Anda di halaman 1dari 9

International Journal of Heat and Mass Transfer 55 (2012) 2151–2159

Contents lists available at SciVerse ScienceDirect

International Journal of Heat and Mass Transfer


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

Water droplet evaporation on Cu-based hydrophobic surfaces with nano-


and micro-structures
Chi Young Lee a,b,⇑, Bong June Zhang b, Jiyeon Park b, Kwang J. Kim b
a
KAERI (Korea Atomic Energy Research Institute), 989-111 Daedeok-daero, Yuseong-gu, Daejeon 305-353, Republic of Korea
b
Low Carbon Green Technology Laboratory, Department of Mechanical Engineering, University of Nevada-Reno, Reno, NV 89557, USA

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

Article history: The characteristics of water droplet evaporation on three different hydrophobic surfaces, PCu (Plain Cop-
Received 28 January 2011 per, h = 115°), MSCu (Micro-Structured Copper, h = 126°) and NSCuO (Nano-Structured Copper Oxide,
Received in revised form 30 November 2011 h = 159°) with coating of the same SAM (Self-Assembled Monolayer) material, were experimentally inves-
Accepted 30 November 2011
tigated. For industrial heat transfer applications, copper material was used as the substrate, and the simple
Available online 11 January 2012
and cost-effective fabrication technique to prepare the superhydrophobic surface, NSCuO, was introduced.
Based on the observations, the behavior of droplet evaporation was divided into three stages: Stage I
Keywords:
(constant contact area stage), Stage II (constant contact angle stage) and Stage III (mixed stage). When
Droplet evaporation
Surface structure
studying the PCu surface, the Stages I, II, and III were observed, consistent with previous reports. For the
(Super)hydrophobic surface MSCu surface, Stages I and III appeared without Stage II, and the pinning period of contact line was the
longest among the test samples due to the formation of Wenzel state droplet. In the case of the superhy-
drophobic NSCuO surface, only Stage III occurred, and the contact line moved freely during the entire
evaporation time because of the formation of Cassie state droplet. The total evaporation time of the NSCuO
was the longest out of all the samples tested. At the last stage of evaporation, the edge of the droplet shrank
at a much faster rate in all surfaces. On the other hand, the shrinking velocity of the droplet height drasti-
cally increased only on the NSCuO, which was considered as the unique behavior of superhydrophobic
surface. In this experiment, it was found that the surface structure determines the motion of the contact
line on the surface, which, in turn, strongly influences the characteristics of the droplet evaporation.
Ó 2011 Elsevier Ltd. All rights reserved.

1. Introduction angle stage’’ also known as the moving contact line stage. As op-
posed to Stage I, the contact area radius recedes in Stage II and
The natural evaporation of a droplet on various surfaces is a the contact angle remains constant. The third stage is the ‘‘mixed
fundamental problem and has been studied intensively for a wide stage’’ (Stage III), an unpinned contact line stage where both the
range of industrial and biological applications; e.g., inkjet printing, contact angle and contact area radius decrease. Birdi and Vu [2]
spray painting, DNA chip fabrication, and cell patterning. In gen- and Birdi et al. [3] investigated the evaporation behavior of water
eral, the surface wettability is divided into two groups: the hydro- and n-octane droplets on smooth glass and Teflon surfaces. Both
philic surface which is the surface with a water contact angle below studies reported that the evaporation of a water droplet on a glass
90°, and the hydrophobic surface when the contact angle is above surface and n-octane on a Teflon surface were stationary processes
90°. The characteristics of droplet evaporation on a hydrophobic each with a constant contact radius and linear evaporation rates.
surface are more complicated than those on a hydrophilic surface. However, when the contact angle was over 90°, a constant contact
Picknett and Bexon [1] reported that droplet evaporation on the angle with a decreasing contact radius and a non-linear evapora-
hydrophobic surface occurs in three distinct stages as shown in tion rate was observed. None of these studies [1–3] considered
Fig. 1. The first stage is the ‘‘constant contact area stage,’’ also the effect of the surface morphology on the behavior of droplet
known as the pinned contact line stage (Stage I). As the droplet evaporation, and mentioned the behavior of droplet evaporation
evaporates, the contact angle decreases, while the contact area on a superhydrophobic surface.
radius remains constant. The second stage is the ‘‘constant contact Recently, studies on superhydrophobic surfaces, which exhibit a
water contact angle larger than 150° [4], have been extensively
⇑ Corresponding author at: KAERI (Korea Atomic Energy Research Institute), 989-
performed. A superhydrophobic surface is notable as it has the fea-
111 Daedeok-daero, Yuseong-gu, Daejeon 305-353, Republic of Korea. Tel.: +82 42
tures of water repellency and low surface energy. These two attri-
868 4587; fax: +82 42 8630565. butes have vast potential in various industrial applications such as
E-mail address: chiyounglee@kaeri.re.kr (C.Y. Lee). anti-sticking, self-cleaning, anti-fouling, anti-corrosion, friction

