You are on page 1of 15

Modeling and Optimization of

Superhydrophobic Condensation
Superhydrophobic microlnanostructured sutfaces for dropwise condensation have recently
Nenad Miljkovic received significant attention due to their potential to enhance heat transfer peiformance by
Department of Mechanical Engineering, shedding water droplets via coalescence-induced droplet jumping at length scales below
Massacliusetts institute ot Technoiogy,
the capillary length. However, achieving optimal surface designs for such behavior requires
77 Massactiusetts Avenue,
capturing the details of transport processes that is currently lacking. While comprehensive
models have been developed for flat Hydrophobie surfaces, they cannot be directly applied
Cannbridge, MA 02139
for condensation on micro/nanostructured surfaces due to the dynamic droplet-structure
interactions. In this work, we developed a unified model for dropwise condensation on
Ryan Enright superhydrophobic structured surfaces by incorporating individuell droplet heat transfer,
Department oí Mectianicai Engineering, size distribution, and wetting morphology. Two droplet size distributions were developed,
Massactiusetts institute ot Tectinoiogy, which are valid for droplets undergoing coalescence-induced droplet jumping, and exhibit-
77 Massactiusetts Avenue, ing either a constant or variable contact angle droplet growth. Distinct emergent droplet
Cambridge, iViA 02139; wetting morphologies, Cassie jumping, Cassie nonjumping, or Wenzel, were determined by
Stoi<es Institute, coupling of the structure geometry with the nucleation density and considering local energy
University ot Limerici<, barriers to wetting. The model results suggest a specific range of geometries (0.5-2 ßm)
Limerick, ireland allowing for the formation of coalescence-induced jumping droplets with a 190% overall
surface heat flux enhancement over conventional flat dropwise condensing surfaces. Subse-
Evelyn N. Wang^ quently, the effects offour typical self-assembled monolayer promoter coatings on overall
Department ot Mectianicai Engineering, heatflta were investigated. Surfaces exhibiting coalescence-induced droplet jumping were
Massachusetts institute ot Technoiogy, not sensitive (<5%) to the coating wetting characteristics (contact angle hysteresis), which
77 Massachusetts Avenue, was in contrast to surfaces relying on gravitational droplet removal. Furthermore, flat
Cambridge, MA 02139 surfaces with low promoter coating contact angle hysteresis (<2 deg) outperformed struc-
e-mail: enwang@mit.edu tured superhydrophobic surfaces when the length scale of the structures was above a cer-
tain size (>2 ¡xm). This work provides a unified model for dropwise condensation on micro/
nanostructwed superhydrophobic surfaces and offers guidelines for the design of structured
surfaces to maximize heat transfer. [DOI: 10.1115/1.4024597]

Keywords: superhydrophobic condensation, jumping droplets, droplet coalescence,


condensation optimization, environmental scanning electron microscopy, microlnano-
scale water condensation, condensation heat transfer

1 Introduction number of small droplets [14] which efficiently transfer the major-
Dropwise condensation has received significant attention since ity of the heat from the surface [8,15-17]. In addition, gravity inde-
its discovery in 1930 by Schmidt et al. [1], due to the superior pendent droplet removal allows utilization of such surfaces in any
heat transfer performance over conventional filmwise condensa- orientation, unlike conventional passive condensing surfaces which
tion [2-7]. The high performance enhancement of passive drop- require gravity for droplet removal, a severe limitation for mobile
wise condensing surfaces is attributed to their ability to form electronics and space appfications [18]. Surface structuring may
discrete nonwetting droplets which, upon growing to a critical therefore be an ideal method to enhance condensation heat transfer.
size ('^lmm), can shed from the surface by gravity, thereby While several groups have demonstrated that properly designed
reducing the overall thermal resistance compared to filmwise con- surfaces can enable stable superhydrophobic condensation
densation. More recently, micro/nanostructured superhydrophobic [19-23], a unified modeling framework to optimize structure
surfaces have been investigated as potential dropwise condensing design that captures the dynamic phase change process is lacking
surfaces for energy conversion [8], atmospheric water harvesting [24]. The early dropwise condensation model of Le Fevre and
[9,10], and high heat fiux thermal management applications [11] Rose [25,26] combined individual droplet heat transfer with drop-
owing to the promise of further improvements in overall heat let size distribution theory. Advanced models have since followed
transfer performance compared to traditional dropwise condensa- this work by including more accurate expressions for the growth
tion. Indeed, a recent study showed that when small droplets of small droplets. Tanaka [27] used population balance theory to
(~10-100,um) merge on superhydrophobic structured surfaces, evaluate the local droplet size by taking into account the two
they can spontaneously jump via the release of excess surface mechanisms of growth: direct vapor accommodation onto the
energy independent of gravity [12]. This phenomenon is attributed droplet and coalescence with neighboring droplets. As a result,
to surface structuring, which enhances the hydrophobicity, and better predictions of the droplet size distribution for small noncoa-
thereby decreases droplet pinning to the surface [13]. Droplet re- lescing droplets were obtained. Abu-Orabi [28] further refined the
moval by this mechanism is highly desirable due to the increased dropwise condensation model by considering all possible thermal
resistances associated with the droplet. More recently, the work of
Kim and Kim [29] extended the previous models by determining
Corresponding author. more accurately the conduction resistance for droplets exhibiting
Contributed by the Heat Transfer Division of ASME for publication in the large contact angles (9 > 90 deg).
JOURNAL OP HEAT TRANSFER. Manuscript received March 31, 2012; final manuscript
received September 2, 2012; published online September 23, 2013. Assoc. Editor: Despite significant developments on dropwise condensation
Sujoy Kumar Saha. modeling, predictive models for condensation on micro/

Journal of Heat Transfer Copyright © 2013 by ASME NOVEMBER2013, Vol. 135 / 111004-1
nanostructured superhydrophobic surfaces are still lacking. Specifi- can exhibit three distinct wetting morphologies: S where con-
cally, three main inconsistencies arise when applying previous densed droplets sit on top of the micro/nanostructure (Eig. l(c))
models to condensation on structured surfaces: (1) Droplet wetting [30], PW where the droplets form a liquid bridge connecting the
morphology cannot be predicted (i.e., Cassie [30], Wenzel (W) base of the droplet (Eig. l{d)) [24], or W where droplets wet the
[31], suspended (S) [24], or partially wetting (PW) [24,32]), (2) cavities ofthe micro/nanostructure (Eig. l(e)) [31].
droplet contact angle can vary during droplet grovrth [14,32,33], To accurately predict the wetting morphology of a single drop-
and (3) droplet size distribution is not valid for droplets with non- let, a nonequilibrium thermodynamic energy criterion is used
constant contact angles and with surfaces undergoing coalescence- which emphasizes the role of contact line pinning on the resultant
induced droplet jumping as the main mode of droplet removal. droplet morphology [22]. By comparing the dimensionless energy
This work develops a model framework to predict dropwise of the advancing Wenzel (cos of = r cos öa) [31] and Cassie
condensation heat transfer for micro/nanostructured superhydro- (cos Ö™ = — 1) [36] droplet morphologies, the expected morphol-
phobic surfaces. Pillar arrays, spanning a wide range of dimen- ogy can be estimated by
sions (~10nm-10/^im), are utilized as the model structured
surface. The current model incorporates prediction capability for coso'iCB
the emergent droplet wetting morphology (Sec. 2), accounts for E* = (1)
COS0" r cos oa
nonconstant contact angle droplet growth (Sec. 3), and extends
the previously developed droplet size distribution theory to both When E*>1 the contact line can overcome the energy barrier to
constant and nonconstant contact angle droplets growing on surfa- depin and a W droplet is formed (Eig. l(e)). If £* < 1 complete
ces experiencing coalescence-induced droplet jumping [12,14] depinning is not possible and the droplet grows upward over
(Sec. 4). The results from the model are subsequently used to the top of the pillar array forming a PW Cassie droplet (Eigs. l(c)
study the effects of surface structure design and size scale (Sec. and l{d)).
6), surface inclination (Sec. 7), and contact angle hysteresis with Droplet coalescence introduces a further length scale depend-
different promoter coatings (Sec. 8) on overall surface heat trans- ency on the emergent droplet wetting morphology. When the av-
fer. Optimization of the structure geometry indicates that surfaces erage condensing droplet spacing (4) approaches that of the pillar
with nanometer length scales and high nucleation densities have spacing /((/c) ~ I), the formation of local pinning barriers is dis-
the potential to enhance overall condensation heat transfer per- rupted. While it may be energetically favorable to form Cassie
formance by 190% when compared to conventional flat dropwise droplets (E* < 1), this may not be possible due to droplet coales-
condensing surfaces. The outcomes of this work create important cence between two adjacent unit cells which bypasses the pinning
regime maps and design guidelines for highly efficient superhy- barriers associated with the pillar sides. This effect results in the
drophobic condensation surfaces. formation of thermodynamically unfavorable W droplets which
are highly pinned to the surface. To avoid formation of W droplets
on Cassie-Baxter (CB) stable surfaces, a critical droplet separa-
2 Droplet Wetting Morphology Model tion distance of at least {k) > 21 must be maintained [22].
During dropwise condensation on a structured superhydropho- Eor the purposes of this model, condensation on the structured
bic surface, droplets can depart either by coalescence-induced surface is assumed to be spatially random. The mean condensing
droplet jumping (if droplet/surface adhesion is small) or by grav- droplet spacing (4) can be related to the nucleation density Ns
ity (if droplet/surface adhesion is large). The individual droplet by [24]
contribution to the heat transfer process is highly sensitive to the
droplet wetting morphology and therefore is necessary to predict 1
the overall surface heat transfer by dropwise condensation.
(4)=- (2)
To study the effects of surface structure geometry on emergent
condensing droplet wetting morphology, we consider a model These two wetting criteria (Eqs. (1) and (2)) have been vali-
structured surface consisting of a micro/nanopillar array with dated experimentally for a wide variety of structured surfaces
heights h, diameters d, and pillar-to-pillar spacings / (solid frac- with a range of length scales and surface energies [2,4,22,24]. The
tion (p = nd^/AP' and roughness factor r = 1 -|- ndh/P) (Eigs. criteria form the basis for determining the emergent condensing
l(a) and 1(0)). To achieve superhydrophobicity, the model surface droplet morphology on the micro/nanostructured surface. How-
is assumed to be coated with a promoter film such as a self- ever, to gain an understanding of overall surface condensation
assembled-monolayer (SAM) having intrinsic advancing/receding performance, the individual droplet heat transfer for each mor-
contact angles of dJ9r [34,35]. During condensation on a struc- phology needs to be incorporated.
tured superhydrophobic surface, the condensing liquid droplets
3 Droplet Heat Transfer Model
To accurately model dropwise condensation on micro/nano-
structured surfaces, individual droplet growth rates and heat trans-
fer are needed. The growth behavior of each droplet morphology
can be obtained by modifying the Kim and Kim model [29] to
account for the pillar geometry and emergent droplet morphology
[24]. It is important to note that most previous models assume a
constant droplet contact angle, 9, dudng growth. This assumption
is appropriate for dropwise condensation on flat hydrophobic
surfaces [28,29,37]; however, it does not apply for structured
superhydrophobic surfaces, since droplets have been observed to
have variable contact angles during growth [14,18,24,32,33,38].
(c) (d) To resolve this discrepancy, a model was developed to predict the
droplet contact angle 0 as a function of the droplet radius R.
Fig. 1 Schematics of the modeled structured surface showing
(a) side view and (b) top view of the characteristic structure
dimensions h, d, and / representing the pillar height, diameter 3.1 Contact Angle Modeling. The droplet contact angle 6
and center-to-center spacing, respectively. Schematics show- was modeled as a function of the droplet wetting morphology (PW,
ing the (c) S, (d) PW, and (e) W morphologies. W, or S) and droplet radius R. When the droplet radius R is lower

