Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

International Journal of Heat and Mass Transfer 181 (2021) 121881

Contents lists available at ScienceDirect

International Journal of Heat and Mass Transfer


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

Three-phase solidification of a liquid compound droplet on a curved


surface
Truong V. Vu
Faculty of Vehicle and Energy Engineering, Phenikaa University, Hanoi, Vietnam
Phenikaa Research and Technology Institute (PRATI), A&A Green Phoenix Group, 167 Hoang Ngan, Hanoi, Vietnam

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

Article history: A compound droplet solidifying on a cold curved surface in a gaseous environment is numerically studied
Received 6 May 2021 by an axisymmetric three-phase front tracking method. The compound droplet containing an inner core
Revised 2 August 2021
of gas is initially assumed to be part of a sphere, and the curved surface is maintained at temperature
Accepted 20 August 2021
below the solidification point of the shell liquid. The effects of the radius of the cold surface, volume
change (in terms of the solid-to-liquid density ratio ρ sl ), supercooling degree (in terms of the Stefan
Keywords: number St), growth angle φ gr and shell liquid thickness on the solidified droplet shape, the solidification
Compound droplet rate and the solidification times are under consideration. Volume expansion (ρ sl < 1.0) induces an apex,
Solidification and volume shrinkage (ρ sl > 1.0) results in a cavity on top of the outer surface while volume change has a
Numerical simulation
minor effect on the inner surface of the solidified droplet. A more convex surface causes the solidification
Curved surface
process to finish earlier, whereas a more concave surface prolongs the entire solidification time. However,
the mean solidification rate, the tip angle at the droplet top and the additional height are not affected by
the radius of the curved surface. The tip angle is also slightly affected by the outer shape of the initial
droplet, the shell thickness and St while these parameters strongly affect the rate of solidification. It is
found that the completion of solidification around the inner core is not affected by the growth angle,
while it takes longer when increasing ρ sl , the size of the inner core or decreasing the Stefan number.
© 2021 Elsevier Ltd. All rights reserved.

1. Introduction propagated from the cold surface to the top of the droplet, and the
frozen droplet exhibited an apex at the droplet top. Marin et al.
Compound droplets with sizes from a few micrometers to a [15] experimentally found that the cone angle formed at the top
few millimeters can be found in many industrial applications and of such a frozen water droplet was independent of the supercool-
processes [1,2]. Some typical applications are light structures [3], ing degree of the surface or the initial water droplet shape. Some
medication technology [4] or semiconductor processing [5]. In such other recent experimental investigations of a water droplet on a
applications, a compound droplet can deform, break up and even cold surface can be found in [16,17]. A key to the formation of a
change in phase, e.g., from liquid to solid [5–7]. Therefore, there is conical top of the frozen water droplet is water denser than ice
no doubt that knowing the phase-change process of a compound [18]. Theoretical models for the frozen water droplet on a cold sur-
droplet becomes a crucial role in both academia and industry. In face were developed by [18,19] in which volume expansion upon
this study, we consider only compound droplets that have an outer freezing and a contact angle was included. However, these theo-
interface that completely encloses one or more inner droplets (or retical models assumed that the solid-liquid interface was flat. In
core), i.e., type-A (or B-) compound droplets [8–10]. Another type contrast, Tembely and Dolatabadi’s theoretical model [20] did not
of compound droplets called a Janus droplet (or partially engulfed limit the freezing interface and more correctly reflected the phys-
compound droplet) is not considered in this study [11–13]. ical interface near the three-phase line as observed in the experi-
A simple droplet may be considered as a special compound ments [15]. Numerical simulation studies with attention to volume
droplet that contains no core or no inner droplet. The liquid-to- change were also conducted for a simple droplet solidifying on a
solid phase-change process of a simple droplet has been well stud- cold plate. For instance, a finite element method was used by Vi-
ied. For example, Enríquez et al. [14] described the freezing pro- rozub et al. [21]. Shetabivash and co-workers [22] used a multi-
cess of a water droplet on a cold surface. The ice-water interface ple level set method in which one level-set was for the gas-liquid
interface and the other one for the freezing interface. A volume
of fluid method implemented in the Fluent software was used by
E-mail address: truong.vuvan@phenikaa-uni.edu.vn Zhang et al. [23]. A Lattice Boltzmann method was also devel-