0017-9310/$ - see front matter Ó 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ijheatmasstransfer.2011.12.019
2152 C.Y. Lee et al. / International Journal of Heat and Mass Transfer 55 (2012) 2151–2159

Nomenclature

A surface area [m2] Subscripts


Bo Bond number [–] A apparent contact angle
F force per unit length [N/m] a advancing contact angle
f fractional flat geometrical area [–] D dynamic contact angle
g gravitational acceleration constant [m/s2] E equilibrium (intrinsic) contact angle
H Height of structure [m] e total evaporation time
h droplet height [m] f roughness factor
Dh droplet height change [m] g gas
P pressure [Pa] h height
DP pressure difference [Pa] i initial
R roughness factor [–] l liquid
r contact area radius [m] lg liquid–gas
Dr contact area radius change [m] p projected surface area
s spacing between structures [m] r radius
t time [s] sg solid–gas
Dt time difference [s] sl solid–liquid
V shrinking velocity [m/s] t total (rough) surface area
UY unbalanced Young force
Greek letters
a a half of nano-structure cone angle [°] Superscript

h contact angle [°] normalized
Dq density difference [kg/m3]
r interfacial tension [N/m]

reduction, and heat transfer enhancement [5–9]. Additionally, completely evaporated. Zhang et al. [12] investigated the droplet
researchers [10–14] have investigated the droplet evaporation on evaporation on superhydrophobic lotus leaf and on biomimetic
the (super)hydrophobic surfaces with structures. Shin et al. [10] polymer surfaces. Both hierarchically structured surfaces appeared
examined the evaporation characteristics of a water droplet on to maintain almost a constant contact radius stage during the
hydrophilic, hydrophobic, and superhydrophobic surfaces. They evaporation. Choi and Kim [13] studied on the evaporative process
used glass (h = 58.64°), OTS (octadecyltricholorosilane, of sessile droplets on superhydrophobic surfaces of the sharp-tip
h = 122.52°), and AKD (alkylketene dimmer, h = 160.59°) surfaces post structures with three kinds of heights (i.e., 100, 300 and
as test samples. On the hydrophilic surface, the contact angle, cen- 500 nm) and a given constant pitch of 230 nm. A pure water
ter-height and volume of droplet decreased linearly over the entire and a protein solution were used for testing. They concluded that
evaporation time. The long pinning time – which is defined as the the superhydrophobicity of surface and the behavior of droplet
total time that the contact area radius remains fixed during the evaporation are strongly influenced by the three-dimensional
evaporation – was distinct and predictable. As a droplet evaporated nano-structured morphology and the surface fouling such as pro-
on the hydrophobic surface, typically three distinct stages (see tein adsorption. McHale et al. [14] reported the water droplet
Fig. 1) were observed, but there was no pinning period for an evap- evaporation process on the superhydrophobic SU-8 patterned
orating droplet on the superhydrophobic surface. This study con- polymer surfaces. Based on the diffusion model of water vapor into
cluded that as the hydrophobicity of a surface became stronger, the surrounding atmosphere, they performed the quantitative
the pinning time became shorter and the total evaporation time analysis of initial pinned contact line phase of evaporation, and
lasted longer. In a later study, Shin et al. [11] examined the evap- estimated the diffusion coefficient and the concentration differ-
oration of a sessile water droplet using non-patterned PDMS (poly- ence. Currently, for the fabrication of well-ordered nano-structure,
dimethylsiloxane) and submicron sized post array silicon surfaces, the silicon substrate and MEMS technique are widely used
each with the same hydrophobic contact angle. On the non-pat- [4,7,8,11,13,14], but they are expensive. For industrial applications,
terned PDMS surface, the contact angle showed the three noted, the fabrication technique of a nano-structure should be convenient
distinct stages during evaporation (see Fig. 1). On the patterned sil- and cost-effective, and should have the ability to be applied to
icon post array surface, however, the contact angle decreased line- large surface areas of a conventional metal substrate (e.g., copper).
arly while the contact area remained constant until the droplet Our group has tried to apply the nano-structured surfaces we

