Anda di halaman 1dari 9

Experimental Thermal and Fluid Science 40 (2012) 150158

Contents lists available at SciVerse ScienceDirect

Experimental Thermal and Fluid Science


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

Morphological change of plain and nano-porous surfaces during boiling


and its effect on nucleate pool boiling heat transfer
Chi Young Lee a,b, Bong June Zhang b, Kwang J. Kim b,
a
b

KAERI (Korea Atomic Energy Research Institute), 989-111 Daedeok-daero, Yuseong-gu, Daejeon 305-353, Republic of Korea
Low Carbon Green Technology Laboratory, Department of Mechanical Engineering (MS312), University of Nevada Reno, Reno, NV 89557, USA

a r t i c l e

i n f o

Article history:
Received 10 October 2010
Received in revised form 30 December 2011
Accepted 24 February 2012
Available online 3 March 2012
Keywords:
Nucleate pool boiling heat transfer
Surface morphology
Aluminum hydroxide

a b s t r a c t
In order to investigate nucleate pool boiling heat transfer characteristics in saturated water, a plain aluminum alloy surface (6061, untreated) and a uniquely fabricated aluminum oxide Nano-Porous Surface
(NPS) sample were prepared. Generally, the NPS exhibited a lower wall superheat at the onset of nucleate
boiling and a higher nucleate boiling heat transfer coefcient than the plain surface. It was also noted that
the nucleate boiling heat transfer coefcient decreased by 30% on the plain surface and by 37% on the NPS
after ve repeated boiling tests. It was found that such performance-degradation in both of the test samples was due primarily to the formation of aluminum hydroxide on the boiling surfaces. It is our belief
that the aluminum hydroxide, formed on both boiling surfaces, signicantly altered the surface morphologies as shown in the microscopic images of the post-boiling surfaces, which resulted in the decrease of
the active nucleation sites. Although NPS is an excellent tool to enhance the nucleate pool boiling heat
transfer coefcient, its life span and usefulness can be impeded by undesirable surface reactions with
working uids.
2012 Elsevier Inc. All rights reserved.

1. Introduction
The development of special surface geometries/structures is a
critical issue for the performance-enhancement of the pool boiling heat transfer. Numerous studies and experiments on nucleate
pool boiling, an effective way to remove high heat ux from a
heated surface, have focused on enhancing the boiling heat transfer coefcient by fabricating micro-scaled surface geometries/
structures [18]. One approach that has success is to create a
number of small micro-porous cavities on the boiling surfaces
to increase the vapor/gas entrapment and the number of active
nucleation sites. These cavities reduce the incipient and nucleate
boiling wall superheats and increase the pool boiling heat transfer
coefcient [9].
While many of the previous studies [18] on the surface structure focused on improving the heat transfer coefcient of pool
boiling in micro-sized dimensions; however, with the evolution
of nanotechnology and nano-materials, our study focused on
well-dened nano-materials suitable for use in pool boiling.
Currently, research of nano-materials used in pool boiling heat
transfer is limited. Vemuri and Kim [9] carried out a brief experimental study on the pool boiling heat transfer of a Nano-Porous

Corresponding author. Tel.: +1 775 784 6931; fax: +1 775 784 1701.
E-mail address: kwangkim@unr.edu (K.J. Kim).
0894-1777/$ - see front matter 2012 Elsevier Inc. All rights reserved.
doi:10.1016/j.expthermusci.2012.02.011

Surface (NPS) using commercially-available aluminum oxide with


a thickness of 70 lm in a saturated FC-72 dielectric uid. Their
study revealed that the incipient wall superheat was reduced by
30% in the input power for the NPS versus the plain surface. This
nding hints that the enhancement could be due to the nanoporous structure increasing both the vapor entrapment volume
and the active nucleation site density, which in turn enhances
the boiling heat transfer performance. It should be noted that
anodizing is a promising candidate for manufacturing a NPS boiling surface since the anodized self-assembled metal oxides are
well-ordered structures on a metal substrate. To make a nano-porous structure on the aluminum alloy, the typical anodizing process
is comprised of three steps as illustrated in Fig. 1: (i) Degreasing,
(ii) electro-polishing, and (iii) anodization.
Recently, Launay et al. [10] investigated the effect of various
surface structures on the boiling heat transfer working with the
uids PF-5060 and De-Ionized (DI) water. In their study, seven different kinds of surface structures were fabricated: A smooth Si
sample; a roughened (dry etched) Si sample; a sample fully coated
with Carbon NanoTubes (CNTs); a silicon pinn microstructure; a
CNT pin n structure grown on silicon; and two 3D silicon structures, one of which was coated with CNTs over its top surface.
The maximum heat uxes achieved in the 3D microstructure without CNTs were reported to be 270 and 1300 kW/m2 for PF-5060
and water, respectively. The CNT-enabled nano-structured surface,
with the inner diameter of the Multi-Walled Carbon NanoTubes