https://doi.org/10.1016/j.ijheatmasstransfer.2021.121881
0017-9310/© 2021 Elsevier Ltd. All rights reserved.
T.V. Vu International Journal of Heat and Mass Transfer 181 (2021) 121881

oped for the freezing process of a droplet on a cold surface [24]. the compound droplet to solidify. This problem frequently happens.
Even though these numerical methods yielded reasonably accurate For instance, a simple droplet comes into contact with a surface,
frozen droplets with the presence of a pointy top, the effect of the a gaseous bubble is trapped in between the droplet and the sur-
growth angle at the three-phase line has not been examined in de- face, e.g., during droplet printing [37,38], which forms a compound
tail. Filling this gap, Vu and co-workers [25,26] extended the front- droplet. More directly, a compound droplet impacts on the surface
tracking method [27] to handle the three-phase issue raised in the [39]. In many applications, the contact surface is not “perfectly flat”
droplet solidification with attention to the effect of the growth an- but curved due to the target surface morphology [40,41] or succes-
gle. It was found that the tip angle on top of the solidified droplet sive deposition (i.e., impact on the other solidified droplets) [42].
decreases with an increase in the value of the growth angle. This
result conformed to the experimental observation reported in Sat- 2. Problem description and numerical method
unkin’s work [28], in which the author found that the growth an-
gle is zero for water, about 12° for silicon, 14° for germanium, 28° We consider a single-core compound droplet on a curved sur-
for indium antimonide and so on. In addition, like water, silicon face as shown in Fig. 1. The problem is assumed to be axisym-
and germanium also have their solid phase lighter than their liquid metric. The compound droplet consists of two interfaces: an inner
phase, thus a conical top surface is also presented after complete interface containing gas – inner gas core – and an outer interface
solidification [28,29]. – outer droplet. Initially, the shell of the compound droplet is fully
Obviously, numerous studies on volume change have been car- liquid and has a melting temperature denoted by Tm . The curved
ried out for a droplet solidifying on a cold plate as previously in- surface has a radius denoted by Rw and is always maintained at the
troduced. However, these studies focused solely on droplets located temperature Tcold  Tm , i.e., cold surface. The curved surface can be
on a flat surface. There are very few works considering droplets concave, as shown in Fig. 1, or convex [32]. To ease understanding
on a curved surface. Jin et al. [30] experimentally examined the of a convex or concave surface, we simply define Rw  0 if concave
freezing process of a water droplet on the cold surface of cylinders and Rw < 0 if convex. The lowest (highest) point of the concave
with a radius varying from 10 mm to 25 mm. The authors found (convex) surface is located at (0, Hw ). The compound droplet stays
that the shape of the frozen droplet was slightly affected by the on the curved surface, and its inner and outer interfaces form, re-
cylinder radius. Ju and co-workers [31] investigated a water droplet spectively, inner and outer angles θ i and θ o at the surface. To sim-
freezing on the cold surface of various spheres. Like Jin et al. [30], plify the problem, at time t = 0, we assume that the droplet has
it was found that the size of the sphere had a minor effect on the a spherical cap for the inner and outer interfaces, and the initial
ice droplet after complete freezing [31]. Unlike Ju et al. [31] and Jin volumes of the inner core and the outer droplet are respectively
et al. [30] where the cold surface was known as a convex one [32], denoted by Vi 0 and Vo 0 . Accordingly, the equivalent radii Ri and Ro
Zhang and co-workers [33] considered a cold concave surface with of the inner core and the outer droplet are defined as
a radius varying in the range of 10 – 25 mm. The authors reported
that the radius of the cold surface did not result in any remarkable Ri = [3Vi0 /(4π )]1/3 , Ro = [3Vo0 /(4π )]1/3 (1)
variation of the ice droplet. Because the temperature of the curved surface is lower than the
The above investigations focused only on simple droplets, while fusion temperature of the shell liquid of the compound droplet, the
phase-change heat transfer processes of compound droplets can shell liquid performs the phase-change process when the droplet is
be found in, e.g., [5,6]. By using jet instability, Vu and co-workers placed on the curved surface. The phase-change interface moves
[6] produced gas-in-molten metal droplets in the air. The shell from the curved surface to the droplet top. During the solidifi-
molten metal of the compound droplet solidified when the droplet cation process, three interfaces including the solid-liquid, liquid-
came into contact with a cold environment or surface. Similarly, gas and solid-gas interfaces meet at the tri-junction lines or triple
Bhagat et al. [5] formed compound droplets of germanium and points as indicated in Fig. 1. At these triple points, a growth angle
found that the compound droplet had a horn on its outer surface φ gr is defined as the angle between the tangent of the liquid-gas
after complete crystallization due to volume expansion. In addi- interface and that of the solid-gas interface (Fig. 1) [25,34]. Like
tion to volume expansion, the apex on the external surface of the our previous studies [36,43], we assume a constant growth angle
droplet is also induced by non-zero growth angles [25,28]. A recent φ gr = constant. Assuming immiscible, incompressible and Newto-
numerical simulation of the liquid-to-solid phase-change process nian fluids, the governing equations, in terms of one-fluid repre-
of a compound droplet was performed by Ho et al. [34]. The front- sentation, include
tracking method was used to track the evolution of the solid-liquid
and gas-liquid interfaces. The authors showed that the compound ∂ (ρ u )   
+ ∇ · ρ uu = −∇ p + ∇ · μ ∇ u + ∇ uT
droplet also had an apex on top of the outer surface after solidifi- ∂t 
cation because of volume expansion. However, this numerical work  
+ σ κδ x − x f n f dS + ρ f + ρ g (2)
was limited to compound droplets on a horizontal surface.
f
Clearly, the liquid-solid phase change of a droplet on a cold sur-
face has received a lot of attention. While cold flat surfaces were