Fig. 1. Typical droplet evaporation stages on smooth hydrophobic surface: (a) Constant contact area (Stage I), (b) constant contact angle (Stage II) and (c) mixed (Stage III)
stages.
C.Y. Lee et al. / International Journal of Heat and Mass Transfer 55 (2012) 2151–2159 2153

introduced for heat transfer applications such as the nucleate boil- 2.2. Experimental set-up and details
ing heat transfer, and achieved some successful and meaningful re-
sults [15,16]. Meanwhile, in order to understand how the liquid Using a syringe, a pre-measured de-ionized water sessile drop-
droplet evaporation can be influenced by the surface material let was placed on test specimens. All droplets were of equal weight.
(chemical heterogeneity) and surface structure (roughness), both Contact angles that changed due to natural evaporation were mea-
effects need to be separately examined. It is believed that a com- sured using a CAM-100 (KSV Instruments Ltd., Finland as shown in
parative study of droplet evaporation between the plain and vari- Fig. 4). The temperature and relative humidity of the environment
ous structured surfaces under the similar conditions (e.g., contact were maintained at 23.5 °C and 18%.
angle and surface material) is one effective way to understand In order to determine if the shape of a sessile droplet on test
the interaction of the surface structure on droplet evaporation. surfaces has a spherical cap, the Bond number, which is a dimen-
The objective of this study is to experimentally investigate the sionless number defined as the ratio of gravitational force to sur-
effect of a surface structure on the evaporation characteristics of a face tension force, is calculated using Eq. (1).
water droplet using plain, micro-, and nano-structured hydrophobic
Dqrgh
surfaces. Copper, a popular metal substrate in industrial heat trans- Bo ¼ ð1Þ
fer applications, is used as the test sample. For a superhydrophobic rlg
surface (Nano-Structured Copper Oxide: NSCuO), a convenient and Here, Dq is the density difference between air and water, g is the
cost-effective technique that creates a well-ordered nano-structure gravitational acceleration constant, r is the contact line radius, h is
on a large surface area is introduced. In order to examine the effect the droplet height, and rlg is the water surface tension [17]. The
of the surface structure on droplet evaporation, (i) the same hydro- Bond numbers calculated in the present experiments are in a range
phobic SAM (Self-Assembled Monolayer) coating is applied on all of 0.011–0.38 for PCu, 0.024–0.36 for MSCu, and 0.005–0.33 for
the sample surfaces, removing the effect of a surface material NSCuO. As all Bond numbers are much less than unity, the surface
(chemical heterogeneity) during evaporation; and (ii) the plain tension force dominates the gravitational force.
and structured surfaces with a similar contact angle are prepared. During evaporation, the images of a droplet are recorded every
The characteristics of an evaporating droplet (i.e., contact angle, 1 min on a computer system. From the recorded images of the
contact area radius, droplet height, shrinking velocities of droplet water droplet on the surfaces (as shown in Fig. 5), the contact angle
height and contact radius) are measured on each surface and are re- (h), droplet height (h) and contact area radius (r) are measured
ported on utilizing the droplet imaging measurement technique. using an image processing technique.
Based on the experimental results and the analysis of the surface
structures, the relationship between the droplet evaporation and
the surface structure is able to be reviewed and discussed in detail. 3. Results and discussion

3.1. General trends of water droplet evaporation on PCu, MSCu, and


2. Experiment NSCuO

2.1. Preparation of test samples In Figs. 6 and 7, the images and contours of a droplet evaporated
on PCu, MSCu, and NSCuO surfaces are shown, respectively. In
The following three test specimens were prepared: PCu, indicat- Fig. 7, the numbers on contours indicate time in second.
ing a Plain (non-structured) Copper surface; MSCu, a Micro-Struc- The changes in the contact angle, the contact area radius and
tured Copper surface ground by power tool; and, NSCuO, Nano- the droplet height measured on PCu, MSCu, and NSCuO surfaces
Structured Copper Oxide surface, whose fabrication procedure is during the water droplet evaporation are shown in Fig. 8(a)–(c),
described presently. respectively.
For NSCuO, the initial surface treatment of the copper foil, ob- Fig. 9(a)–(c) show the trends of normalized contact angle, con-
tained from Alfa Aesar (99.98%, thickness 500 lm), power tool tact area radius and droplet height with normalized evaporation
grinding (Grit-120) is used. A specimen, then, is cleaned in acid time, respectively. The normalized parameters are defined in Eq.
(HCl:HNO3 = 1:1 by volume) to remove organic residues. The spec- (2).
imen is placed under ultra-sonication for 10 min and thoroughly hðtÞ rðtÞ hðtÞ t

