrication and
when the wall superheat is increased near to 30 °C. characterization of silver thin films using physical vapor deposition, and the
3. The downward-facing heating surface (inclination of 180°) is investigation of annealing effects on their structures, Mater. Res. Express 6
(2019) 1–11.
not recommended for using in heat flux over 25 kW/m2 be- [22] N. Tran, S.R. Sheng, C.C. Wang, An experimental study and empirical correla-
cause the bubbles will be accumulated on the surface and tions to describe the effect of lubricant oil on the nucleate boiling heat trans-
a vapor film will be formed, deteriorating the heat transfer fer performance for R-1234ze and R-134a, Int. Commun. Heat Mass Transf. 97
(2018) 78–84.
significantly.
[23] M.S. El-Genk, J.L. Parker, Enhanced boiling of HFE-7100 dielectric liquid on
4. The surface roughness not only promotes the heat transfer porous graphite, Energy Convers. Manag. 46 (2005) 2455–2481.
coefficient but also can decrease the wall superheat signif- [24] M.S. El-Genk and J.L. Parker, 3MTM NovecTM 7200 engineered fluid. Avail-
able: https://multimedia.3m.com/mws/media/199819O/3mtm- novectm- 7200-
icantly. The wall superheat is less for the sandblasted and
engineered-fluid.pdf
machined surfaces compared to the polished surface. [25] W.X. Chu, P.H. Tseng, C.C. Wang, Utilization of low-melting temperature alloy
5. Correlations, which are proposed in this study, could predict with confined seal for reducing thermal contact resistance, Appl. Therm. Eng.
the nucleate pool boiling heat transfer coefficients of HFE- 163 (2019) 114438 12/25/ 2019.
[26] D. Cooke, S.G. Kandlikar, Effect of open microchannel geometry on pool boiling
7200 dielectric fluid for polished, machined and sandblasted enhancement, Int. J. Heat Mass Transf. 55 (2012) 1004–1013.
rough surfaces at the inclinations of 0° (horizontal facing up) [27] Y. Zhu, H. Hu, G. Ding, D. Zhuang, H. Peng, Heat transfer enhancement by
and 90° (vertical) with the predictive ability to describe 96% metal foam during nucleate pool boiling of refrigerant/oil mixture at a wide
range of oil concentration, HVAC&R Res. 18 (2012) 377–389.
of the test data to be within ±10%. [28] S. Gong, P. Cheng, Lattice Boltzmann simulations for surface wettability ef-
fects in saturated pool boiling heat transfer, Int. J. Heat Mass Transf. 85 (2015)
Acknowledgment 635–646.
[29] S. Gong, P. Cheng, X. Quan, Two-dimensional mesoscale simulations of satu-
rated pool boiling from rough surfaces. part I: Bubble nucleation in a single
This work was supported by the Ministry of Science Technology cavity at low superheats, Int. J. Heat Mass Transf. 100 (2016) 927–937.
of Taiwan, under grant numbers: MOST 106-2811-E-009-013 and [30] S. Gong, P. Cheng, Two-dimensional mesoscale simulations of saturated pool
boiling from rough surfaces. Part II: bubble interactions above multi-cavities,
107-2622-E-0 09-0 02-CC2. The supports are deeply appreciated.
Int. J. Heat Mass Transf. 100 (2016) 938–948.
[31] X. Ma, P. Cheng, 3D simulations of pool boiling above smooth horizontal
References heated surfaces by a phase-change lattice Boltzmann method, Int. J. Heat Mass
Transf. 131 (2019) 1095–1108.
[1] V.P. Carey, Liquid Vapor Phase Change Phenomena: An Introduction to the [32] X. Ma, P. Cheng, Dry spot dynamics and wet area fractions in pool boiling
Thermophysics of Vaporization and Condensation Processes in Heat Transfer on micro-pillar and micro-cavity hydrophilic heaters: a 3D lattice boltzmann
Equipment, CRC Press, 2018. phase-change study, Int. J. Heat Mass Transf. 141 (2019) 407–418.