well conducted, only a few researches considered a curved surface. ∂ (ρ C p T )
+ ∇ · ( ρ C p T u ) = ∇ · ( k∇ T ) + q˙ δ (x − xf )dS + ρC p H
Furthermore, to our best knowledge, no numerical work has been ∂t
carried out so far for a compound droplet on a curved surface. Ac- f
cordingly, it is necessary to bridge this gap in understanding the (3)
solidification of compound droplets because it is very important in
many applications of compound droplets, e.g., [5,6,35]. This is the
1
1 1
  
primary focus of the present study, which employs a numerical ap- ∇ ·u= − δ x − x f q˙ dS (4)
proach [36] to track the motion of the solidification interface and Lh ρs ρl
f
reveal the effects of some typical parameters on the final prod-
uct of a compound droplet on a curved surface. The investigated Here, ρ and μ are respectively the density and viscosity. Cp and
compound droplet surrounded by a gas environment consists of a k are respectively the heat capacity and thermal conductivity. The
shell liquid enclosing a gas bubble, which is partially wetted on velocity vector and pressure are respectively denoted by u = (u, v)
a cold curved surface. This cold surface causes the shell liquid of and p. σ denotes the interfacial tension coefficient, and κ is twice

2
T.V. Vu International Journal of Heat and Mass Transfer 181 (2021) 121881

Fig. 1. Schematic sketch of an axisymmetric single-core compound droplet solidifying on a curved surface held at temperature Tcold below the solidification point Tm of
the shell liquid. The curved surface has a radius denoted by Rw . Initially, the droplet is placed on the surface at an inner wetting angle θ i and outer angle θ o . During
solidification, solid, liquid and gas meet at three-phase lines, i.e., triple points, at which the growth angle φ gr is specified. In terms of the front-tracking technique, the
interface is represented by discrete points on a Eulerian background grid.