rinsed with DI water several times. Next, the specimen is cleaned h ¼ ; r ¼ ; h ¼ ; t ¼ ð2Þ
hi ri hi te
in acetone for 10 min under ultra-sonication. Subsequently, the
specimen is immersed in NH4OH solution (0.05 M) for a day at where h, r, h and t are contact angle, contact area radius, droplet
55 °C. The self-assembled copper oxide black coating is then homo- height and time, respectively. Subscripts of i and e indicate initial
geneously applied to the surface. Thorough DI water rinsing is car- value and total evaporation, respectively. A superscript of ⁄ signifies
ried out several times. The specimen is dried in the oven at 90 °C. In a normalized value.
order to remove the chemical heterogeneity on the surface and The initial contact angle of PCu is 115°. As shown in Fig. 8, until
achieve the hydrophobicity, all the samples are coated by SAM in 1250 s, the contact angle steadily decreases to approximately 75°.
a process of immersion in 1.1 mM dodecanethiol for 1 h. By devel- At this angle, the radius of the area between the water droplet and
oping these surfaces and coating the samples, only surface struc- the surface is almost the same as the initial value of 1.44 mm with
tures influence the behavior of droplet evaporation. only a decrease in the droplet height. In this region, the change in
The initial contact angles of all samples and the FESEM (Field contact angle is about 40°. Although the droplet height continues
Emission Scanning Electron Microscopy) images of the MSCu and to decrease, the constant contact angle of approximately 75° is
NSCuO surfaces are shown in Figs. 2 and 3, respectively. The PCu maintained for about 1000 s. In this region, the contact area radius
and MSCu had the initial contact angles of 115° and 126°, respec- does being to decrease. After 2250 s, the contact angle, the contact
tively. Both are hydrophobic surfaces that have similar initial con- area radius, and the droplet height all decrease. The total evapora-
tact angles (the contact angle difference is approximately 11°). It tion time of PCu is 2541 s.
should be noted that NSCuO has the well-ordered nano-structure The present experimental data of PCu is in agreement with the
(see Fig. 3(b)) and becomes a superhydrophobic surface with the trend of previous research [1,10,11]: the behavior of droplet evap-
initial contact angle of 159°. oration for 0–1250 s, 1250–2250 s and 2250–2541 s indicate the
2154 C.Y. Lee et al. / International Journal of Heat and Mass Transfer 55 (2012) 2151–2159

Fig. 2. Contact angle measurement of test specimens: (a) PCu, (b) MSCu and (c) NSCuO.

Fig. 3. FESEM image of test specimens: (a) MSCu and (b) NSCuO.

Fig. 4. CAM-100 contact angle measurement equipment. Fig. 5. Recorded image of water droplet on surface.

constant contact area, constant contact angle, and mixed stages, 80% and 20% of the evaporation time. In other words, when the
respectively. Based on Fig. 9, approximately 50%, 40% and 10% in water droplet on MSCu evaporates, the contact line – a perimeter
total evaporation time occur in Stage I, II, and III, respectively. contacted between air, water droplet, and substrate – is pinned
In the MSCu study, the contact angle steadily decreases during over 80% of the time.
the entire evaporation time without having a region of a constant Although the initial contact angle of MSCu (126°) was similar to
contact angle (i.e., Stage II), as shown in Fig. 8(a). The contact area that of PCu (115°), the detailed behavior of droplet evaporation on
radius remains constant until 2000 s (see in Fig. 8(b)); this behav- both of MSCu and PCu surfaces differs greatly.
ior corresponds to Stage I, and the change in contact angle is about
80°. Then, both the contact angle and the contact area radius de- (i) Stage II, the constant contact angle stage, where the contact
crease (i.e., Stage III). The droplet height decreases over the whole area radius recedes and the contact angle remains constant,
evaporation period. Total evaporation time is 2388 s. As shown in does not occur in MSCu. Birdi and co-workers [2,3] reported
Fig. 9(a) and (b), the portions of Stage I and III are approximately that the constant contact area stage for h < 90° and the con-
C.Y. Lee et al. / International Journal of Heat and Mass Transfer 55 (2012) 2151–2159 2155

Fig. 7. Contours of droplet evaporated on (a) PCu, (b) MSCu, and (c) NSCuO.

ever, Stage II is observed on PCu, consistent with previous


reports, implying that the surface structure is more impor-
tant than the initial contact angle when determining the
droplet evaporation process.
(ii) Stage I of MSCu lasts longer than that of PCu. Shin et al. [10]
reported that, as the hydrophobicity of the surfaces became
Fig. 6. Images of droplet evaporated on (a) PCu, (b) MSCu, and (c) NSCuO. stronger (as the contact angle became larger), the pinning
time became shorter. In this experiment, the MSCu has a
similar (or slightly larger) contact angle compared to PCu,
stant contact angle stage for h > 90° can dominate during the but the MSCu stays at Stage I much longer. Our belief is that
process. Based on the present observations of MSCu, the con- this difference is due to the effect of the surface structure on
tact angle is 126° (>90°), and Stage II does not appear. How- droplet evaporation.
2156 C.Y. Lee et al. / International Journal of Heat and Mass Transfer 55 (2012) 2151–2159

(a) 180 (a) 1.2

150 1.0

Normalized Contact Angle


Contact Angle (deg.)