111004-2 / Vol. 135, NOVEMBER 2013 Transactions of the ASME


than the structure scale (R < T), the droplet is assumed to have the tact angle oscillates about a mean value with an amplitude as high
intrinsic hydrophobic surface coating contact angle 6.^,. When the as~15deg at length scales R/l-^ 1-3, the oscillation amplitude
droplet grows to a size comparable to the structure scale {R ~ /), it will decay at larger sizes (R/l > 3) as the droplet grows laterally
begins to interact with the pillars confining it and fills the unit cell during condensation [22]. Due to the relatively small amplitude of
to the top of the structures. At this point, depending on the energy oscillation and decay, we neglected the contact angle deviation to
criterion F* and nucleation density A's, the droplet can propagate simplify the analysis. In addition, the magnitude of the W droplet
above the unit cell and form a PW droplet, or it can propagate later- contact angle oscillation (~l-15deg) is less than that of the PW
ally and form a W droplet (Sec. 2). In addition, due to the spatially droplet variable contact angle difference which can approach
random nature of nucleation, droplets nucleating on Cassie stable 90deg.
surfaces can do so on the tips of pillars to form S droplets. The contact angle behavior of the PW, W, and S droplet mor-
The PW droplet contact angle was modeled as a nonlinear func- phologies can be summarized by
tion of droplet radius R [2,4,24]. Previous experimental studies
using environmental scanning electron microscopy (ESEM) and R<1
optical microscopy revealed the characteristic diameter of the
pinned neck of the condensing PW droplets is approximately (3)
2 x 2 unit cells [22,24]. At later times, the PW droplet has an
increasingly large apparent contact angle öpw characteristic of the e?, R>
Cassie morphology (Fig. 2(b)) [2,24]. , R<1
Wenzel and S droplets were modeled as having a constant con- (4)
', R>1
tact angle characteristic of the W and S morphologies (Figs. 2{a)
/iCB
and 2(c)). Although previous studies have shown that the W con- Ôs{R) = (5)

Fig. 2 Time-lapse schematics of (a) S, (b) PW, and (c) W dropiet morphoiogies
during growtii on the structured surface. To the right of the schematics are corre-
sponding environmentai scanning electron microscopy (ESEiM) images of dropiets
with the different morphoiogies on a nanostructured surface (/7 = 6.1 ftm, t = 2fim,
d= 300 nm) [23,24]. Schematics not to scaie.

Journal of Heat Transfer NOVEMBER2013, Vol. 135 / 111004-3


The temperature drop between the saturated vapor and liquid
interface (ATj) is given by

(7)
' /!i27l/?2(l -COSS)

where q is the heat transfer rate through the droplet and hi is the
condensation interfacial heat transfer coefficient given by [42,43]
Pillars
hi
(8)

(C) where /fg is the specific gas constant and Vg is the water vapor spe-
cific volume. The condensation coefficient a is the ratio of vapor
molecules that will be captured by the liquid phase to the total
number of vapor molecules reaching the liquid surface (ranging
from 0 t o i ) .
Once the vapor condenses on the droplet interface, the latent
heat must be conducted through the droplet to the substrate. This
resistance is modeled as a pure conduction resistance which leads
to a temperature drop (AT¡i) given by [29]

Fig. 3 (a) Schematic of the dropiet on the condensing surface (9)


growing in the PW morphoiogy. (b) Dropiet thermai resistance ^ sin ö
diagram showing the iiquid-vapor interface (R/), dropiet con-
duction (Ra), hydrophobic promoter coating {Rhc), piilar (/7p), where Tbi is the liquid temperature of the droplet base (Fig.
and gap (flg) thermai resistances, (c) Thermai resistance net-
work in the dropiet and piilar structure. The schematic shows and A'w is the condensed water thermal conductivity. The tempera-
the paraiiei path of heat f iowing through (i) the coating (/7HC) fol- ture drop due to the promoter coating is calculated using a con-
iowed by the piilar (Rp) and (ii) the iiquid bridge (Rg) foiiowed by duction resistance given by
the coating (HHC)- Schematics not to scale.
"HC
(10)
where ^^ = (p{cos9^ + \ ) - 1, /?max = / / s i n ( 7 t - and uc sin 6

where T^^o is the temperature of the silicon pillars beneath the


3.2 Droplet Growth Modeling. At the scales considered in coating (Fig. 3{b)), ¿HC is the coating thickness, cp is the struc-
this work (~1 ßvn), the dominant mode of droplet growth is due to tured surface solid fraction, and k^c is the coating thermal
the direct accommodation of vapor molecules at the droplet inter- conductivity.
face [39]. For a droplet with radius R(t) on a structured superhy-
The conduction resistance through the pillars is dependent on
drophobic surface (Fig. 3{a)), the contact angle 9 varies with the
the wetting morphology of the droplet. For the S morphology, the
droplet radius according to Eqs. (3)-(5). The local vapor (Tsat) and
temperature drop associated with the conduction resistance is
surface (TJ temperatures are assumed to be constant throughout
given by
the growth process. The individual droplet heat transfer, q, is
determined by considering all thermal resistances from the satu-
rated vapor through the condensing droplet to the substrate qh
(Fig. 3). All thermal resistances associated with the droplet are = 7*2 —Ts — (11)
presented in terms of individual temperature drops: the liquid-
vapor interfacial resistance due to direct vapor molecule accom-
modation at the droplet interface (ATi), the conduction resistance where T^ is the substrate temperature, and kp is the piUar thermal
through the droplet {AT¿), the conduction resistance through the conductivity.
pillars (ATps) or liquid bridge and pillars (ATppw), the promoter For PW droplets, the conduction resistance temperature drop
coating resistance (ATHC), and the resistance due to the curvature through the pillar and coating structure is calculated by consider-
of the droplet (ATc). Intemal droplet convection was neglected in ing a parallel heat transfer pathway from the base of the droplet to
the model since the droplets were sufficiently small so that con- the substrate surface (Fig. 3(c)) given by
duction is the primary mode of heat transfer through the droplet
[27,40].
The temperature drop due to droplet curvature (Arc) is given
by [41] kp(p
(12)
9 [<5HC^P +
(6)
It is important to note that the PW conduction temperature drop
given by Eq. (12) becomes the S temperature drop when
where T^( is the water vapor saturation temperature, a is the water ¿w = /:v ~ 0 W/mK, where k^ is the water vapor thermal conductiv-
surface tension, hfg is the latent heat of vaporization, p„ is the liq- ity. In this case, there is no liquid bridge available for heat flow.
uid water density, and R^in is the droplet nucleation radius Accounting for all of the temperature drops, the individual
droplet heat transfer rate is

111004-4 / Vol. 135, NOVEMBER 2013 Transactions of the ASME


(13)
1 kp(p
2/!i(l-COS0) sino + ¿HC*W

The droplet heat transfer is related to the droplet growth rate {dRI 4.1 W and Flat Hydrophobic Surface Droplet Size
dt) by the latent heat of phase change Distribution. Structured superhydrophobic surfaces exhibiting W
droplet growth have a droplet size distribution analogous to that
in Q)^ñih = h ~ of a flat hydrophobic surface. W droplets grow and merge until
'" dt reaching a size characteristic of the capillary length before being
swept off by gravity. Sweeping droplets roll down the surface and
= -p„hfc,— \{l-cos9)(2 + cos6)R^\ (14) remove all droplets in their path, cleaning the condensing surface
3 ai L J for new droplets to renucleate. The droplet sweeping mechanism
allows for small, more effective, droplets to populate the surface
Differentiating Eq. (14), we obtain an explicit term for dR/dt and thereby minimizes the condensation heat transfer resistance.
Due to the significant droplet/surface adhesion exhibited by W
q{R, 6) = ;ip^Afg/?2^ droplets, coalescence-induced droplet removal is not possible,
dt dR rather gravity assisted sweeping of droplets from the surface dom-
inates [12,14,22,24].
+ (1-cos0)^(2 +cos (15) For small W droplets undergoing noninteracting growth, the
population balance theory can be used to determine the droplet
size distribution [27-29]. The steady-state W droplet size distribu-
Equation (15) has been shown to have excellent agreement with
tion is determined from the conservation of number of droplets
experimental results for a variety of differing droplet morpholo-
entering a size range containing droplets with radii Ri to /?2, i.e.,
gies including PW, S, W, and droplets growing on flat surfaces
the number of droplets entering this size range must equal to
[2,24]. In order to determine the overall surface heat transfer per-
the number of droplets leaving. The W droplet growth rate is
formance, the individual droplet heat transfer must be combined
defined as
with the morphology dependent droplet size distribution.