mean curvature. g is the gravitational acceleration, and f and H de- - For the interface point belongs to the gas-liquid interface, Vn is
note the momentum and energy forcing terms [44]. T denotes tem- an area-weighted interpolation from the velocities on four near-
perature except for superscript T denoting the transpose. Subscript est background grid nodes [27];
f denotes interface, and subscripts s and l denote solid and liquid, - For the interface point belongs to the solid-liquid interface
respectively. δ is the Dirac delta function, and x and xf denote the q˙ f
position vector and that indicating the location of the interface. nf Vn = (7)
is the unit vector normal to the interface. Lh is the latent heat of
ρs L h
solidification and q˙ – heat flux at the solid-liquid interface – is de- Here, q˙ f is calculated from Eq. (5) by a probe technique [45]:
fined as
1
∂T ∂T q˙ f = (ks (Ts − Tm ) − kl (Tm − Tl ) ) (8)
q˙ = ks − kl (5) h
∂n s
∂n l In Eq. (8), h is the background grid spacing. Ts is the tempera-
Eq. (4) stands for volume change upon solidification [25]. ture at the point in the solid, which is on the normal to the inter-
To solve the problem, we use an axisymmetric front-tracking face and at a distance h away from the interface point. Similarly,
method with the momentum and energy forcing terms as used in Tl is the temperature of the point on the normal drawn to the liq-
our previous study [34]. Accordingly, the left boundary of the com- uid phase and at the same distance h from the interface. Note that
putational domain is specified by the axisymmetric condition. The Eq. (6) applies for the gas-liquid and liquid-solid interfaces. The
top and right boundaries are defined with the full-slip condition point on the gas-solid interface is the intersection of the gas-liquid
while the bottom is non-slip. The temperature Tcold of the curved and liquid-solid interfaces along with applying a constant growth
surface is kept constant by the energy forcing term H, and the no- angle at the triple points.
slip boundary condition on the solid-liquid and solid-gas interfaces Gas, liquid, solid and their properties (e.g., ρ ) are specified by
is enforced by the momentum forcing term f, as described in detail two indicator functions I1 and I2 that are reconstructed from the
in [44]. location of the interface points [34]:
At a certain time t, the droplet interfaces including the liquid- ρ = I1 ρs + (1 − I1 )(I2 ρl + (1 − I2 )ρg ) (9)
gas, solid-gas and liquid-solid (or solidifying) interfaces are repre-
sented by Lagrangian grid points xf (t) on the fixed, rectangular In Eq. (9), subscript g represents gas. I1 equals 1 in the solid
background grid (Fig. 1). At the next time t + t, the new position phase and 0 in the other phases, and I2 equals 1 in the liquid phase
xf (t + t) of an interface point is given as and 0 in the other phases. The points belonging to the liquid-gas
interface are also used to compute the interfacial tension force.
x f (t + t ) = x f (t ) + tVn n f (6)
The rectangular background grid is used to solve the energy
In Eq. (6), Vn is specified as follows: and momentum equations by discretizing the spatial derivatives

3
T.V. Vu International Journal of Heat and Mass Transfer 181 (2021) 121881

τ = 0 the compound droplet and its surrounding gas are set at


Tn 0 = Tn (τ = 0) = 1.0. Here, τ and Tn are the dimensionless time
and normalized temperature defined as
 
τ = t kl / ρl Cpl R2o , Tn = (T − Tcold )/(Tm − Tcold ) (13)
The length scale is Ro and the velocity scale is Uc , Uc =
Fig. 2. Validation for a water droplet freezing on a cold curved surface [31]. The kl /(ρl C pl Ro ).
surface temperature is – 9.5°, and the surface radius is 15 mm. In (a) and (b), the
left side is the experimental droplet, and the right side is the computation droplet.
3. Grid and method validation

by central differences and a predictor-corrector method for time In the present study, we use the domain with a size specified as
integration. For more details on the method, one can find in, e.g., 3Ro × 6Ro and a grid resolution of 192 × 384. The domain size and
[25,27,34]. the grid resolution are based on our previous work [34] in which
Like our previous study [34], the following parameters are used: the grid resolution was coarser but still yielded reliable results.
Because detailed experimental data on the solidification pro-
C pl μl C pl (Tm − Tcold ) ρl gR2o μl cess of a compound droplet on a curved surface do not exist, we
Pr = , St = , Bo = ,Oh = (10)
kl Lh σ ρl R o σ here validate the method by performing the simulation of a sim-
ple droplet on a convex surface reported by Ju et al. [31]. As de-
Ri ρs ρg μg scribed in Ju et al. [31], at the beginning of solidification, the wa-
Rio = ,ρ = ,ρ = ,μ = (11) ter droplet is static, and the radius of the curved surface is 15 mm
Ro sl ρl gl ρl gl μl
(Fig. 2a). The surface was maintained at – 9.5°. The frozen droplets
ks kg C ps C pg are shown in Fig. 2b. The computed freezing time is about 22.6 s
ksl = , k = , C psl = ,C = (12) that results in a relative error of less than 8%, as compared to
kl gl kl C pl pgl C pl
the experimental value [31]. Basing on the properties of water,
Because of the presence of the curved surface and the ini- we use the following parameters for this comparison, Pr = 7.0,
tial and solidified shapes of the droplet specified by the wet- Oh = 0.01, St = 0.12, Bo = 0.29, ρ sl = 0.9, ρ gl = μgl = 0.05,
ting and growth angles, the present study is also characterized ksl = 4.0, Cpsl = 0.5, kgl = 0.05, Cpgl = 0.2 and φ gr = 0°. This com-
by Rw /Ro – the radius ratio of the curved surface to the outer parison shows that the numerical method predicts well the frozen
droplet, φ gr , θ i and θ o . To reduce the number of the parameters droplet after complete freezing. Other validations for a droplet on
under consideration, we fix the values of the following parameters: a flat surface can be found in [36,43], in which the morphology of
Pr = 0.01, Oh = 0.01, Bo = 1.0, ρ gl = μgl = 0.05, ksl = Cpsl = 1.0, the freezing front and droplet at the intermediate moments during
kgl = 0.005 and Cpgl = 0.24. We also assume θ i = 90° and at time the freezing process was provided. For compound droplets without