C.Y. Lee et al. / Experimental Thermal and Fluid Science 40 (2012) 150158

151

Nomenclature
A
C1
C2
h
I
i
P
Q
r
T
DT
V
y

heat transfer surface area (m2)


constant value in Eq. (6) ()
constant value in Eq. (6) ()
heat transfer coefcient (W/m2K)
current (A)
enthalpy (J/kg)
pressure (Pa)
power input (W)
radius (m)
temperature (K)
temperature difference (K)
voltage (V)
distance from heating surface (m)

Greek Letters
d
thermal boundary layer thickness (m)
h
contact angle ()
q
density (kg/m3)
r
surface tension (N/m)

(MWCNTs) ranged between 10 and 20 nm; the wall thickness was


10 nm; and the pitch between individual nanotubes was 50
70 nm, appeared to improve the pool boiling heat transfer at low
superheat conditions when compared to a smooth surface. Ahn
et al. [11] also conducted the experiments in the nucleate and lm
boiling regimes to investigate the effect of the MWCNT height on
the pool boiling heat transfer performance with PF-5060 as the
working liquid. They used two silicon wafer substrates coated with
vertically aligned MWCNT forests as the test surfaces. The
MWCNT forests with the different heights of 9 and 25 lm were
synthesized on the silicon wafer substrates using a Chemical Vapor
Deposition (CVD) process. They reported that the MWCNT forests
signicantly enhanced the nucleate boiling heat transfer and Critical Heat Flux (CHF) when compared to smooth bare silicon wafer
(without MWCNT coating). Also, in the lm-boiling regime,
MWCNT with 25 lm height achieved a 57% higher heat ux at
Leidenfrost Point when compared to smooth bare silicon wafer.
The longevity of the enhanced heat transfer is an important
parameter for various heat transfer applications. Chaudhri and
McDougall [12] mentioned the degradation of heat transfer surfaces and pointed out that the varying activity in minute pits,
grooves and other small unwetted spots on the solid surface, deposition of a fouling lm, and corrosion are important factors to be
considered. A recent study by Lee et al. [13] reported the signicant
performance-degradation of aluminum oxide NPS in pool boiling
with saturated water: while the pool boiling occurs vigorously on
the heated surface, the NPS experiences the chemical reaction between the heated surface and the uid, which may considerably affect the change of nucleate boiling heat transfer behavior. Hence,

Subscripts
ave
average value
b
bubble
bottom at the bottom of test section
c
cavity
f
liquid
fg
liquidgas
left
at the left of test section
max
maximum
min
minimum
right
at the right of test section
sat
saturation
sub
subcooling
t
thermal boundary layer
top
at the top of test section
v
vapor
w
wall

further studies on the heat transfer coefcient and long-term performance of the surface in the nucleate pool boiling are needed.
The objective of the present study is to investigate the morphological change of NPS and plain (non-structured) aluminum alloy
(6061) surfaces in a saturated water pool boiling condition, and
the surfaces inuence on the nucleate boiling heat transfer. The
NPS was fabricated utilizing an anodizing technique. Through a
series of repeated tests, the nucleate boiling heat transfers on the
NPS and plain surfaces were monitored. The relationship between
the change of the nucleate boiling heat transfer and the transformation of the surface morphology by the chemical reaction between such surfaces and water were also observed.