4 Droplet Size Distribution Theory dR


G= (16)
'dt
On a flat hydrophobic surface, droplet nucleation and growth
proceeds through two mechanisms; (1) direct (noninteracting)
growth where droplets nucleate and grow on spatially random The number of droplets entering the size range {R1-R2) in a
high energy sites by direct deposition from the vapor onto the time increment dt is An^Gidt, where A is the surface area and «1
droplet surface [22], and (2) coalescence-dominated growth where represents the number of droplets of size Rj. Similarly, the num-
the distances between neighboring droplets (4) become smaller ber of droplets leaving the size range is An2G2dt. The number of
and coalescence occurs until the droplet is large enough to be droplets swept off the surface by droplet shedding is equal to
swept off the surface by gravity [16]. The falling droplet is able to Sni.2dRdt, where S is the sweeping rate at which the substrate sur-
sweep away the droplets beneath it and clean the condensing sur- face is renewed by falling droplets, and «1.2 is the average popula-
face so that new droplets can renucleate. As a result of this droplet tion density in the size range. Applying droplet conservation, we
growth from nucleation to departure, a wide range of droplet sizes obtain [28,29]
exist on the condensing surface [11,29].
Due to the dynamic nature of droplet growth on structured
superhydrophobic surfaces [22,24], the flat surface growth and (17)
departure mechanisms are not adequate to accurately predict the
droplet size distribution. Three main inconsistencies arise: (1) In the limit of dR approaching zero, Eq. (17) can be reduced to
Droplet departure on structured superhydrophobic surfaces may
occur via coalescence-induced droplet jumping as size scales
well below the capillary length; (2) droplets can have a range of djGn) ^ « _
(18)
wetting morphologies (Figs. l(c) and l(e)) depending on the dR T
structure geometry and size scale (Sec. 2); and (3) droplet con-
tact angles may not be constant during growth depending on the where T is the sweeping period (T=A/S).
emergent droplet wetting morphology (Sec. 3.1). In light of these The solution to Eq. (18) can be determined by first solving for
differences, new morphology dependent droplet size distribu- the droplet growth rate G. Relating the latent heat of phase change
tions were derived that are valid for structured superhydrophobic (Eq. (15)) to the individual droplet growth (Eq. (13)), G for W
surfaces. droplets is

AT--
G= (19)
- cos 0)^(2 + cos 0) 1 RB 1
2/zi ( 1 - cos 0) sin 0 in^ 0 [pnckp + hknc

Journal of Heat Transfer NOVEMBER2013, Vol. 135 / 111004-5


"min
1 - Ns=1x10'°
(20) 10I Ns=1x10"
Ns= 1x10"
Ns = 1x10"

where

AT
(21)
- cos ö)-(2 + cos 0)

(22)
, sin 0 10^'
0.1 1 10 50
1 R(nm)
2/îi(l-cos0)
Fig. 4 Droplet size distribution for a surface with droplet re-
moval by gravity (flat hydrophobic surfaces) as a function of
•-I-- (23) droplet radius R for various nucleation densities /Vs. The popu-
Â:HC sin
lation density is shown for small droplets (/7(R)) with color
curves and large coalescing droplets {N{PÍ¡) with the black
Solving Eq. (18) analytically, we obtain an expression for n(R) curve. Higher nucleation densities lead to earlier droplet coa-
which is identical for growth of noninteracting droplets on a fiat lescence and smaller coalescence lengths (l^ = 2Re). The popu-
hydrophobic surface [29]. Note that in this case, W droplets are lation of small (R<Re) noninteracting droplets is higher than
assumed to grow with constant apparent contact angles, making large (R> Re) droplets because large droplets experience coa-
the analysis analogous to droplet condensation on a flat hydropho- lescence in addition to being swept off the surface. Model pa-
bic surface. rameters: h = 10fim, l=4nm, d=300nm, \T=Tsat - 7's = 5K,
For large coalescing droplets, the droplet size distribution N(R) 0a/0r=121.6deg/86.1deg.
was established by Glicksman and Rose [25,26]
Where
-(2/3)
1
(24)
^?-Ä\„ r/f R) R^ I ^ ^

where R is the average maximum droplet radius (departure ra- (28)


dius), and Re is the radius when droplets growing by direct vapor
addition begin to merge and grow by droplet coalescence. R can R "^^ fp P /? I (29)
"2 = — ¡ - «e - n - «min ñ
be estimated by a force balance on the droplet contact line T4 [ - «min
between gravity, Fg = (2 — 3 cos öe -\- cos^ 9s)nR^pg cos 0 / 3 ,
and surface tension, F^ = 2cr^sin0e(cos0r - cos0a), given by T = • (30)
[44,45] 1 (1 M2/?2 - 14A2ÄeÄm

1/2 Figure 4 shows the droplet size distribution n(R) as a function


6(cos 0, — cos 0a) sin 9¡.a
R= of droplet radius R. The distribution at small droplet sizes (colored
(25)
7t(2 — 3 cos 0e dt)pg COS 0 lines) (A < /?e) is approximately constant since droplets do not
interact with each other. In addition, smaller droplets have the
where 0a and 0^ are the apparent structured surface advancing and highest population density due to the renewal of the surface by the
receding contact angles [36,46], 0e is the apparent equilibrium droplet sweeping mechanism. In contrast, coalescing droplets
contact angle 0e = cos"'(O.5cos0a + O.5cos0r), and 0 is the (black line) (R > Rg) have a reduced population density because
surface inclination from the vertical, i.e., 0 = 0 deg corresponds both sweeping and coalescence acts to remove them from the sur-
to a vertical condensing surface and 0 = 90 deg corresponds to a face [28,29].
horizontal surface.
The radius when droplets begin to merge on the surface R^ is
determined by assuming the nucleation process is random (Pois- 4.2 S Droplet Size Distribution. The growth and departure
son) [22,24]. Relating Rg to the nucleation density, we obtain process of S droplets on a structured surface differs markedly
from that of W droplets. A recent study showed that when small S
and PW droplets (~ 10-100 ^m) merge on superhydrophobic
(26)
structured surfaces, they can spontaneously jump via the release
of excess surface energy independent of gravity [12]. Droplet re-
where l^ is the droplet coalescence length and Ns is the droplet moval by this mechanism is highly desirable due to the increased
nucleation density per unit area of condensing surface. It is impor- number of small droplets [14] which efficiently transfer the major-
tant to note that the assumption of a random Poisson distribution ity of the heat from the surface [8,15,16]. However, if the surface
of nucleation sites results in a droplet interaction radius that is solid fraction is too high (<p > 0.1), coalescence-induced droplet
half of the interaction radius if we assume a square array of nucle- jumping is not observed due to higher droplet surface adhesion. It
ation sites [29]. is important to note that the chosen cutoff (cpi^O.l) for droplet
jumping is a representative estimate and may not be exact. In
The droplet size distribution (Eq. (18)) can be analytically
addition, the role of apparent contact angle 0 on droplet jumping
solved for by assuming the noninteracting and large droplet size
is currently not well understood, and is not considered in this
distributions are equal {n{R¿) = N{RJ). The solution is given by
model. In reality, a droplet jumping stable surface may exist
[29]
where jumping is not possible for low apparent contact angles
-(2/3) (IJl -^ 2), but may exist for larger apparent contact angles
1 R{Rc - Rrr
n{R) = exp(ßi +B2) (¡Jl —> oo) on the same surface.
R — Rx 4- Aj
In order to model the droplet size distribution, a similar
(27) approach to the W droplet model is used. Suspended droplets

111004-6 / Vol. 135, NOVEMBER 2013 Transactions of the ASME


nucleating on the tips of a structured surface grow and, once
reaching a size large enough to begin interacting, merge and are
5x10''^
removed from the surface (cp < 0.1). This process makes the non-
interacting growth mechanism dominant and the population bal- 2x10^5 \
ance theory valid for the entire growth range [24]. In addition,
1x10''^
' • • \

droplet sweeping is nonexistent on a surface exhibiting 4 = 10 pm


coalescence-induced droplet departure due to the removal of small s- 5x10^"
droplets before they reach the capillary length. Applying droplet
conservation for a size range Rj-R2, we obtain 2x10^"
/„ = 20|jm

(31) 5x10"
' •

In the limit of dR approaching zero, Eq. (31) can be reduced to


0.3 0.50.7 1 2 3 4 5 67810
ff (um)
= 0 (32)
dR
(") 1.55
In this case, since S droplets undergo growth with a constant
contact angle, an analytical solution for the droplet size distribu- 1.45
tion can be obtained. Equations (19)-(23), which outline the deri-
% 1.35 \ /-»Ik
vation of the droplet growth rate G, are valid in this case for S
droplets. Applying the chain rule and integrating Eq. (32), we
obtain
r 1.25
\
1.15

1.05

Eor large coalescing droplets {R>Rg), n(R) = O due to the


coalescence-induced droplet departure. The maximum droplet 2 4 6 8 10 12 14 16 18 20
size R for coalescence-induced jumping droplets is given by ¡c, 2Re (Mtn)