Fig. 3. Evolution of the solidifying front with the normalized velocity and temperature fields from (a) the beginning τ = 0 to (f) the end of solidification τ = τ o = 7.3. Other
times show (b) τ = 1.0, (c) τ = 1.7, (d) τ = τ i = 5.2 – right after completion of solidification around the inner core and (e) τ = 5.9. (g) Temporal variations of the average
height Hs of the solidifying front and its average velocity Vnsl . The horizontal arrow with its magnitude in (b) – (f) shows the reference vector. The dot lines in (g) show τ i
and τ o . The parameters are provided in the text.

4
T.V. Vu International Journal of Heat and Mass Transfer 181 (2021) 121881

Fig. 4. Effect of the outer wetting contact angle θ o . (a) Comparison between θ o = 60° and θ o = 120° at τ = 2.1 for the droplet shape with the normalized temperature and
normalized velocity fields. (b) Solidified droplet shape for various θ o . (c) Temporal variation of the average height Hs of the solidifying interface for various θ o . (d) Variations
with respect to θ o of the outer solidification time τ o , the inner solidification time τ i and the average space-time velocity Usl of the solidifying interface. (e) Variations with
respect to θ o of the tip angle α t and the additional height Hsf – H0 of the droplet. The horizontal arrow with its magnitude in (a) shows the reference vector. The lines in
(d) and (e) show the data fitting. The parameters are provided in the text.

phase change, validations can be found in, e.g., [46]. These valida- time” of the compound droplet. The tip angle at the solidified
tions support the accuracy of the computed results presented in droplet top is denoted by α t (Fig. 3f). It is found that α t is around
this paper. 65°.
Like the droplet on the horizontal surface [25,34], the droplet
on the curved surface has a high solidification rate during the ini-
4. Results and discussion tial stage of solidification. Therefore, the average growth velocity
Vnsl of the solidifying front, i.e., the average value of the veloci-
A typical case of solidification on a curved surface with St = 0.1, ties of all points on the solid-liquid interface, is rather high during
Rw /Ro = 10, Ri /Ro = 0.5, ρ sl = 0.9, φ gr = 0° and θ o = 90° is shown this stage (the circle line in Fig. 3g). This results in a high slop of
in Fig. 3. Rw /Ro = 10 indicates that the cold surface is concave and the average height Hs of the solidifying interface (the triangle line
thus, as compared to that on the horizontal surface [25,34], the in Fig. 3g). We also see a very high Vnsl during the last stage of
compound droplet on this surface is higher at τ = 0.0 (Fig. 3a). solidification when the solid-liquid interface is almost at the top.
More liquid accumulates at the center as compared to that on As mentioned in our previous work [34], unlike the simple droplet
the flat surface. Accordingly, we observe a rather strong velocity [25], we see a sudden change in the slope of the line Hs (τ ) before
field at τ = 1.0 and 1.7 before the liquid phase completely solid- and after the moment of completing solidification around the in-
ifies around the inner core (at τ = τ i = 5.2, Fig. 3d). The veloc- ner core, i.e., at time τ i . This is because no heat directly releases to
ity at these times also indicates that the droplet oscillates during the gas phase in the inner core that has a lower heat transfer rate
this stage of solidification. This oscillation results in the wavy sur- as compared to the solid phase after τ = τ i . After complete solid-
face on the outer droplet after solidification [34]. The time τ i , i.e., ification, the height of the droplet is added by a value called the
time for complete solidification around the inner core, is called “additional height” Hsf – H0 . Here, Hsf is the height of the solidified
the “inner solidification time”. After this time, the droplet top is droplet, Hsf = Hs (τ = τ o ), (Fig. 3f) and H0 is the initial height of
less oscillated, and there is only the presence of the flow induced the liquid compound droplet. It is found that Hsf – H0 is equal to
by volume expansion because of solid less dense than liquid (i.e., about 0.156H0 while the volume of the droplet increases about 11%
ρ sl = 0.9, Fig. 3e) [25,34]. Consequently, the droplet forms an apex after complete solidification because of volume expansion.
at the top after complete solidification at τ = τ o = 7.3 (Fig. 3f). τ o In the following, the influences of some main parameters will
is time for completing the entire solidification process and called be revealed.
the “outer solidification time”, or simply called “the solidification