2. Experiment
2.1. Sample preparation
Fig. 2 shows the schematic diagram and a photo of the boiling
test section. An aluminum alloy rod (6061, 19.1 mm in diameter
and 101.6 mm in length), with a center hole (9.5 mm wide and
99.1 mm long), was used. Four holes, each 2.4 mm in diameter
and 70.0 mm in depth, were drilled at the top, middle, and the bottom 90 apart along the periphery of the test section. In order to
heat the test section, a cartridge heater (diameter of 9.5 mm and
a length of 76.2 mm) was inserted into the center hole with a high
thermal conductive paste (Omega 201). T-type thermocouples
(Omega) were inserted into the small exterior holes with the high
thermal conductive paste to measure the surface temperatures.

Fig. 1. Typical anodizing process to make nano-porous structure on aluminum alloy (6061).

152

C.Y. Lee et al. / Experimental Thermal and Fluid Science 40 (2012) 150158

Fig. 2. Schematic diagram of test section (unit: mm).

Fig. 3. Images of fabricated NPS: (a) FESEM and (b) optical microscope.

The two kinds of boiling surfaces on the test sections were prepared by electro-polishing (plain surface) and multi-step anodizing
(NPS) techniques, respectively. The fabrication process for the
nano-porous coating on aluminum alloy test section was performed in two steps; electro-polishing and subsequent anodizing.
The aluminum alloy specimens were annealed at 400 C for four
hours and then cooled to room temperature. The heat-treated
specimens were electro-polished in perchloric acid and ethanol
(1:4 by vol) at 12 V for ve minutes at 0 C. The nished specimens
were thoroughly rinsed with ethanol and distilled water several
times to remove any organic residues. In order to obtain a uniform
nano-porous growth, multi-step anodizing was carried out in 0.3 M
sulfuric acid. In the pre-anodizing step, the specimen was submerged and anodized in 0.3 M sulfuric acid at 25 V at near 0 C,
and vigorously stirred for 15 h. Then, the sample was placed in a
mixture of 5 wt.% phosphoric acid and 18 wt.% chromic acid (1:1
by vol) at 50 C for an hour. The patterned specimen was thoroughly rinsed with distilled water several times. Under the same
experimental conditions, anodizing was carried out for an additional hour. Then, pore-widening was accomplished using 5 wt.%
phosphoric acid at room temperature for 15 min. The specimen
was again thoroughly washed with distilled water.
In Fig. 3a, the FESEM (Field Emission Scanning Electron
Microscopy, Hitachi S-4700) image of the NPS created by the
multi-step anodizing fabrication process is shown. The wellordered nano-pores, which are approximately 30 nm in diameter,
appeared on the surface. In Fig. 3b, the micro-scale image of the
NPS by digital microscope (VHX-100, KEYENCE Inc.) is also
provided. The dent-like morphology appeared on the top of the
surface in the test section, which could be caused by the pitting
corrosion under the acidic conditions used.

inner diameter of 229 mm, an outer diameter of 254 mm, and a


305 mm long cylindrical acrylic cell and two Teon end-plates.
The test section was mounted on the side wall of the test chamber
using a specialized tting. Two T-type thermocouples were used to
measure the temperatures of the liquid pool and vapor. Three auxiliary heaters (2000 W each) were installed to provide additional
heating during the degassing process. An O-ring was used on the
Teon plates to ensure a good seal. The test chamber was thoroughly insulated to prevent possible heat transfer from the surrounding area. An AC variable transformer was used to supply a
controlled amount of power to the cartridge heater. Two high
accuracy digital multi-meters (Agilent 34410A and Fluke 87 V,
respectively) were used to measure the voltage and the current
in order to calculate the applied power. To condense the vapor
generated from the boiling chamber, a reux glass condenser and
a constant temperature controller (PolyScience) were installed.

2.2. Experimental set-up of pool boiling


In Fig. 4, the schematic diagram for the experimental pool boiling set-up is shown. The chamber of the pool boiling set-up had an

Fig. 4. Schematic diagram of pool boiling experimental set-up.

C.Y. Lee et al. / Experimental Thermal and Fluid Science 40 (2012) 150158

All of the temperatures in the experimental set-up were monitored


using a data acquisition system (IOtech).

uncertainties in the power input and heat ux were within 8.1%


and 8.6%, respectively.
3. Result and discussion

2.3. Data reduction


The local heat transfer coefcient, h, at each wall thermocouple
location was calculated using the following equation.