Fig. 5 (a) Droplet population densities for surfaces exhibiting


(34) coalescence-induced droplet removal as a function of radius Rfor
a variety of nucleation densities Afe with constant contact angle
ns(fl) (solid lines) and variable contact angle «^(R) (dotted lines).
Higher nucleation densities result in earlier droplet coalescence
The droplet size distribution (Eq. (33)) can be analytically
and smaller coalescence lengths (/c = 2He). Inset: schematic
solved by assuming the noninteracting and interacting droplet size showing coalescence length (t). (Ö) Ratio of the dynamic surface
distributions are equal (n(R)=NiR)) at R=Re. The solution is heat flux q^ (Eq. (46)) to the static surface heat flux q4' (Eq. (45))
given by as a function of droplet coalescence length ( y and structured
surface pillar-to-pillar spacing (/). The shaded region includes the
results for the different pillar spacings (0.5/im </< 8/im). The
'R,y^^''^R{R-, AjR + Aj
static contact angle droplet model overpredicts the surface heat
ns{R) = (35)
RJ R-R flux at small departure sizes, which shows the importance of
using the dynamic contact model for predicting PW droplet per-
formance. Model parameters: /7=10/im, l=ifim, d=300nm,
The droplet size distribution (n^(R)) can be understood physi-
A r = 5 K, ejOr = 121.6 deg/86.1 deg.
cally in terms of an asymptotic solution of the classical noninter-
acting droplet size distribution n{R) (Eq. (27)). By assuming the
sweeping time approaches infinity (T K¿ CO) or the sweeping fre- a droplet population. Droplets larger than 7?« 1.2/xm grow at a
quency approaches zero (SwO); the sweeping mechanism is reduced rate due to increasing conduction thermal resistance
removed from the population balance analysis. Including this through the droplet; therefore, the droplet distribution increases
assumption for Eq. (27), the exponential term disappears and Eq. with R up to the departure length {R = R^ = 4/2). In addition, the
(35) is obtained. It is important to note, Eq. (35) is valid for any lack of a sweeping mechanism to remove small droplets allows
surface exhibiting coalescence-induced droplet departure where for the overall population density to be higher than the classical
the droplet contact angle can be approximated as constant case (Eig. 4).
throughout the growth, and is not exclusively valid for S droplets
only.
Eigure 5(a) shows the droplet size distribution, n^iR) as a func- 4.3 PW Droplet Size Distribution. In certain cases, droplets
tion of droplet radius R for a variety of nucleation densities A^^. growing during condensation cannot be characterized by a con-
The three curves (solid lines) correspond to coalescence lengths stant contact angle 9 during growth, as in the case of PW droplets.
of 4 = 10, 15, and 20 ^m. The droplet size distribution is distinct Previous studies have shown that once PW droplets reach a size
from the derived solution for noncoalescence-induced droplet comparable to the stmcture (R^l), they begin to undergo a
departing surfaces (Eig. 4). The results indicate that droplets larger growth regime where the contact angle varies while the basal area
than the coalescence length {R > R^) are nonexistent, which is remains constant [14,24,32,33,38]. To capture this behavior in our
consistent with the physical interpretation of droplet departure model, a modified droplet size distribution was derived for surfa-
due to coalescence. Eor all three cases, the distributions n^iR) ces undergoing variable contact angle droplet growth (PW).
have a minimum at a droplet radius RKÍI.2 ¡im. This result is due Partially wetting droplets nucleate randomly on a structured
to the slower droplet growth for R<l.2fim caused by the increas- surface, grow, and upon reaching a size large enough to begin
ing contribution from the droplet curvature resistance. As a conse- interacting {R=R^), merge and are removed from the surface via
quence of the slower growth, the droplet population density is coalescence-induced departure. Droplet sweeping is nonexistent
increased due to the conservation of droplets entering and leaving on this surface due to the removal of small droplets before they

Journal of Heat Transfer NOVEMBER 2013, Vol.135 / 111004-7


reach the capillary length. Applying droplet conservation for a important since the constant contact angle solution, n^iR), tends to
size range R1-R2, we obtain overestimate the distribution by as much as 100% in some cases.
Similar to n^iR), droplets larger than the coalescence length
(/? > i?e) do not existent due to coalescence-induced droplet depar-
ture. In addition, the distribution n¿(R) has a minimum identical to
where n^ is the "dynamic" droplet size distribution for droplets that of n^iR) at a droplet radius RRil.2ßm due to the curvature
undergoing variable contact angle growth (PW), and Gpw is the thermal resistance included in the growth model (see Sec. 4.2).
PW droplet growth rate Gp^ = dRpJdt. The term dynamic is used To accurately predict the overall structured surface dropwise
here due to the nonexclusiveness of this derivation to PW droplets condensation heat transfer performance, the droplet size distribu-
only. This formulation can be applied to any droplet morphologies tions derived in this section are combined with the individual
undergoing variable contact angle growth. droplet growth rates to determine the overall surface heat flux.
In the limit as dR approaches zero, Eq. (36) can be reduced to
5 Overall Surface Heat Flux
= 0 (37)
The overall surface heat flux, q", can be obtained at steady state
dR
by combining the morphology dependent individual droplet heat
Since PW droplets grow with a variable contact angle, an ana- transfer rate (Eq. (13)) with the droplet size distributions (Eqs.
lytical solution for the droplet size distribution n¿{R) cannot be (27), (35), and (41)). For flat hydrophobic surfaces or structured
obtained. In addition, Eqs. (19)-(23), which represent the droplet surfaces showing the W droplet morphology, the surface heat flux
growth rate Gp„ must be rederived due to the contact angle var- is given by
iance (Sec. 3.1). The latent heat of phase change for PW droplets
is expressed by
f q{R)N{R)dR (44)

- cos^ 0) sm9—R
For structured surfaces showing the S droplet morphology, or
+ ( 1 - c o s 0)^(2-I-cos Ö) (38) coalescence-induced droplet departure with a static droplet con-
tact angle, the surface heat flux is given by

By relating Eq. (38) to the droplet heat transfer Eq. (19), we can
solve explicitly for Gpw q{R)n,{R)dR (45)

Rn
1 - For structured surfaces showing the PW droplet morphology, or
R (39) surfaces exhibiting nonconstant contact angle behavior, the sur-
A2R+A3 face heat flux is given by
Where
(46)
AT
(1 - cos 0)^(2 + cos 0) To compare the utility of the developed dynamic contact angle
model, we determined the predicted surface heat flux ratio of the
(40)
constant contact angle with the variable contact angle formula-
tions for a surface undergoing PW droplet growth. Figure 5{b)
Because Gp» is a function of R and 0, Eq. (37) is first expanded shows the results of the comparison, indicating that the static con-
and solved numerically. The differential equation to be solved is tact angle formulation is a good approximation for variable con-
given by tact angle droplets at departure lengths larger than approximately
8 /^m. However, at smaller departure lengths, the static formula-
1
(41) tion deviates significantly from the dynamic solution, with error in
dR dR excess of 50% at coalescence lengths below 2 ßm. Although the
error is large at small length scales {R<S /(m), the static approxi-
The boundary condition for this case is identical to that of the S mation works fairly well at larger coalescence lengths, showing
case, where the droplet size distribution a.t R=Re is equal to the the error to be within 10%.
Rose distribution [25] given by In order to gain a better understanding of structured surface
design, the dropwise condensation models developed were applied
-(2/3) to an optimization scheme to examine the interplay between sur-
(42) face structure, droplet morphology, removal mechanisms and
overall surface heat transfer performance.

where
6 Desigu and Optimization of Structured Surfaces
R - ' ' - (43) A number of recent works have fabricated superhydrophobic
structured surfaces for the purpose of enhanced dropwise
condensation via coalescence-induced droplet removal
Figure 5 (a) shows the droplet size distribution, n¿(R) as a func- [10,19-21,23,31,47-50]. These surfaces were designed to be
tion of droplet radius R for a variety of nucleation densities. The Cassie stable such that PW or S droplets are formed on micro/
three plotted curves (dashed lines) correspond to coalescence nanostructures [30] to have minimal contact line pinning. How-
lengths of /c = 10, 15, and 20 ßm. The dynamic droplet size distri- ever, the design methodology of these surfaces is not very well
bution varies from the "static" or S solution (Eq. (35)). The plot understood. In an effort to provide a rational basis for structured
indicates that inclusion of the droplet contact angle variability is surface design, prior to and post droplet departure, we utilized the