120 0.8

90 0.6

60 0.4

NSCuO NSCuO
30 0.2
MSCu MSCu
PCu PCu
0 0.0
0 500 1000 1500 2000 2500 3000 3500 0.0 0.2 0.4 0.6 0.8 1.0 1.2
Time (sec) Normalized Time

(b) 2.0 (b) 1.2


NSCuO
MSCu 1.0

Normalized Contact Area Radius


PCu
Contact Area Radius (mm)

1.5
0.8

1.0 0.6

0.4
0.5
NSCuO
0.2
MSCu
PCu
0.0 0.0
0 500 1000 1500 2000 2500 3000 3500 0.0 0.2 0.4 0.6 0.8 1.0 1.2
Time (sec) Normalized Time

(c) 3.0 (c) 1.2


NSCuO NSCuO
2.5 MSCu 1.0 MSCu
PCu PCu
Normalized Droplet Height
Droplet Height (mm)

2.0 0.8

1.5 0.6

1.0 0.4

0.5 0.2

0.0 0.0
0 500 1000 1500 2000 2500 3000 3500 0.0 0.2 0.4 0.6 0.8 1.0 1.2
Time (sec) Normalized Time

Fig. 8. Changes in (a) contact angle, (b) contact area radius and (c) droplet height. Fig. 9. Normalized time vs. (a) normalized contact angle, (b) normalized contact
area radius and (c) normalized droplet height.

The droplet evaporation behaviors of the superhydrophobic during the entire evaporation period. The spherical shape of the
NSCuO are vastly different from those of the PCu and MSCu. The water droplet can prevent the thin liquid film near the edge of
contact angle of the NSCuO gradually reduces due to evaporation droplet from being formed, which leads to a suppression of the
with a decreasing contact area radius and droplet height until evaporation [10]. In the evaporation process of NSCuO, only Stage
3250 s. Then, the contact angle, the contact area radius, and the III is observed and can be further divided into two sub-stages: Slow
droplet height all decrease dramatically. The droplet completely and Fast. Until 3250 s (about 95% of total evaporation time), the
evaporates at 3450 s, which is the longest period of evaporation contact angle and contact area radius slowly and steadily decrease
time among the samples tested in the present study. The slow (referred to as Slow Stage III), then, both the angle and the radius
evaporation is because the droplet maintains its spherical shape drastically shrink (Fast Stage III). The contact line on NSCuO is
C.Y. Lee et al. / International Journal of Heat and Mass Transfer 55 (2012) 2151–2159 2157

(a) 0.012 surface structures. In the following section, this will be discussed
in detail.
NSCuO
0.010 MSCu
Radius Shrinking Velocity (mm/s)

PCu 3.2. Effect of surface structure on water droplet evaporation


0.008
In PCu, the Stages I, II, and III are observed, and the contact line
is pinned during 50% of the total evaporation time. In MSCu, the
0.006
Stages I and III occur, and the contact line is pinned during 80%
of the total evaporation time. In NSCuO, only Stage III appears,
0.004 and the contact line moves freely during the whole evaporation
time. Observing the results, it can be determined that the behavior
0.002 of droplet evaporation is closely related to the motion of the con-
tact line, which, in turn, is influenced by the surface structure.
0.000 When a droplet is placed on the surface, the equilibrium contact
angle is determined by Young’s equation, Eq. (4), which is derived
-0.002 from the force balance acting on the contact line, as shown in
0.0 0.2 0.4 0.6 0.8 1.0 1.2 Fig. 11.
Normalized Time ðrsg  rsl Þ
coshE ¼ ð4Þ
rlg
(b) 0.012
NSCuO Here, rsg, rsl and rlg are interfacial tensions of solid–gas, solid–li-
0.010 MSCu quid, and liquid–gas, respectively. At the initial moment of wetting,
Height Shrinking Velocity (mm/s)

PCu the droplet appears to be at or near the equilibrium contact angle.