Q =A
h
T w  T sat

Where Q, A, Tw, and Tsat are power input to the cartridge heater, heat
transfer surface area, wall temperature, and saturation temperature,
respectively. The input power, Q, supplied to the cartridge heater of
test section was estimated utilizing Eq. (2).

Q VI

Where V and I are applied electric voltage and current,


respectively.
The average heat transfer coefcient, have, measured at a given
heat ux, can be determined by averaging the four local heat transfer coefcients, as seen in the following equation

have

153

htop hleft hright hbottom


4

In this study, in order to evaluate the average heat transfer coefcients, the temperatures measured at each thermocouple location
were used as the wall temperatures in Eq. (1).
2.4. Experimental details
The experimental procedure, in detail, is described below:
 Distilled water was heated to the pre-determined saturation
temperature and boiled to remove dissolved gases.
 The auxiliary cartridge heaters were turned off, and the power
to the cartridge heater of the test section was initiated.
 The voltage of power supply was increased gradually from 20 to
136 V in intervals of 4 V.
 At each increase of voltage, the experimental data was obtained
under the steady state at the atmospheric pressure condition.
The pool temperature for all experiments was maintained within 0.8 C of the saturation temperature. The nominal uncertainty of
the T-type thermocouple was 0.5 C, and the voltage and current
measurements uncertainties were 0.5 V and 0.04 A, respectively.
Based on the method proposed by Kline [14], the estimated

3.1. The rst pool boiling test


The pool boiling curve and heat transfer coefcients, found during the rst test, for the plain surface and the NPS are provided in
Fig. 5. As shown in Fig. 5a, the NPS resulted in the lower wall superheat when compared to the plain surface under the same heat ux
conditions. The Onset of Nucleate Boiling (ONB) appeared at the
wall superheats at approximately 5 C for the plain surface and
approximately 3 C for the NPS. This result implies that the NPS
has the effective cavities to initiate the nucleation at the lower wall
superheat condition. The heat transfer coefcients of both surfaces
became higher with an increase in the heat ux, as shown in
Fig. 5b. The NPS achieved higher nucleate boiling heat transfer
coefcients than the plain surface.
Hsu [15] developed a criterion for the bubble nucleation. The
illustrations of the ebullition cycle model and the inception criteria
are shown in Figs. 6 and 7, respectively. Hsus criteria takes into account the bubble residing on the conical cavity mouth, which is assumed to be a chopped-sphere shape, and assumes it to be a part of
the sphere. According to his criteria, for an embryo to evolve into a
bubble, the temperature surrounding the bubble embryo should
equal or exceed the saturation temperature that corresponds to
the pressure condition inside the bubble. The pressure inside the
bubble is higher than the surrounding liquid, and the pressure difference between the inside and outside of bubble can be expressed
by the YoungLaplace equation for a spherical bubble. Therefore,
based on ClausiusClapeyron and YoungLaplace equations, the
relationship between the equilibrium superheat and the bubble radius can be obtained as follows:

T  T sat

2rT sat
r b qv ifg

Where r, rib, qv, and ifg are surface tension, bubble radius, vapor
density, and latent heat of vaporization, respectively. If the temperature change in the thermal boundary layer with thickness, dt, is assumed to be linear, the temperature prole can be expressed as
below:



y
T  T f T w  T f 1 
dt

Notations, Tf and y, are liquid temperature and distance from the


heating surface, respectively. The temperature proles predicted
by Eqs. (4) and (5) are plotted on the same graph, as shown in

Fig. 5. (a) Pool boiling curve and (b) heat transfer coefcient of rst test for plain surface and NPS.

154

C.Y. Lee et al. / Experimental Thermal and Fluid Science 40 (2012) 150158

Fig. 6. Illustration of ebullition cycle model [16].

Fig. 8. Prediction of active cavity size range using Hsus analysis.

could be explained by a micro-galvanic process between these


phases and the matrix alloy. While the phases often act as local
cathodes, the surrounding aluminum matrix can undergo a localized attack [1719].
Based on Hsus analysis, some micro-cavities formed on the NPS
by unavoidable pitting corrosion may play a role as active nucleation sites, and the NPS can have more micro-cavities than the
plain surface. In addition, the nano-porous structure, with its large
interfacial area, could contribute to enhancing the nucleate boiling
heat transfer coefcient by facilitating the formation of a larger
area to evaporate the liquid micro-layer inside the porous structure [20]. Therefore, it can be interpreted that the nucleate boiling
heat transfer coefcients of a NPS appears higher than those of a
plain surface.