111004-8 / Vol. 135, NOVEMBER 2013 Transactions of the ASME


(a) 5 ration or Ar. The results indicate distinct regions of droplet mor-
phology in accordance to the wetting criteria outlined in Sec. 2. For
d/l < 0.16, the W wetting morphology was favored due to the low
energy barrier of liquid propagation in the lateral direction. At
larger d/l ratios (0.16<á//<0.36), the Cassie (PW) morphology
became favored with coalescence-induced jumping as the main
Id I mode of droplet removal. At even larger d/l ratios (d/l > 0.36), the
PW morphology remained favorable; however, the coalescence-
induced droplet jumping behavior significantly diminished in favor
of conventional droplet sweeping due to high droplet surface pin-
ning (cp > 0.1).
Figure 6(a) show the dynamic nature of 9 for all three droplet
morphologies (W, PW jumping, and PW shedding). At IJl < 2
(not shown), 9 remains constant since the droplet radius is not
laige enough to fill the structure. In this regime, droplets grow
individually within the unit cells and take on the intrinsic advanc-
ing contact angle, 0^. Upon reaching a size comparable to 2 x 2
unit cells {ljl = 2), droplets begin to either (1) emerge from the
structure and grow in the eonstant-basal-area/variable-contact-
angle PW droplet mode («'//> 0.16) [2,32,38] or (2) spread later-
ally and grow in the constant contact angle W droplet mode
(<¿//<0.16). For the jumping region (0.16 <c///< 0.36),
9 increased with increasing IJl due to the differing droplet depar-
ture radii. The region of nonjumping PW droplets (d/l > 0.36)
showed constant 9 that decreased with increasing d/l due to
increasing solid fraction (p. The W regime (d/l < 0.16) showed
increased 6 with increasing d/l due to larger surface roughness r.
In addition to having a large influence on the droplet contact
angle, the disthict regions of different droplet morphologies cre-
ated a large variance in the droplet departure radius R (Fig. 6(b)).
Fig. 6 (a) Condensing dropiet apparent contact angie 0 as a The W regime had the highest R due to the high contact angle hys-
function of coalescence length (IJI) and ratio of piiiar diameter teresis and droplet surface adhesion. W droplets rely on the gravi-
to center-to-center spacing {d/l). Distinct regions of differing tational force to be removed from the surface, reaching sizes
droplet wetting morphologies exist based on the wetting crite- comparable to the capillary length (/f « 2 mm) before departing.
ria (Sec. 2). For £///>0.36, the PW droplet morphology is In contrast, PW droplets exhibiting jumping behavior (0.16 < d/l
favored; however, droplet jumping is not possible due to the
high solid fraction ((p>0^) and high contact iine pinning to the < 0.36) depart, from the surface at length scales well below the capü-
surface structure. For ljl<2 (not-shown), iiquid fiims and lary length (R <C 1mm). As d/l increased further (d/l > 0.36), the
pinned W droplets are formed due to droplet merging within the PW nonjumping regime becomes favored. However, the departure
unit ceil of the structure, (b) Condensing droplet departure ra- radii of these nonjumping PW droplets are well below that of W
dius f? as a function of coaiescence length {IJt} and ratio of pii- droplets due to the significantly smaller contact angle hysteresis and
lar diameter to center-to-center spacing (cUi). Regimes of W droplet surface adhesion associated with the Cassie morphology. At
droplet formation have higher departure radii than PW droplets coalescence lengths of IJl < 2 (not shown), droplets merge within
due to higher surface adhesion and contact angie hysteresis. the structure to form liquid films and highly irregular W droplets that
Model parameters: h=-\Ofim, t=4ftm, £í=300nm, A r = 5 K , depart via gravitational shedding.
0a/ör = 121.6deg/86.1deg, ^HC ~ 0.2W/mK [28], /fp = 150W/mK,
ÔHC = 1 nm. Insets: emergent droplet morphology schematics
for each region.
6.1 Optimization of Surface Heat Flux. The optimal drop-
wise condensation enhancement was determined based on the
newly developed model to evaluate the effects of surface geome- overall surface heat flux for a variety of structured surface geome-
try, scaling down the size scale of the structures, and intrinsic sur- tries. Figure 7(a) shows the normalized surface condensation heat
face wettability (promoter coating) on overall heat transfer flux, q"/qmJ', as a function structure geometry (d/l) and coales-
performance. cence length (IJl). Distinct regions of operation exist due to vary-
The droplet wetting morphology model was combined numeri- ing droplet morphologies and their associated departure and
cally with the growth rate and size distribution models. In the case contact angle characteristics (Fig. 6). As expected, the region
of Cassie droplet formation, we assumed the PW morphology to favoring PW jumping droplets (0.16 < d / / < 0.36) showed maxi-
emerge due to the use of smooth pillars and the ability of S droplets mum heat flux for the entire range of IJL Wenzel droplets
to transition to PW droplets [24]. This assumption was used to pro- (d/l < 0.16) showed the lowest heat flux due to their relatively
vide an upper bound for surface heat transfer performance when large departure radii indicating that structured surfaces having
compared to conventional flat hydrophobic surfaces. Figure 6 very low solid fractions can be far from optimal in terms of drop-
shows the droplet contact angle 9 (Fig. 6(a)) and departure radius R wise condensation performance. The regime of nonjumping PW
(Fig. 6(b)) as a function of structure geometry (d/l) and coalescence droplet formation (rf//>0.36) showed a decrease in performance
length (IJl). It is important to note, the coalescence length (or compared to the jumping regime. However, at increased solid
nucleation density, Eq. 34) is used as a variable parameter in the fractions (d/l > 0.8), the heat flux became comparable due to
model due to its dependence on the vapor to surface temperature favorable departure conditions and high droplet-base contact
difference AT. Although classical nucleation theory states that the (high individual droplet growth rate prior to departure). The
coalescence length is highly dependent on Ar, accurate prediction results indicate that an optimum structure geometry exists to max-
of nucleation density (based on AT) is difficult to achieve due to imize the overall surface heat flux, and that the unified model can
the presence of numerous high energy defect sites in the hydropho- be used to find this optimum. In this case, designing d/l to be
bic coating [4,22]. This makes model analysis based on coalescence within the range of 0.16 < d/l < 0.36 is highly favorable in terms
length desirable over extemal control parameters such as supersatu- of overall heat flux performance. However, this criteria may not

Journal of Heat Transfer NOVEMBER2013, Vol. 135 / 111004-9


(a) 5

dll
lall

Fig. 8 Structured surface steady-state wetting morphology as


a function of the pillar diameter to center-to-center spacing ra-
tio {dll) and the center-to-center spacing to pillar height ratio (//
h). Scaling down the surface structure {Ith) broadens the d/l re-
gime where PW jumping droplets are observed. Insets: emer-
gent droplet morphology schematics for each region.

h=l /im) for surfaces with smaller scale structures due to the
reduced structure thermal resistance. In addition, reduction of the
pillar height acts to extend the peak performance of the surface
for a larger range of coalescence lengths {IJl).
A second important advantage of scaling down the micro/nano-
structure is seen by the broadening of the dll ratio for the PW
jumping regime. Figure 8 shows the structured surface steady-
state wetting morphology as a function of the pillar diameter to
center-to-center spacing ratio {dll) and the center-to-center spac-
ing to pillar height ratio (///;). As / is reduced for a fixed h, the
region of dll where PW jumping droplets are favored expanded.
This increase of the dll range for PW jumping stability has very
important implications for nanostructured surface design, since
the structure can consist of highly irregular pillar type protrusions
arising from nonuniform fabrication techniques such as oxidation
[4] or self-assembly [33]. The larger range of dll at smaller length
scales facilitates tolerance of structure irregularity and imperfec-
tion, allowing for less costly, scalable, and robust fabrication tech-
niques tobe used [4].
Although the heat flux increases as the structure scale is
decreased from micro to nano, the corresponding required coales-
cence lengths are also decreased. This effect may pose a problem
Fig. 7 Normalized overall steady-state surface heat flux in realistic structured surface design since the coalescence-
cf'lcfmax, as a function of coalescence length {IJI} and ratio of induced droplet jumping mechanism has been shown to have a
pillar diameter to center-to-center spacing (dtl) for (a) h = 5fim, minimum size limit [12].
\b) h = 2tim, and (c) h=^ /im. Scaling down the surface struc-
ture ((a) to (c)) enhances performance due to the reduced
micro/nanostructure thermal resistance. Regions favoring PW 7 Intrinsic Contact Angle
jumping droplet removal show peak heat fluxes for all three
cases ((a) to (c)). qfmax" was determined from examining the Dropwise condensation of water on metal/metal oxide surfaces
peak heat flux in all three cases, which occurred for the small- is rarely observed in natural conditions due to their high surface
est scale structure (c), t/max" = 342.12 kW/m^. Model parameters: energies (the exception being noble metals such as gold and plati-
h/l=2, A r = 5 K , 0a/0r = 121.6 deg/86.1 deg, /CHC = 0.2 W/mK, num due to adsorption of atmospheric hydrocarbons [51-53]). To
/fp = 150 W/mK, áHc = 1nm. Insets: emergent droplet morphol- promote dropwise condensation, the use of SAMs has emerged as
ogy schematics for each region. popular and robust technique to obtain reduced surface energies
[4,10,12,14,19,22-24,29,32,38,47,54,55]. The use of SAMs as
hold for different surfaces, since the morphology is strongly de- surface promoter coatings has three advantages: (1) The SAM
pendent on the solid fraction and surface roughness. thickness is on the order of a few nanometers, resulting in minimal
resistance to heat flow [29]; (2) a range of SAMs can form a cova-
lent bond with metal oxide surfaces greatly increasing robustness;
6.2 Effect of Micro/Nanostructure Scale. Scaling down of and (3) the highly nonwetting properties of fluorinated SAMs act
the structured surface has been shown to potentially increase the as a suitable promoter for dropwise condensation. The developed
overall performance due to the reduction in thermal resistance model is utilized to quantify the effect of SAM coating properties
between the base of the emerging droplet and the structured sur- for selecting an optimal dropwise condensation promoter.
face substrate [2,24]. To study the effects of structure scale on the Four chemistries for SAM formation were examined, with
overall surface heat transfer performance, the unified model was varying chain lengths and intrinsic contact angles on smooth
used to simulate condensation on a surface having h/l = 2 for a va- surfaces. The advancing and receding contact angles used
riety of different pillar heights h. Figures l{a)-l{c) show the nor- were: 0^/0^= 121.6 deg/86.1 deg, 0^/0^= 110.8 deg/85.8 deg, and
malized surface heat flux, q"lq"m¡,,,, for pillar heights oí h = 5 /im, 0a/0r = 103.8 deg/102.7 deg corresponding to typical wetting prop-
h = 2ßm, and h=\ /tm, respectively. The results show enhanced erties for deposited films of (tridecafluoro-l,l,2,2-tetrahydrooc-
heat transfer performance (up to 22.5% from h = 5ßm to tyl)-l-trichlorosilane (SAMl), octadecyltrichlorosilane (SAM2),