0.008 As the droplet begins to evaporate, the droplet height and contact
angle decrease. Therefore, it is no longer the initial equilibrium con-
0.006 dition and the unbalanced Young force can be expressed as below.
F UY ¼ rlg ðcoshD  coshE Þ ð5Þ
0.004
where, hD is the dynamic contact angle by evaporation. According to
Eq. (5), the contact line tends to recede, and the unbalanced Young
0.002
force makes the contact line pinned or de-pinned. An opposing fric-
tional force between the liquid and the solid surface, which leads to
0.000
the pinning of the contact line, is induced by the surface roughness,
and chemical heterogeneities of the solid surface. Whether the con-
-0.002
0.0 0.2 0.4 0.6 0.8 1.0 1.2
tact line is pinning or de-pinning is determined by the competition
between the unbalanced Young force and frictional forces. When
Normalized Time
the unbalanced Young force overcomes the frictional force, the con-
Fig. 10. Normalized time vs. shrinking velocities of (a) radius and (b) height of tact line can be de-pinned [18].
droplet. In general, when a water droplet resides on a rough surface, (i)
the liquid can completely wet the cavities of rough surface or (ii)
the liquid cannot penetrate into the cavities.
not pinned over the entire evaporation time. This trend may be If the droplet wets the surface structure (without an air-pocket
considered as a unique behavior of the superhydrophobic surface. inside the cavity), its apparent contact angle is given by Wenzel’s
In Fig. 10, the shrinking velocities of contact area radius (Vr) and model [19]. Wenzel [19] proposed the relationship between the
height (Vh) of droplet on PCu, MSCu and NSCuO are shown. Vr and apparent (hA) and the intrinsic (hE) contact angles through the
Vh are defined as Eq. (3) modification of Young’s equation to utilize a roughness factor (Rf)
that is defined as a ratio of projection and total surface areas, as
Dr Dh
Vr ¼ ; Vh ¼ ð3Þ follows.
Dt Dt  
The shrinking radius velocities of all the samples are drastically
At rsg  rsl
cos hA ¼ ¼ Rf cos hE ð6Þ
increased at the last stage of droplet evaporation as shown in
Ap rlg
Fig. 10(a). Based on the measurements, the edge of droplets shrink
very fast for the last 10% period of the total evaporation time. How-
ever, the trend of the shrinking droplet height velocity seems to
vary from that of the shrinking radius velocity. At the last evapora-
tion stage, the shrinking velocity of the droplet height in NSCuO
rises more sharply than that of PCu and MSCu, as shown in
Fig. 10(b). This difference is due to the droplet on NSCuO maintain-
ing a spherical shape during the entire evaporation time. This
spherical formation is owing to the low surface energy and the free
moving contact line on the superhydrophobic surface, which is dif-
ferent from that on PCu and MSCu (see Figs. 6 and 7). This peculiar
behavior of water droplet evaporation on the superhydrophobic
surface is unique.
Based on the present observations, the evaporation characteris-
tics of a water droplet on a surface are significantly influenced by Fig. 11. Force balance at contact line during the evaporation [18].
2158 C.Y. Lee et al. / International Journal of Heat and Mass Transfer 55 (2012) 2151–2159