Fig. 7. Inception criteria.

Fig. 7, two points intersected by both Eqs. (4) and (5) indicate the
minimum, rc,min, and maximum, rc,max, values of the mouth cavity
radius, rc. Eq. (6) is then used to predict the size-range of active
nucleation cavities for a given wall superheat.

dt C 2
DT w
fr c;min ; r c;max g
2C 1 DT w DT sub
"
s#
8C 1 rT sat Pf DT w DT sub
 1 1
qv ifg dt DT w 2

Where, C1, C2, DTw and DTsub indicate 1 + cos h, sin h, Tw  Tsat and
Tsat  Tf, respectively. It should be noted that the minus and plus
signs on the right side in Eq. (6) correspond to rc,min and rc,max,
respectively.
In Fig. 8, the prediction of the active cavity size range, employing Eq. (6), is shown. This prediction utilized cavities with microsize dimensions, instead of nano-sized dimensions. These results
suggest, according to Hsus analysis, that bubble nucleation may
not occur in a nano-pore of 30 nm within the present experimental
parameters.
During the anodizing NPS fabrication, the dent-like morphology
appeared on the top of the surface of the test section, as shown in
Fig. 3b. This could be caused by the pitting corrosion under the
acidic conditions used: Acids, such as sulfuric acids, can cause such
morphological changes. It is known that aluminum alloy is vulnerable to pitting corrosion, which is due to a lack of oxygen around a
small area. These oxygen-deprived areas become anodes, while
areas with excess of oxygen become cathodes, leading to localized,
galvanic corrosion. The corrosion penetrates the mass of metal and
leads to the creation of small holes (pitting) on the surface.
Although the passive layer is formed on the surface during the
anodizing fabrication, the metal is still porous and can be susceptible to corrosion in harsh environments (e.g., pH less than 4 or
pH greater than 9). Since most commercial alloys contain several
types of inter-metallic phases, the corrosion on aluminum alloys

3.2. Morphological change of heat transfer surfaces


Fig. 9 shows the changes of the boiling curves for the plain surface and NPS with the number of repeated tests. The time interval
between tests was about 10 h. As the tests were repeated, the boiling curves for both surfaces considerably shifted to the right (at the
higher wall superheat condition), and the ONB occurred at a higher
wall superheat condition. The wall superheats of Test 1 and Test 5
at the ONB were approximately 5 C and 11 C in the plain surface,
respectively. In order to achieve those same results in the NPS, the
temperatures had to be 3 C and 11 C, respectively. In the nucleate
boiling regime for both surfaces, the slope of the boiling curve of
Test 1 was lower than those of other four tests. In Fig. 10, the heat
transfer coefcients of the plain surface and NPS, with the repeated
tests, are shown. For each test, the heat transfer coefcients became higher as the heat ux increased. However, when the number
of repeated tests was increased, the nucleate boiling heat transfer
coefcients for both surfaces decreased. For a natural convection
region, the heat transfer coefcients were insensitive to the number of repeated tests. The average degradations of the nucleate
boiling heat transfer coefcients between Test 1 and Test 5 appeared to be approximately 30% for the plain surface and approximately 37% for the NPS. In the present experimental range, at the
low heat ux region (below approximately q00 = 60 kW/m2), the
degradation of the heat transfer coefcients between Tests 1 and
2 appeared larger than the other tests. The FESEM images revealed
that the changes of the pool boiling heat transfer performance may
be due to the change of boiling surface caused by the chemical
reactions between aluminum and aluminum oxide interacting
with water (i.e., phase changes of aluminum and aluminum oxide).
According to Petrovic and Thomas [21], the reactions of the aluminum in water are plausible. They noted that the aluminum
hydroxide (such as bayerite (Al(OH)3), boehmite (AlO(OH))) and
aluminum oxide (Al2O3) can be formed on the aluminum surface.

C.Y. Lee et al. / Experimental Thermal and Fluid Science 40 (2012) 150158

155

Fig. 9. Boiling curves of (a) plain surface and (b) NPS with repeated tests.

Fig. 10. Heat transfer coefcients of (a) plain surface and (b) NPS with repeated tests.