111004-10 / Vol. 135, NOVEMBER 2013 Transactions of the ASME


100

80

^E 60

20
SAM1
swa
THHDL
0
SAM1;a/6l=121.6786.
SAM2;Éya = 110.e785.8 0 10 20 30 40 50 60 70 80 90
THOL; 9 J f l = 121.17106.3

Fig. 10 Structured surface steady-state heat flux qr" as a func-


tion of the surface inclination angie 0 for SAM coated (SAM1,
SAM2, and THIOL) structured surfaces exhibiting coaiescence-
induced dropiet jumping (d//=0.3) and gravity based dropiet
shedding (i*/=0.4). Jumping surfaces showed little sensitivity
to the surface orientation owing to their ability to shed droplets
at length scales well below the capillary length (R <c 1 mm).
Surfaces exhibiting gravity based shedding showed a strong
dependence on 0, due to the cos0 dependence of the gravita-
tionai body force acting on the condensing droplets needed to
overcome the surface tension force (Eq. (25)). Inset: condens-
ing surface orientation schematic. Model parameters: h = 5 fim,
/=2.5/im, A r = 5 K , /c = 10/im,fcHc= 0.2W/mK, /fp = 150W/mK,
and (5HC = 1 nm.
SAM1; «,/«(= 121.6786.1"
SAM2:9>/íl=110.S786.8"
SAM3; (?a/'^= 103.87102.7
THOL; to a size large enough to overcome the adhesive surface tension
force. The adhesion force is highly dependent on contact angle
hysteresis {A9 = 9¡,-9,) of the coating (Eq. (25)); the larger the
hysteresis, the larger the adhesion to the surface [44,45]. The
lower the hysteresis, the smaller the pinning force and average
Fig. 9 Structured surface steady-state heat flux qr" as a func- droplet size before departure occurs, resulting in a larger popula-
tion of coalescence length 1^ for four different promoter coat-
ings with (a) coalescence-induced droplet jumping (no tion of smaller droplets and enhanced q". Figure 9(b) shows that
sweeping) and (b) gravitational dropiet removai (sweeping). the smallest contact angle hysteresis coatings have the best per^
The surface heat flux is not sensitive to the promoter coating formance (SAM3 has the best performance (AÖSAM3 =1-1 deg)
for surfaces with coalescence-induced droplet departure. Heat followed in order by THIOL (A0THIOL= 14.8 deg), SAM2
flux (qr") is highly dependent on the promoter coating for surfa- (AÖSAM2 = 25 deg), and SAMl (ASSAMI = 35.5 deg)).
ces relying on gravity for droplet removal due to the strong de-
pendence of droplet/surface adhesion on contact angle
hysteresis. Insets: Surface heat flux (qr") as a function of tem-
perature difference {AT= Tsat - Ti) for the four different pro- 8 Condensing Surface Orientation
moter coatings and model parameters: ft = 5/im, l=2.5fim, Structured surfaces undergoing coalescence-induced droplet
A r = 5 K , /c = 7.5/im, lfHc = 0.2W/mK, /fp = 150W/mK, and jumping can be utilized in a variety of applications where conven-
¿Hc = 1 nm. tional flat hydrophobic surfaces (gravity based droplet shedding)
cannot, including space, mobile electronics, and thennal diodes
[18]. To study the effect of surface orientation on overall surface
and dichlorodimethylsilane (SAM3), respectively [22]. In addi- heat flux, the surface inclination angle, 0 , was varied from
tion, a thiolated SAM coating (THIOL) was analyzed with intrin- 0 = Odeg (vertical) to 0 = 90deg (horizontal) in the model.
sic advancing and receding contact angles of 9J9j= 121.1 deg/ Figure 10 shows the surface heat flux as a function of the inclina-
106.3 deg. tion angle for SAM coated structured surfaces exhibiting PW
Figures 9(a) and 9{b) show the surface heat flux q" as a function coalescence-induced droplet jumping {d/l = 0.3) and PW gravity
of coalescence length l¡. for a structured surface coated with the based droplet shedding (d/l = 0.4). Surfaces exhibiting jumping
four coatings with (a) d/l = 0.3 and (b) d/l = 0.4. While both surfa- showed little sensitivity to the surface orientation owing to their
ces favor formation of PW droplets, Figs. 9(a) and 9{b) underwent ability to shed droplets at length scales well below the capillary
coalescence-induced droplet jumping and gravity based droplet length (R < 1 mm). Although, in the horizontal orientation
shedding, respectively. For all four coatings, as 4 decreased, q" ( 0 = 90 deg), jumping droplets can return to the surface via gravi-
increased due to the increase in population of small droplets tational force and reduce heat transfer, this effect can be neglected
(R < 4). The impact of the coatings on q" showed different sensi- due to the ease of small droplet ( ~ 10/tm) advection or entrain-
tivity depending on the droplet morphology and departure mecha- ment in the bulk water vapor flow above the surface.
nism. In the case of coalescence-induced droplet removal In contrast to jumping droplets, heat flux perfonnance of surfa-
(Fig. 9(a)), the effect of intrinsic wetting angle (coating type) ces exhibiting gravity based droplet shedding was reduced due to
on q" was small. This result is due to the fact that droplets grow- larger droplet departure size (Fig. 6(b)), and showed large sensi-
ing on the structured surface depart at length scales well below tivity to surface orientation. The main mechanism of droplet re-
the capillary length (4 < 1 mm), where contact line hysteresis is moval in this case is highly dependent on the gravitational body
very important in the removal mechanism of droplets from the force acting on the condensing droplets. Before they can be
surface. removed, droplets need to grow to a size large enough for gravity
When the droplet removal mechanism depends on gravity to overcome the contact line surface tension force holding them
(Fig. 9(è)), the intrinsic wetting angle of the coating has a signifi- on to the surface. The surface tension force arises due to contact
cant effect on q". Droplets being removed by gravity need to grow angle hysteresis and acts in a direction parallel with the surface

Journal of Heat Transfer NOVEMBER2013, Vol. 135 / 111004-11


Substrate. Therefore, to maximize the gravitation force, the vertical denser sizes and cost [51]; (2) overall performance enhancement of
surface orientation is highly favored and any deviation toward the devices such as heat pipes and thermal ground planes for applica-
horizontal orientation results in a larger departing droplet size (R) tions requiring maximization of evaporator area and minimization
and lower overall surface heat flux (particularly at inclination of condenser area [56]; and (3) use of cooling devices previously
angles above 30 deg). As expected, the SAM coatings with the not possible for local high heat flux electronic devices [18,57].
smallest contact angle hysteresis have the best performance (SAMl Although the added benefit of surface structuring may outweigh
has the worst performance (AÖSAMI = 35.5 deg) followed in order its increased complexity, some practical limitations remain. The
by SAM2 (AÖSAM2 = 25 deg), and THIOL (AÔTHIOL =14.8 deg)). scalability of the surface structuring process is a concem that has
to be addressed before practical designs are implemented in indus-
trial settings. Recent research into novel and scalable copper oxi-
9 Flat Versus Structured Surfaces dation nanostructuring techniques has alleviated some of this
In an effort to address the question: can a flat hydrophobic sur- concem [4,58,59]; however, the robustness and durability after
face with low contact angle hysteresis, AÖ = oa — 0r. outperform a long hours of operation is unknown given the relatively harsh
structured superhydrophobic surface exhibiting droplet jumping, working environments encountered in industry.
we used the developed model to compare the heat transfer of struc-
tured surfaces exhibiting a range of length scales (lnm<h
10 Conclusions
<5|im, h/l = 2) with flat hydrophobic surfaces with a range of A model framework of dropwise condensation heat transfer for
intrinsic surface contact angle hysteresis values micro/nanostructured superhydrophobic surfaces was developed.
(Odeg < AÖ < 55 deg) (Fig. (11)). Eigure 11 shows that as the con- Unlike previous works, the current model is able to (1) predict the
tact angle hysteresis is reduced for a flat surface (öa=120deg, emergent droplet wetting morphology (PW jumping, PW non-
of = variable), the overall heat flux is increased due to the reduction jumping, or Wenzel) via coupling of the structure geometry and
in the required size of the droplet to overcome surface tension nucleation density by considering local energy barriers to wetting;
forces to initiate shedding (Fig. 11 inset). In addition, the results (2) model nonconstant contact angle droplet growth; and (3)
show that at low enough contact angle hysteresis, the flat surface extend the previously developed droplet size distribution theory to
can potentially have better performance than a jumping superhy- both constant and nonconstant contact angle droplets growing on
drophobic surface. However, as the length scale of the superhydro- surfaces experiencing coalescence-induced droplet jumping. The
phobic surface is reduced, the required hysteresis to maintain model was used to compute the overall surface heat transfer rate
enhancement is also reduced (A0<27deg for h = 5iim, and study the effects of surface geometry and scale, nucleation
A9 < 2.5 deg for A = 2 fan, AÖ < 1 deg for /¡ = 1 pm). This result is density, and promoter coating. The results suggest the importance
of emergent droplet wetting morphology on overall condensation
in accordance with the theory and subsequently the jumping sur-
heat flux. Specifically, distinct geometries existed which allowed
face heat fluxes are increased (Sec. 6.2). Therefore, structured
for the formation of coalescence-induced jumping droplets having
surfaces with relatively large structure scale features may not be
up to 110% overall surface heat flux enhancement over dropwise
advantageous when compared to a flat hydrophobic surface. How-
condensing geometries favoring W droplets, and 190% enhance-
ever, provided that the surface structure design is optimized ment over flat hydrophobic surfaces. In addition, the results
(Fig. 7(c)) with low coalescence lengths (2 < IJl < 4), the heat flux showed that scaling down the stmcture dimensions, while main-
performance enhancement of peak superhydrophobic condensation taining droplet coalescence at small length scales helps to sustain
(<7max = 342.12kW/m^ at A7'=5K) compared to that of an ideal coalescence-induced droplet jumping over a larger range of geo-
flat surface (A0 < 2.5 deg) approaches 110%. In addition, using the metries and maximize heat transfer performance enhancement.
model results to compare the peak superhydrophobic heat flux
(<?max") to a more realistic flat dropwise condensing surface (SAMl Subsequently, the model was used to study the effects of four
coated) shows an even greater enhancement of up to 190%. self-assembled monolayer promoter coatings on overall heat trans-
The results indicate that structured surfaces may be the ideal fer, showing that surfaces exhibiting coalescence-induced droplet
pathway to achieving even higher dropwise condensation heat jumping are not sensitive to the intrinsic promoter coating wetting
fluxes over conventional flat hydrophobic surfaces. Such surfaces characteristics, which is in contrast to surfaces relying on gravity.
may therefore enable (1) substantial reduction in industrial con- Similarly, the impact of surface inclination with respect to gravity
does not change the heat transfer characteristics of surfaces exhibit-
ing jumping droplets, which is not the case for gravity based shed-
178
2.5 ding surfaces, particularly at inclination angles in excess of 30 deg.
166 The results indicate that stmctured surfaces may be the ideal path-
154 way to achieving high heat flux dropwise condensation. This work
142 provides guidelines for the rational design of stmctured superhydro-
130 phobic surfaces to maximize dropwise condensation heat transfer.
/ ^
118 0 10 20 30 40 50