[2] V. Dhir, Boiling heat transfer, Annu. Rev. Fluid Mech. 30 (1998) 365–401. [33] D. Jung, J. Venart, A. Sousa, Effects of enhanced surfaces and surface orienta-
[3] D.E. Kim, D.I. Yu, D.W. Jerng, M.H. Kim, H.S. Ahn, Review of boiling heat trans- tion on nucleate and film boiling heat transfer in R-11, Int. J. Heat Mass Transf.
fer enhancement on micro/nanostructured surfaces, Exp. Therm. Fluid Sci. 66 30 (1987) 2627–2639.
(2015) 173–196. [34] J. Chang, S. You, Heater orientation effects on pool boiling of micro-porous-en-
[4] E. Pop, Energy dissipation and transport in nanoscale devices, Nano Res. 3 hanced surfaces in saturated FC-72, J. Heat Transf. 118 (1996) 937–943.
(2010) 147–169. [35] M.G. Cooper, Heat flow rates in saturated nucleate pool boiling-a wide-rang-
[5] J.R. Thome, State-of-the-art overview of boiling and two-phase flows in mi- ing examination using reduced properties, Adv. Heat Transf. 16 (1984) 158–
crochannels, Heat Transf. Eng. 27 (2006) 4–19. 237.
[6] S. Sinha-Ray, W. Zhang, R.P. Sahu, S. Sinha-Ray, A.L. Yarin, Pool boiling of [36] M.S. El-Genk, A. Suszko, Saturation nucleate boiling and correlations for
novec 7300 and di water on nano-textured heater covered with supersoni- PF-5060 dielectric liquid on inclined rough copper surfaces, J. Heat Transf. 136
cally-blown or electrospun polymer nanofibers, Int. J. Heat Mass Transf. 106 (2014) 081503.
(2017) 482–490. [37] A. Suszko, M.S. El-Genk, Saturation boiling of PF-5060 on rough cu surfaces:
[7] K.H. Chu, R. Enright, E.N. Wang, Structured surfaces for enhanced pool boiling Bubbles transient growth, departure diameter and detachment frequency, Int.
heat transfer, Appl. Phys. Lett. 100 (2012) 241603. J. Heat Mass Transf. 91 (2015) 363–373.
[8] R. Pastuszko, Pool boiling heat transfer on micro-fins with wire mesh–Experi- [38] C.K. Yu, D.C. Lu, T.C. Cheng, Pool boiling heat transfer on artificial micro-cavity
ments and heat flux prediction, Int. J. Therm. Sci. 125 (2018) 197–209. surfaces in dielectric fluid FC-72, J. Micromech. Microeng. 16 (2006) 2092.
[9] Z. Cao, B. Liu, C. Preger, Z. Wu, Y. Zhang, X. Wang, et al., Pool boiling heat [39] B.J. Jones, J.P. McHale, S.V. Garimella, The influence of surface roughness on
transfer of FC-72 on pin-fin silicon surfaces with nanoparticle deposition, Int. nucleate pool boiling heat transfer, J. Heat Transf. 131 (2009) 121009.
J. Heat Mass Transf. 126 (2018) 1019–1033. [40] R.P. Sahu, S. Sinha-Ray, S. Sinha-Ray, A.L. Yarin, Pool boiling of novec 7300
[10] A. Walunj, A. Sathyabhama, Comparative study of pool boiling heat transfer and self-rewetting fluids on electrically-assisted supersonically solution-blown,
from various microchannel geometries, Appl. Therm. Eng. 128 (2018) 672–683. copper-plated nanofibers, Int. J. Heat Mass Transf. 95 (2016) 83–93.
[11] L. Zhou, W. Li, T. Ma, X. Du, Experimental study on boiling heat transfer of [41] T.Y. Kim, J.A. Weibel, S.V. Garimella, A free-particles-based technique for boil-
a self-rewetting fluid on copper foams with pore-density gradient structures, ing heat transfer enhancement in a wetting liquid, Int. J. Heat Mass Transf. 71
Int. J. Heat Mass Transf. 124 (2018) 210–219. (2014) 808–817.

You might also like