5
T.V. Vu International Journal of Heat and Mass Transfer 181 (2021) 121881

Fig. 5. Effect of the surface radius Rw /Ro . (a) Comparison between Rw /Ro = −5 and Rw /Ro = 5 at τ = 2.2 for the droplet shape with the normalized temperature and
normalized velocity fields. (b) Solidified droplet shape for various Rw /Ro . (c) Temporal variation of the average height Hs of the solidifying interface for various Rw /Ro . (d)
Variations with respect to Rw /Ro of the outer solidification time τ o , the inner solidification time τ i and the average space-time velocity Usl of the solidifying interface. (e)
Variations with respect to Rw /Ro of the tip angle α t and the additional height Hsf – H0 of the droplet. The horizontal arrow with its magnitude in (a) shows the reference
vector. The lines in (d) and (e) show the data fitting. The parameters are provided in the text.

4.1. Effect of the outer wetting angle θ o In Eq. (14), Vnsl is the average velocity of all points of the solid-
ifying interface at time t. to is the solidification time or the outer
When a droplet stays on a surface, the droplet may form on the solidification time, i.e., time to complete the entire solidification
surface at a certain wetting angle which depends on the droplet process. We also use ti called “the inner solidification time” which
properties and surface morphology. The influences of the outer is time to complete the solidification process around the inner
droplet shape at the beginning of solidification, in terms of the core. In other words, ti is the time for the triple point on the inner
outer wetting angle, are depicted in Fig. 4. Note that the value interface to move from the bottom to the top of the inner core.
of θ o is varied but the drop volume and thus the size Ro remain The corresponding dimensionless times are τ o and τ i (Fig. 4d), as
unchanged. The parameters are St = 0.1, Rw /Ro = 10, Ri /Ro = 0.5, earlier mentioned.
ρ sl = 0.9, φ gr = 0° and θ o = 60° – 130°. For θ o = 60°, the droplet A linear fit to the variation of Usl with respect to θ o (in radian)
has a wide spreading with a large area of contact with the surface gives Usl /Uc = – 0.1722θ o + 0.5165 with the maximum relative er-
(the left side of Fig. 4a). Increasing the value of θ o to 120° while ror of 5.45% (the dash-dot line in Fig. 4d). In accordance with a
keeping the inner core unchanged decreases the contact area and reduction in Usl , the inner and outer solidification times τ i and τ o
makes more liquid accumulate in the center region. Consequently, are increased with an increase in the outer wetting angle, as in-
at time τ = 2.1, the liquid comes down more strongly in the case dicated by the circles (τ i ) and the squares (for τ o ) in Fig. 4d. The
of the large θ o than in the case of the small θ o , as illustrated in second-order polynomial fits to these times provide τ o = 8.1112θ o 2
Fig. 4a. As a result of a smaller contact area, the heat released from – 18.056θ o + 15.342 (the dash line, Fig. 4d) and τ i = 3.869θ o 2 –
the liquid decreases during solidification by increasing the contact 9.14θ o + 9.7744 (the solid line, Fig. 4d). The maximum errors of
angle. Therefore, Fig. 4a indicates, at this time, that the location these fits are 8.5% and 6.5%, respectively, for τ o and τ i .
of the solidifying front for θ o = 60° has progressed further than Because of the same volume, increasing the outer wetting an-
for θ o = 120°. It confirms that increasing the contact angle slows gle increases the initial height H0 of the compound droplet. Hence,
down the solidification process as shown in Fig. 4c. Therefore, an the height of the compound droplet (i.e., the height of the solid-
increase in θ o leads to a decrease in the averaged space-time ve- ified droplet Hsf ) after complete solidification increases with θ o
locity of the solidifying interface Usl (the triangle in Fig. 4d) de- (Fig. 4b and c). However, the additional height of the droplet at
fined as the end of the solidification process, i.e., Hsf – H0 , is decreased
with θ o (the circles in Fig. 4e). The presence of a non-zero Hsf –
 H0 is caused by volume expansion due to density difference be-
1
Usl = Vnsl dt (14) tween the solid and liquid phases, i.e., ρ sl = 0.9 [28,47]. As indi-
to