106
Acknowledgment
94 This work was supported as part of the MIT S3TEC Center, an
82 Fiai Surface
Energy Erontier Research Center funded by the U.S. Department of
70
Jumping Surfece Energy, Office of Science, Basic Energy Sciences under Award #
0 6 12 18 24 30 36 42 48 54 DE-FG02-09ER46577. R.E. acknowledges support from the Irish
Research Council for Science, Engineering, and Technology,
cofunded by Marie Curie Actions under EP7. This work was per-
Fig. 11 Structured (jumping) and flat (gravity shedding) surface formed in part at the Center for Nanoscale Systems (CNS), a mem-
steady-state heat flux </' as a function of intrinsic promoter coat- ber of the National Nanotechnology Infrastmcture Network (NNIN),
ing contact angle hysteresis A0 for three structured surfaces which is supported by the National Science Foundation under NSF
coated with the SAM1 promoter. As Ad decreases for the flat sur-
award number ECS-0335765. CNS is part of Harvard University.
face, cf' increases due to the lower droplet adhesion to the
surface and lower departure radii (inset). As a result, the flat
surfaces begin to show enhanced qf" compared to the structured Nomenclature
surfaces. Inset: Droplet departure diameter {R), as a function of
intrinsic flat surface contact angle hysteresis (Ad). Model param- A = area of condensing surface (m^)
eters: A r = 5 K , /c = 10/<m, /fHc = 0.2 W/mK, kp = 150 W/mK, d = pillar diameter (m)
áHc = 1 nm, and SAMl coating: 0a/ör = 121.6deg/86.1 deg. E* = wetting-state energy ratio

111004-12 / Vol. 135, NOVEMBER 2013 Transactions of the ASME


gravitational force acting on a droplet (N) 9s = suspended droplet contact angle (deg)
surface tension force acting on a droplet (N) öpw = partially wetting droplet contact angle (deg)
g gravitational acceleration, (9.81 N/kg) öa = advancing contact angle (deg)
G droplet growth rate (m/s) of = receding contact angle (deg)
partially wetting (variable contact angle) droplet growth AÖ = intrinsic contact angle hysteresis (deg)
rate (m/s) 0 = surface inclination angle fïom the vertical (deg)
h = pillar height (m) cp = solid fraction
latent heat of vaporization (J/kg) a = surface tension (N/m)
liquid-vapor interface heat transfer coefficient (W/m K) PH. = water density (kg/m"*)
hydrophobic promoter coating thermal conductivity (W/ Vg = water vapor specific volume (m^/kg)
mK) ô = thickness (m)
pillar/substrate thermal conductivity (W/m K) T = sweeping period (s)
water vapor thermal conductivity (W/m K)
water thermal conductivity (W/m K) Subscripts
/ = pillar center-to-center spacing (m)
4, = droplet coalescence length (m) a= advancing
(¡c) = mean droplet coalescence length (m) c= curvature, critical, coalescence
m = mass rate of condensate formation (kg/s) CB = Cassie Baxter
N = large droplet population density (m" ) d= droplet
n = small droplet population density (m^') e= equilibrium, effective
constant contact angle small droplet population density F= flat
g= pinned liquid region under droplet
«(J = variable contact angle small droplet population density HC = hydrophobic promoter coating
(m~') i= liquid-vapor interface
Ns = number of nucleation sites per unit area (m ^) p= pillar
P = vapor saturation pressure (Pa) PW = partially wetting
q = individual droplet heat transfer (W) P-C-G = pillar, coating, pinned liquid region
qp-w = partially wetting morphology drop heat transfer (W) , r= receding
qs = suspended morphology drop heat transfer (W) S= suspended
«iw = Wenzel morphology drop heat transfer (W) w= water
q" = heat flux (W/m^) W= Wenzel
9rnax = maximum structured surface heat flux (W/m^)
r = surface roughness Superscripts
R = droplet radius (m)
Rmin = minimum droplet nucleation radius (m) CB = Cassie-Baxter
^max = droplet radius when variable contact angle growth ends ESEM= environmental scanning electron microscopy
(PW droplets) (m) PW = partially wetting
R = effective maximum droplet radius (m) S= suspended
/?e = droplet interaction radius (m) W= Wenzel
Re = droplet curvature thermal resistance (K/W) SAM = self-assembled monolayer
R^a = critical radius when R^ = R, (m)
Ri = liquid-vapor interfacial thermal resistance (K/W)
R¿ = droplet conduction thermal resistance (K/W)
References
[1] Schmidt, E., Schurig. W.. and Sellschopp, W.. 1930, "Versuche über die Kon-
Rhc = hydrophobic promoter coating thermal resistance (K/W) densation von Wasserdampf in Film- und Tropfenform," Forsch. Ingenieurwes.,
Rp = pillar structure thermal resistance (K/W) 1(2), pp. 53-63.
Rg = pinned liquid region thermal resistance (K/W) [2] Miljkovic. N., Enright, R., and Wang, E. N., 2012, "Growth Dynamics During
R = gas constant (J/mol K) Dropwise Condensation on Nanostructured Superhydrophobic Surfaces." 3rd
5 = droplet surface area removal rate due to large droplet Micro/Nanoscale Heat and Mass Transfer Intemational Conference, Atlanta,
GA.
sweeping (m^/s) [3] Ma, X. H., Rose, J. W.. Xu, D. Q., Lin, J. F., and Wang, B. X., 2000. "Advances
t = time (s) in Dropwise Condensation Heat Transfer; Chinese Research." Chem. Eng. J.,
78(2-3), pp. 87-93.
Ar = surface subcooling temperature (K) [4] Enright, R., Dou. N., Miljkovic, N., Nam, Y., and Wang, E. N., 2012,
ATc = droplet curvature temperature drop (K] "Condensation on Superhydrophobic Copper Oxide Nanostructures," ASME J.
ATiic = coating layer conduction temperature drop (K) Heat Transfer, 135(9), p. 091012.
ATi = liquid-vapor interfacial temperature drop (K) [5] Le Fevre, E. J., and Rose, J. W., 1964, "Heat-Transfer Measurements During
Dropwise Condensation of Steam," Int. J. Heat Mass Transfer, 7, pp. 272-273.
AT¿ = droplet body conduction temperature drop (K) [6] Welch, J., and Westwater, J. W.. 1961, "Microscopic Study of Dropwise Con-
A7p s = suspended droplet pillar temperature drop (K) densation," Proceedings of the Second Intemational Heat Transfer Conference,
ATp2 = pillar, coating and gap temperature drop (K) Vol. 2, pp. 302-309.
Thi = liquid temperature at the droplet base (K) [7] Ma, X. H., Wang, S. F., Lan, Z., Peng, B. L., Ma, H. B., and Cheng, P., 2012,
7'b2 = temperature of pillar tops beneath coating (K) "Wetting Mode Evolution of Steam Dropwise Condensation on Superhydro-
phobic Surface in the Presence of Noncondensable Gas," ASME J. Heat Trans-
Ti = liquid-vapor interfacial temperature (K) fer, 134(2), p. 021501.
Tsat = vapor saturation temperature (K) [8] Glicksman, L. R., and Hunt, A. W., 1972, "Numerical Simulation of Dropwise
Tg = substrate/wall temperature (K) Condensation," Int. J. Heat Mass Transfer, 15(11), pp. 2251-2269.
[9] Love, J. C . Estroff, L. A.. Kriebel, J. K., Nuzzo, R. G., and Whitesides, G. M.,
Ts = substrate subcooled region temperature (K) 2005, "Self-Assembled Monolayers of Thiolates on Metals as a Form of Nano-
V = volume (m^) technology," Chem. Rev., 105(4), pp. 1103-1169.
[10] Andrews, H. G., Eccles, E. A., Schofield, W. C. E., and Badyal, J. P. S., 2011.
"Three-Dimensional Hierarchical Structures for Fog Harvesting," Langmuir,
Greek Symbols 27(7), pp. 3798-3802.
a = condensation coefficient [11] Leach, R. N., Stevens, F., Langford, S. C, and Dickinson, J. T., 2006,
"Dropwise Condensation: Experiments and Simulations of Nucleation and
0 = contact angle (deg) Growth of Water Drops in a Cooling System," Langmuir, 22(21), pp.
0w = Wenzel droplet contact angle (deg) 8864-8872.