where, At and Ap are the total surface and projection areas, respec- likely too large to be maintained. In this case, the liquid may touch
tively, and, therefore, the value of Rf is greater than unity. Wenzel’s the bottom surface of the cavity, and the surface structures may
equation in Eq. (6) implies that the free energy of the liquid–solid then be flooded. From these findings, we conclude that Cassie and
interface on the rough surface is Rf times larger than that on the Wenzel state droplets are formed on the NSCuO and MSCu,
perfectly smooth surface. The Wenzel’s model of Eq. (6) predicts respectively.
that, if the intrinsic contact angle is larger than 90°, the apparent Two different states of a droplet on the structured surface, the
contact angle increases with an increase in the roughness factor. Wenzel and the Cassie, can affect the behavior of the evaporation.
If the intrinsic contact angle is less than 90°, the roughness factor The stages of droplet evaporation are closely related to whether
leads to a decrease in the apparent contact angle. the contact line is pinned or not. The motion of the contact line de-
On the other hand, the droplet can sit on the peaks of the sur- pends on the droplet states on the structured surface and the dif-
face roughness (with an air-pocket inside the cavity). In this case, ferent frictional forces of both drop states. In the droplet of
a liquid–solid–gas interface is formed, and Wenzel’s model [19] Cassie state on a superhydrophobic nano-structured surface, a slip
can be modified by considering the fractional areas of wetted of liquid can occur due to the air trapped between the nano-struc-
and non-wetted surfaces. The apparent contact angle is given by tures [8]. In this case, the frictional force becomes smaller due to
Cassie and Baxter’s model [20]. the small shear stress and the small contact area on the surface
when compared to the droplet in Wenzel state. Moreover, in the
cos hA ¼ Rf  flg ðRf cos hE þ 1Þ ð7Þ Wenzel droplet state, the surface structure may play a role as the
Here, flg is the fractional flat geometrical area of the liquid–gas defect and severely prevents the contact line from moving. Conse-
interface under the droplet. quently, Cassie state droplet on the nano-structured surface ap-
Two different droplet states, Wenzel and Cassie, can be deter- pears to have a smaller contact angle hysteresis and a weaker
mined by the surface structures and may be related to the detailed pinning of the contact line. In the NSCuO, which exhibits only Stage
behavior of the droplet evaporation. Therefore, it is worthwhile to III, the contact line is observed to move freely, and the shrinking
evaluate the droplet state formed on the each structured surface velocity of droplet height drastically increases at the last stage of
we used. evaporation due to the formation of Cassie droplet state. The
Choi et al. [8] proposed the criteria of structure spacing (s) and MSCuO has the longest period of Stage I, where the contact line
height (H) for liquid meniscus to withstand the pressure of liquid is pinned in test samples due to the formation of the Wenzel drop-
using a simple geometrical calculation and the Laplace–Young let state.
equation, as follows: In summary, the surface morphology has a great influence on
the behavior of droplet evaporation. The surface structure deter-
pffiffiffi j cosðha  aÞj mines the state of the droplet placed on the surface, and this affects
s < 2 2rlg ð8Þ
DP the motion of the contact line. The behavior of the contact line (i.e.,
whether it is pinned or de-pinned) results in a change of the details
1  sinðha  aÞ in droplet evaporation, for example, a change of the contact angle,
H > pffiffiffi s ð9Þ
2j cosðha  aÞj contact area radius, droplet height, shrinking velocities of droplet
radius and height.
The idealized surface structure in Fig. 12 is different from that in the
present surface structure, but Eqs. (8) and (9) may be a good indica-
tor when evaluating the spacing and height of structure for Cassie 4. Conclusion
drop. In order to estimate the drop states on the present structured
surfaces, the spacing and height of the structure were calculated at In the present experimental study, the characteristics of droplet
the following condition [8]: rsl = 0.0727 N/m, ha  a = 120° and DP evaporation on three kinds of hydrophobic surfaces were investi-
(=Pl  Pg) = 1 atm. It is figured that the spacing between the struc- gated utilizing PCu, MSCu and NSCuO surfaces, all coated by the
tures is less than 1.0 lm, and the height of structures is larger than same hydrophobic SAM material.
0.2 lm. Based on Fig. 3(b), the NSCuO structure is satisfied with the In the PCu experiment, Stages I, II, and III were observed, which
s < 1.0 lm and H > 0.2 lm proposed by Eqs. (8) and (9). However, was consistent with previous reports. In the MSCu experiment,
the MSCu structure does not seem to meet the criteria. The struc- Stages I and III were observed without Stage II – in spite of a similar
ture spacing in y-direction is above about tens of lm, and the inter- contact angle to the PCu experiment. The pinning period of the
face curvature of x-direction in Fig. 3(a) between liquid and gas is contact line was longest (80% of total evaporation time) in the
MSCu test sample. This was due to the Wenzel state droplet that
was formed on the MSCu surface which prevents the contact line
from moving due to the large frictional force. In the NSCuO exper-
iment, only Stage III occurred. During the total evaporation time,
the contact line was able to move freely due to the sliding effect
of the contact line caused by the formation of the Cassie state drop-
let. The total evaporation of NSCuO lasted the longest out of the
three tested surfaces. While the shrinking velocity of the contact
area radius drastically increased for all surfaces at the last evapo-
ration stage, the droplet height shrunk sharply for only the NSCuO
surface, which was considered as one of the unique evaporation
behaviors of the superhydrophobic surface. These findings are pre-
sumed to be caused by the droplet on such a surface maintaining
its spherical shape.
In this study, the superhydrophobic surface can be fabricated
utilizing simple and cost-effective techniques to create well-con-
trolled nano-structures on a large surface area. It is found that
Fig. 12. Gas–liquid interface on an idealized structured hydrophobic surface. the surface structure is a more important factor than the initial
C.Y. Lee et al. / International Journal of Heat and Mass Transfer 55 (2012) 2151–2159 2159