Fig. 11. FESEM image of plain surface: (a) after Test 1 and (b) after Test 5.

All these reactions are thermodynamically favorable and are exothermic. They also reported that aluminum hydroxide could appear on the surface because of the reaction between the
aluminum oxide and water. Vedder and Vermilyea [22] identied
the three essential steps of growth mechanism of aluminum
hydroxide caused by the chemical reaction between the aluminum
and water: (i) The amorphous oxide formation, (ii) the dissolution
of the amorphous oxide, and (iii) precipitation of the aluminum
hydroxide. Alwitt [23] rened this model and concluded that steps
(i) and (ii) were probably limited by the diffusion of H2O (potentially acting as OH and H+) into the oxide surface. The outer layer
of the aluminum oxide developed under these conditions is typi-

cally converted into aluminum hydroxide through the hydration


process.
In Fig. 11, FESEM images of the plain surface, after Test 1 and
Test 5, are shown. The surface after the rst boiling test was tightly
covered with a sharp, ower-like (vertically overlapping petals)
structure due primarily to the formation of aluminum hydroxide
(in Fig. 11a). As the tests were repeated, the sharp ower-like
structures on the surface become denser, and a part of the test surface turned smooth (in Fig. 11b).
In Fig. 12, FESEM images of the specimens obtained on the NPS
after Tests 1 and 5 are shown. After the rst pool boiling experiment, the surface was greatly transformed from a high, well-or-

156

C.Y. Lee et al. / Experimental Thermal and Fluid Science 40 (2012) 150158

Fig. 12. FESEM image of NPS: (a) after Test 1 and (b) after Test 5.

dered nano-pore structure (in Fig. 3a) to a blunt, ower-like structure (in Fig. 12a). As the number of repeated tests was increased,
the structures became more densely populated; the pores between
the nano-structures were lled and smoothed. (in Fig. 12b).
Based on previous research on the chemical reactions of aluminum and aluminum oxide with water and the surface images
(Figs. 11 and 12), the change of the boiling heat transfer performance on the plain surface and NPS is most likely caused by the
formation of the aluminum hydroxide on both surfaces. The possible reasons are the following: (i) An increase in the thermal resistance due to the aluminum hydroxide and (ii) the transformation
of the boiling surface morphology caused by the phase changes
of aluminum and aluminum oxide.
In order to estimate the thickness of the aluminum hydroxide
formed on the surface, at specimens of the NPS were prepared.
The specimens were then boiled in saturated water for 8 h and
observed using FESEM. The specimens were then boiled for an
additional 42 h and observed again using FESEM. In Fig. 13, the
cross-sectional FESEM images of the NPS at the different conditions, before boiling, after boiling for 8 h and 50 h, are shown.
The before-boiling NPS had a number of vertical columnar
nano-structures, shown in Fig. 13a, and the thickness of the aluminum oxide layer (i.e., nano-porous layer) was approximately
1.5 lm. After the two sets of pool boiling sessions, the aluminum
hydroxide (phase change of aluminum oxide) propagated to
approximately 0.5 lm (in Fig. 13b) and 1.5 lm (in Fig. 13c),
respectively. The growth rate of the aluminum hydroxide on the
NPS was approximately 30 nm per hour. The growth height of
the aluminum hydroxide after 50 h of boiling in a plain surface
was similar to that on the NPS. Based on these observations,
the thermal resistance caused by the formation of the aluminum
hydroxide appears to have little inuence on the nucleate boiling
heat transfer coefcient, since the thickness of the aluminum
hydroxide lm produced would be negligible. Therefore, the
change of the nucleate boiling heat transfer performance in
the plain surface and NPS is attributed to the transformation of
the boiling surface morphology caused by the phase changes of
the aluminum and aluminum oxide. This transformation contributes to the decrease of the active nucleation site density by ooding the cavities.
Fig. 14a and b shows the visualization-results of the pool boiling
using a NPS for Test 1 and Test 5 at q00 = 36.2 kW/m2. It can be
clearly seen that the pool boiling during Test 1 was more vigorous
than Test 5. This nding implies that the NPS of Test 5 had less active nucleation sites at the same heat ux condition. The visualization result of the plain surface was not shown, but was similar to
that of the NPS.
In summary, the aluminum hydroxide, which can be formed on
both surfaces at the water boiling condition, made the active

nucleation sites decrease. This decrease resulted in a decrease in


the nucleate boiling heat transfer coefcient and created an in-

Fig. 13. Cross-sectional FESEM image of NPS: (a) before boiling, (b) after 8 h boiling
and (c) after 50 h boiling.