Journal of Heat Transfer NOVEMBER2013, Vol. 135 / 111004-13


[12] Boreyko, J. B.. and Chen, C. H., 2009, "Self-Propelled Dropwise Condensate [36] Choi, W., Tuteja, A., Mabry, J. M., Cohen, R. E., and McKinley, G. H., 2009,
on Superhydrophobic Surfaces," Phys. Rev. Lett., 103(18), p. 184501. "A Modified Cassie-Baxter Relationship to Explain Contact Angle Hysteresis
[13] Lafuma, A., and Quere, D., 2003, "Superhydrophobic States," Nature Mater., and Anisotropy on Non-Wetting Textured Surfaces," J. Colloid Interface Sei.,
2(7), pp. 457-460. 339(1), pp. 208-216.
[14] Dietz, C., Rykaczewski, K., Fedorov, A. G., and Joshi, Y., 2010, "Visualization [37] Anand, S., and Son, S. Y., 2010, "Sub-Micrometer Dropwise Condensation
of Droplet Departure on a Superhydrophobic Surface and Implications to Heat Under Superheated and Rarefied Vapor Condition," Langmuir, 26(22), pp.
Transfer Enhancement During Dropwise Condensation," Appl. Phys. Lett., 17100-17110.
97(3), p. 033104. [38] Rykaczewski, K., Scott, J. H. J., and Fedorov, A. G., 2011, "Electron Beam
[15] Graham, C, and Griffith, P., 1973, "Drop Size Distributions and Heat-Transfer Heating Effects During Environmental Scanning Electron Microscopy Imaging
in Dropwise Condensation." Int. J. Heat Mass Transfer, 16(2), pp. 337-346. of Water Condensation on Superhydrophobic Surfaces," Appl. Phys. Lett.,
[16] Rose, J. W., 1967, "On the Mechanism of Dropwise Condensation," Int. J. Heat 98(9), p. 093106.
Mass Transfer, 10. pp. 755-762. [39] Kaschiev. D., 2000, Nucleation: Basic Theory With Applications, Butterworth
[17] Anderson, D. M., Gupta, M. K., Voevodin, A. A., Hunter, C. N., Putnam, S. A., Heinemann, Oxford, UK.
Tsukruk, V. V., and Fedorov, A. G., 2012, "Using Amphiphilic Nanostructures [40] Tam, D., von Arnim, V., McKinley, G. H., and Hosoi, A. E., 2009, "Marangoni
To Enable Long-Range Ensemble Coalescence and Surface Rejuvenation in Convection in Droplets on Superhydrophobic Surfaces," J. Fluid Mech., 624,
Dropwise Condensation," ACS Nano, 6(4), pp. 3262-3268. pp. 101-123.
[18] Boreyko, 1. B., Zhao, Y. J., and Chen, C. H., 2011, "Planar Jumping-Drop Ther- [41] Carey, V. P., 2008, Liquid-Vapor Phase-Change Phenomena: An Introduction
mal Diodes," Appl. Phys. Lett., 99(23), p. 234105. to the Thermophysics of Vaporization and Condensation Processes in Heat
[19] Dietz, C , Rykaczewski, K., Fedorov, A., and Joshi, Y., 2010, "ESEM Imaging Transfer Equipment, Taylor and Francis, New York.
of Condensation on a Nanostructured Superhydrophobic Surface," ASME J. [42] Schräge, R. W., 1953, A Theoretical Study of Interphase Mass Transfer, Colum-
Heat Transfer, 132(8), p. 080904. bia University Press, New York.
[20] Narhe, R. D., and Beysens, D. A., 2006, "Water Condensation on a Super- [43] Umur, A., and Griffith, P., 1965, "Mechanism of Dropwise Condensation,"
Hydrophobic Spike Surface," Europhys. Lett., 7S(1), pp. 98-104. ASME J. Heat Transfer, 87(2), pp. 275-282.
[21] Varanasi, K. K., Hsu, M., Bhate, N., Yang, W. S.. and Deng, T., 2009, "Spatial [44] Dimitrakopoulos, P., and Higdon, J. J. L., 1999, "On The Gravitational Dis-
Control in the Heterogeneous Nucleation of Water," Appl. Phys. Lett., 95(9), placement of Three-Dimensional Ruid Droplets From Inclined Solid Surfaces,"
pp.094101. J. Fluid Mech., 395, pp. 181-209.
[22] Enright, R., Miljkovic, N., Al-Obeidi, A., Thompson, C. V., and Wang, E. N., [45] Kim, H. Y., Lee, H. J., and Kang, B. H., 2002, "Sliding of Liquid Drops Down
2012, "Condensation on Superhydrophobic Surfaces: The Role of Local an Inclined Solid Surface," J. Colloid Interface Sei, 247(2), pp. 372-380.
Energy Barriers and Structure Length Scale," Langmuir, 28(40), pp. [46] Raj, R., Enright, R., Zhu, Y., Adera, S., and Wang, E. N., 2012, "Unified Model
14424-14432. for Contact Angle Hysteresis on Heterogeneous and Superhydrophobic
[23] Miljkovic, N., Enright, R., Maroo, S. C , Cho, H. J., and Wang, E. N., 2011, Surfaces," Langmuir, 28(45), pp. 15777-15788.
"Liquid Evaporation on Superhydrophobic and Superhydrophilic Nanostruc- [47] Chen, C. H., Cai, Q. J., Tsai, C. L., Chen, C. L., Xiong, G. Y., Yu, Y., and Ren,
tured Surfaces," J. Heat Transfer, 133(8), p. 080903. Z. F., 2007, "Dropwise Condensation on Superhydrophobic Surfaces With
[24] Miljkovic, N., Enright, R., and Wang, E. N., 2012, "Effect of Droplet Two-Tier Roughness," Appl. Phys. Lett., 90(17), p. 173108.
Morphology on Growth Dynamics and Heat Transfer During Condensation [48] Lau, K. K. S., Bico, J., Teo, K. B. K., Chhowalla, M., Amaratunga, G. A. J.,
on Superhydrophobic Nanostructured Surfaces," ACS Nano, 6(2), pp. Milne, W. I., McKinley, G. H., and Gleason, K. K., 2003, "Superhydrophobic
1776-1785. Carbon Nanotube Forests," Nano Lett, 3(12), pp. 1701-1705.
[25] Le Fevre, E. J., and Rose, J. W., 1966, "A Theory of Heat Transfer by Dropwise [49] Rykaczewski, K., 2012, "Microdroplet Growth Mechanism During Water Con-
Condensation," Proceedings of the Third Intemational Heat Transfer Confer- densation on Superhydrophobic Surfaces," Langmuir, 28(20), pp. 7720-7729.
ence, Vol. 2, pp. 362-375. [50] Rykaczewski, K., Osbom, W. A., Chinn, J., Walker, M. L., Scott, J. H. J., Jones,
[26] Rose, J. W., and Glicksman, L. R., 1973, "Dropwise Condensation—The Distri- W., Hao, C, Yao, S., and Wang, Z., 2012, "How Nanorough is Rough Enough
bution of Drop Sizes," Int. J. Heat Mass Transfer, 16, pp. 411^25. to Make a Surface Superhydrophobic During Water Condensation?," Soft Mat-
[27] Tanaka, H., and Tsuruta, T., 1984, "A Microscopic Study of Dropwise Con- ter, 8, pp. 8786-8794.
densation," Int. J. Heat Mass Transfer, 27(3), pp. 327-335. [51] Rose, J. W., 2002, "Dropwise Condensation Theory and Experiment: A
[28] AbuOrabi, M., 1998, "Modeling of Heat Transfer in Dropwise Condensation," Review," Proc. Inst. Mech. Eng., A, 216(A2), pp. 115-128.
Int. J. Heat Mass Transfer, 41(1), pp. 81-87. [52] Woodruff. D. W., and Westwater, J. W., 1981, "Steam Condensation on Vari-
[29] Kim, S., and Kim, K. J., 2011, "Dropwise Condensation Modeling Suitable for ous Gold Surfaces," ASME J. Heat Transfer, 103(4), pp. 685-692.
Superhydrophobic Surfaces," ASME J. Heat Transfer, 133(8), p. 081502. [53] Wilkins, D. G., Bromley, L. A., and Read, S. M., 1973, "Dropwise and Film-
[30] Cassie, A. B. D., and Baxter, S., 1944, "Wettability of Porous Surfaces," Trans. wise Condensation of Water Vapor on Gold," AlChE J., 19(1), pp. 119-123.
Faraday Soc, 40, pp. 546-551. [54] Chen, X., Wu, J., Ma, R., Hua, M., Koratkar, N., Yao, S., and Wang, Z., 2011,
[31] Wenzel, R. N., 1936, "Resistance of Solid Surfaces to Wetting by Water," Ind. "Nanograssed Micropyramidal Architectures for Continuous Dropwise Con-
Eng. ehem., 28, pp. 988-994. densation," Adv. Funct. Mater., 21, pp. 4617^623.
[32] Rykaezewski, K., and Scott, J. H. J., 2011, "Methodology for Imaging Nano-to- [55] Miljkovic, N., Enright, R., and Wang, E. N., 2012, "Liquid Freezing Dynamics
Microscale Water Condensation Dynamics on Complex Nanostrucmres," ACS on Hydrophobic and Superhydrophobic Surfaces," ASME J. Heat Transfer,
Nano, 5(7), pp. 5962-5968. 134(8), p. 080902.
[33] Rykaczewski, K., Scott, J. H. J., Rajauria, S., Chinn, J., Chinn, A. M., and [56] Miljkovic, N., and Wang, E. N., 2011, "Modeling and Optimization of Hybrid
Jones, W., 2011, "Three Dimensional Aspects of Droplet Coalescence During Solar Thermoelectric Systems With Thermosyphons," Sol. Energy, 85(11), pp.
Dropwise Condensation on Superhydrophobic Surfaces," Soft Matter, 7(19), 2843-2855.
pp. 8749-8752. [57] McCarthy, M., Peters, T., Allison, J., Espinosa, A., Jenicek, D,, Kariya, A,,
[34] Das, A. K., Kilty, H. P., Marto, P. J., Andeen, G. B., and Kumar, A., 2000, Koveal, C, Brisson, J, G., Lang, J. H., and Wang, E. N., 2010, "Design and
"The Use of an Organic Self-Assembled Monolayer Coating to Promote Drop- Analysis of High-Performance Air-Cooled Heat Exchanger With an Integrated
wise Condensation of Steam on Horizontal Tubes," ASME J. Heat Transfer, Capillary-Pumped Loop Heat Pipe," Intersoc C Thermal T.
122(2), pp. 278-286. [58] Miljkovic, N., Enright, R., Nam, Y., Lopez, K., Dou, N., Sack, J., and Wang, E.
[35] Vemuri, S., Kim, K. J., Wood, B. D., Govindaraju, S., and Bell, T. W., 2006, N., 2013, "Jumping-Droplet-Enhanced Condensation on Scalable Superhydro-
"Long Term Testing for Dropwise Condensation Using Self-Assembled Mono- phobic Nanostructured Surfaces," Nano Letters, 13(1), pp. 179-187.
layer Coatings of n-Octadecyl Mercaptan," Appl. Therm. Eng., 26(4), pp. [59] Miljkovic, N., and Wang, E. N., 2013, "Condensation Heat Transfer on Super-
421-429. hydrophobic Surfaces," MRS Bulletin, 38(5), pp. 397^06.

111004-14 / Vol. 135, NOVEMBER 2013 Transactions of the ASME


Copyright of Journal of Heat Transfer is the property of American Society of Mechanical
Engineers and its content may not be copied or emailed to multiple sites or posted to a listserv
without the copyright holder's express written permission. However, users may print,
download, or email articles for individual use.

You might also like