contact angle when determining the characteristics of droplet [7] David Qúeŕe, Non-sticking drops, Rep. Prog. Phys. 68 (11) (2005) 2495–2532.
[8] C.H. Choi, C.J. Kim, Large slip of aqueous liquid flow over a nanoengineered
evaporation. This study implies that the modification of surface
superhydrophobic surface, Phys. Rev. Lett. 96 (2006) 066001.
structure is one of ways to control the evaporation rate in the heat [9] J.B. Boreyko, C.H. Chen, Self-propelled dropwise condensate on
transfer application. superhydrophobic surfaces, Phys. Rev. Lett. 103 (2009) 184501.
[10] D.H. Shin, S.H. Lee, J.Y. Jung, J.Y. Yoo, Evaporating characteristics of sessile
droplet on hydrophobic and hydrophilic surfaces, Microelectron. Eng. 86 (4–6)
Acknowledgements (2009) 1350–1353.
[11] D.H. Shin, S.H. Lee, C.K. Choi, S. Retterer, Evaporation and wetting dynamics of
This material is based upon work partially supported by the Na- sessile water droplets on submicron-scale patterned silicon hydrophobic
surfaces, J. Micromech. Microeng. 20 (5) (2010) 055021.
tional Science Foundation under Grant No. (#0923869) via a STTR [12] X.Y. Zhang, S.X. Tan, N. Zhao, X.L. Guo, X.L. Zhang, Y.J. Zhang, J. Xu, Evaporation
Phase II program of Advanced Materials and Devices Inc. (AMAD), of sessile water–droplet on natural lotus and biomimetic polymer surface with
Reno, NV (subcontracted to the University of Nevada, Reno; PI: superhydrophobicity, Chem. Phys. Chem. 7 (2006) 2067–2070.
[13] C.H. Choi, C.J. Kim, Droplet evaporation of pure water and protein solution on
KJK). Special thanks go to Dr. Y. Liu of AMAD. Also, Nano-Struc- nanostructured superhydrophobic surfaces of varying heights, Langmuir 25
tured Copper Oxide structures were developed under a DOE Grant (2009) 7561–7567.
DE-EE0003231 to the University of Nevada, Reno (PI: KJK). [14] G. McHale, S. Aqil, N.J. Shirtcliffe, M.I. Newton, H.Y. Erbil, Analysis of droplet
evaporation on a superhydrophobic surface, Langmuir 21 (2005) 11053–
11060.
References [15] C.Y. Lee, M.M.H. Bhuiya, K.J. Kim, Pool boiling heat transfer with nano-porous
surface, Int. J. Heat Mass Transfer 53 (19–20) (2010) 4274–4279.
[1] R.G. Picknett, R. Bexon, The evaporation of sessile or pendant drops in still air, J. [16] C.Y. Lee, B.J. Zhang, K.J. Kim, Morphological change of plain and nano-porous
Colloid Interface Sci. 61 (2) (1976) 336–350. surfaces during pool boiling and its effect on nucleate boiling heat transfer,
[2] K.S. Birdi, D.T. Vu, Wettability and the evaporation rates of fluids from solid Exp. Therm. Fluid Sci. (Under review).
surface, J. Adhesion Sci. Technol. 7 (6) (1993) 485–493. [17] H. Hu, R.G. Larson, Evaporation of a sessile droplet on a substrate, J. Phys.
[3] K.S. Birdi, D.T. Vu, A. Winter, A study of the evaporation rates of small water Chem. B 106 (2002) 1334–1344.
drops placed on a solid surface, J. Phys. Chem. 93 (9) (1989) 3702–3703. [18] L. Shi, P. Shen, D. Zhang, Q. Lin, Q. Jiang, Wetting and evaporation behaviors of
[4] B. Bhushan, Y.C. Jung, Wetting study of patterned surfaces for water–ethanol sessile drops on PTFE surfaces, Surf. Interface Anal. 41 (2009)
superhydrophobicity, Ultramicroscopy 107 (10–11) (2007) 1033–1041. 951–955.
[5] H. Zhang, R. Lamb, J. Lewis, Engineering nanoscale roughness on hydrophobic [19] R.N. Wenzel, Surface roughness and contact angle (letter), J. Phys. Coll. Chem.
surface-preliminary assessment of fouling behaviour, Sci. Technol. Adv. Mater. 53 (9) (1949) 1466.
6 (3–4) (2005) 236–239. [20] A. Cassie, S. Baxter, Wettability of porous surfaces, Trans. Faraday Soc. 40
[6] G.R.J. Artus, S. Jung, J. Zimmermann, H.P. Gautschi, K. Marquardt, S. Seeger, (1944) 546–551.
Silicone nanofilaments and their application as superhydrophobic coatings,
Adv. Mater. 18 (20) (2006) 2758–2762.

Anda mungkin juga menyukai