C.Y. Lee et al. / Experimental Thermal and Fluid Science 40 (2012) 150158

157

Fig. 14. Visualization of pool boiling with NPS for (a) Test 1 and (b) Test 5 at q00 = 36.2 kW/m2.

crease in the wall superheat at the ONB. At the early stage of the
pool boiling experiment for both surfaces (below q00 = 60 kW/m2
for Test 1), the aluminum hydroxide may be formed on the test
sections. Therefore, the slope of the boiling curve for Test 1 in both
surfaces appeared lower than those for the other tests in the nucleate boiling regime (Fig. 9), and the degradation of the heat transfer
coefcient between Test 1 and Test 2 was greater (Fig. 10). Based
on observations, in the rst boiling test results in Fig. 5, the inuence of transformation occurred in a part of heated surface during
degassing process before main boiling test may be included. However, the difference of boiling performance between plain surface
and NPS is attributed to the effect of nano-structure, because in
this work, all experiments were carried out at the same conditions
of degassing time and time interval between tests for plain surface
and NPS.
It is important to note that the morphologies of the aluminum
hydroxide formed on the surfaces during the pool boiling were
the ower-like nano-structures as shown in Figs. 11a and 12a.
The boiling curves of both surfaces consistently shifted to the right
(at the higher wall superheat condition) with the increase of the
number of repeated tests. This nding could imply that solely utilizing the nano-porous structure is insufcient to improve the
nucleate pool boiling heat transfer, and some passages, such as micro-cavities, that are formed inside the porous structure are instrumental in aiding in the escape of vapor [20]. Also, the nucleate pool
boiling heat transfer could potentially be inuenced by the surface
material (e.g., surface wettability), inside structure of the nanoporous layer, the cavities conguration, shape, and size. The
open-connected transformed structure in Figs. 11a and 12a may
not play a role as the active nucleate sites, because they may be
easy to be ooded due to a combination effect of the hydrophilic
surface of aluminum hydroxide and the open-connected structure.
The relationship between nano-structure and nucleate boiling heat
transfer is still not unclear. Therefore, further studies are required
to investigate the effect of the nano-structure on the nucleate boiling heat transfer.
4. Conclusion
In the present experimental study, the change of surface
morphologydue to chemical reactions occurring on the heated
aluminum alloy/aluminum oxide surfaces in a saturated water
pool boiling conditionwas investigated to determine its inuence
on the nucleate boiling heat transfer. Of particular interest are
aluminum alloy (6061) plain surface and the aluminum oxide
NPS surface that was chemically fabricated. The NPS showed a
lower wall superheat at the ONB as well as a higher nucleate
boiling heat transfer coefcient than the plain surface. Through a

series of boiling experiments, both the wall superheats at the


ONB for the plain surface and NPS rose. The nucleate boiling heat
transfer coefcients for the plain surface and NPS decreased by
approximately 30% and approximately 37%, respectively. These
decreases are attributed to aluminum hydroxide formation. The
aluminum hydroxide, formed on both test surfaces, signicantly
changed the surface morphologies and resulted in the degradation
of the pool boiling heat transfer performance by decreasing the
active nucleation sites. The experimental result in this study
reveals that the transformation of surface morphology by the
chemical reaction between the tested surface and the uid is an
important parameter to determine the nucleate boiling heat transfer performance.

Acknowledgments
This material is based upon work supported by the National
Science Foundation under Grant No. (#0923869) via a STTR Phase
II program of Advanced Materials and Devices Inc. (AMAD), Reno,
NV. Special thank goes to Dr. Y. Liu of AMAD.

References
[1] N.H. Kim, K.K. Choi, Nucleate pool boiling on structured enhanced tubes having
pores with connecting gaps, Int. J. Heat Mass Transfer 44 (1) (2001) 1728.
[2] B.J. Jones, J.P. McHale, S.V. Garimella, The inuence of surface roughness on
nucleate pool boiling heat transfer, J. Heat Transfer 131 (2009) 121009-1121009-14.
[3] A.K. Das, P.K. Das, P. Saha, Nucleate boiling of water from plain and structured
surfaces, Exp. Therm. Fluid Sci. 31 (8) (2007) 967977.
[4] W. Nakayama, T. Daikoku, H. Kuwahara, T. Nakajima, Dynamic model of
enhanced boiling heat transfer on porous surfacespart I: experimental
investigation, J. Heat Transfer 102 (3) (1980) 445450.
[5] A. Sloan, S. Penley, R.A. Wirtz, Sub-atmospheric pressure pool boiling of water
on a screen-laminate enhanced surface, in: Proceedings Semi-Therm 2009,
IEEE CFP09SEM-PRT, 2009. pp. 246253.
[6] M.G. Kang, Effect of surface roughness on pool boiling heat transfer, Int. J. Heat
Mass Transfer 43 (22) (2000) 40734085.
[7] A.E. Bergles, Some perspectives on enhanced heat transfer-second generation
heat transfer technology, J. Heat Transfer 110 (4b) (1988) 10821096.
[8] G. Guglielmini, M. Misale, C. Schenone, Boiling of saturated FC-72 on square
pin n arrays, Int. J. Therm. Sci. 41 (7) (2002) 599608.
[9] S. Vemuri, K.J. Kim, Pool boiling of saturated FC-72 on nano-porous surface, Int.
Commun. Heat Mass Transfer 32 (12) (2005) 2731.
[10] S. Launay, A. Fedorov, Y. Joshi, A. Cao, P. Ajayan, Hybrid micro-nano structured
thermal interface for pool boiling heat transfer enhancement, Microelectron. J.
37 (11) (2006) 11581164.
[11] H.S. Ahn, N. Sinha, M. Zhang, D. Banerjee, S. Fang, R.H. Baughman, Pool boiling
experiments on multiwalled carbon nanotube (mwcnt) forests, J. Heat Transfer
128 (12) (2006) 13351342.
[12] I.H. Chaudhri, I.R. McDougall, Ageing studies in nucleate pool boiling of
isopropyl acetate and perchloroethylene, Int. J. Heat Mass Transfer 12 (6)
(1969) 681688.

158

C.Y. Lee et al. / Experimental Thermal and Fluid Science 40 (2012) 150158

[13] C.Y. Lee, M.M.H. Bhuiya, K.J. Kim, Pool boiling heat transfer with nano-porous
surface, Int. J. Heat Mass Transfer 53 (1920) (2010) 42744279.
[14] S.J. Kline, The purposes of uncertainty analysis, J. Fluids Eng. 107 (1985) 153
160.
[15] Y.Y. Hsu, On the size range of active nucleation cavities on a heating surface, J.
Heat Transfer 84 (1962) 207216.
[16] V.P. Carey, LiquidVapor Phase-Change Phenomena, Hemisphere Publishing
Corporation, New York, NY, 1992.
[17] M. Pourbaix, Atlas of Electrochemical Equilibria in Aqueous Solutions, NACE
Cebelcor, Huston, 1974.
[18] K. Nisancioglu, Corrosion of aluminium alloys. in: Proceedings of 3rd
International Conference on Aluminum Alloys, vol. 3, 1992. pp. 239259.

[19] G.M. Scamans, J.A. Hunter, N.J.H. Holroyd, Corrosion of aluminum a new approach,
in: Proceedings of 8th International Light Metals Congress, 1987. pp. 699705.
[20] S. Li, R. Furberg, M.S. Toprak, B. Palm, M. Muhammed, Nature-inspired boiling
enhancement by novel nanostructured macroporous surfaces, Adv. Funct.
Mater. 18 (15) (2008) 22152220.
[21] J. Petrovic, G. Thomas, Reaction of aluminum with water to produce hydrogen:
a study of issues related to the use of aluminum for on-board vehicular
hydrogen storage, DOE (US Department of Energy), 2008.
[22] W. Vedder, D.A. Vermilyea, Aluminum + water reaction, Trans. Farady Soc. 65
(1969) 561584.
[23] R.S. Alwitt, Growth of hydrous oxide lms on aluminum, J. Electrochem. Soc.
121 (1974) 13221328.

Anda mungkin juga menyukai