6
T.V. Vu International Journal of Heat and Mass Transfer 181 (2021) 121881

Fig. 6. Effect of the solid-to-liquid density ratio ρ sl . (a) Comparison between ρ sl = 0.8 and ρ sl = 1.1 at τ = 3.6 for the droplet shape with the normalized temperature and
normalized velocity fields. (b) Solidified droplet shape for various ρ sl . (c) Temporal variation of the average height Hs of the solidifying interface for various ρ sl . (d) Variations
with respect to ρ sl of the outer solidification time τ o , the inner solidification time τ i and the average space-time velocity Usl of the solidifying interface. (e) Variations with
respect to ρ sl of the tip angle α t and the additional height Hsf – H0 of the droplet. The horizontal arrow with its magnitude in (a) shows the reference vector. The lines in
(d) and (e) show the data fitting. The parameters are provided in the text.

cated in Fig. 4e, it is found that a decrease in Hsf – H0 with θ o (in 4.2. Effect of the curved surface radius Rw
radian) follows (Hsf – H0 )/Ro = – 0.1019θ o 2 + 0.1997θ o + 0.1605
with the maximum error of 1.2% (the solid line in Fig. 4e). More- Fig. 5a shows a compound droplet solidifying at τ = 2.2 on a
over, as a result of volume expansion, the solidified droplet has an convex surface with Rw /Ro = –5 (left side) in comparison with that
apex at the droplet top [34]. This apex corresponds to a tip an- on a concave surface with Rw /Ro = 5 (right side). The other param-
gle α t . It is found that increasing θ o in the range of 60° – 130° eters are St = 0.1, Ri /Ro = 0.5, ρ sl = 0.9, φ gr = 0° and θ o = 100°. At
has a minor effect on the tip angle (the squares in Fig. 4e). It is this time, the liquid phase is moving up, but with different mag-
consistent with the experimental observations for simple droplets nitudes for two droplets. The droplet on the concave surface (in
[31]. We find that the droplet with θ o in the range of 60° – the right side of Fig. 5a) has a stronger flow field than that in the
130° yields a tip angle α t = 64.7° ± 1.43°. Our result for the left one because the droplet on the concave surface has more liq-
tip angle can be referred to water droplets that also have simi- uid accumulating at the droplet center than that on the convex
lar volume expansion (i.e., ρ sl ≈ 0.9) and the growth angle (i.e., surface. As a result of increasing the value of Ro /Rw from a more
φ gr = 0°) upon freezing [15]. Marin and co-workers [15] experi- convex surface (Rw < 0) to a more concave surface (Rw > 0), the
mentally found that for frozen water drops the tip angle α t is var- compound droplet becomes higher with more liquid accumulating
ied from 65.5° to 73.5°. Theoretically, these authors found α t = 64° at the center, and thus the solidified droplet height increases with
– 69°. It is evident that our computational tip angles are in close Ro /Rw (Fig. 5b). Consequently, the solidification process needs more
agreement with the theoretical and experimental results of Marin time τ o to complete the entire fluid of the shell when Ro /Rw in-
et al. [15]. Even though θ o has an effect on the inner solidification creases from – 1/5 to 1/5, as indicated in Fig. 5d (the square sym-
time but the variation of θ o from 60° to 130° makes no change bols). It is found that τ o = 1.2301Ro /Rw + 8.3756 with the max-
in the shape of the inner core after solidification, as exemplified imum error of 0.67% (the dash line in Fig. 5d). Even though τ o
in Fig. 4b. increases with Ro /Rw , the mean solidification rate is slightly in-
Moreover, Fig. 4a and b indicates that the droplet with fluenced by the shape of the surface (the triangle symbols with
θ o = 120° causes stronger undulation on the outer surface of the the dash-dot line keeping horizontal). This is not what we had ex-
solidified drop after complete solidification as compared to that pected (i.e., increasing Ro /Rw would decrease the rate of solidifica-
with a lower wetting angle. This is because increasing the value tion).
of the wetting angle causes the liquid to accumulate more near the However, we see that the inner triple point has gone closer to
axis of symmetry, and thus causes the droplet top to more oscillate the axis of symmetry for Rw /Ro = 5 than that for Rw /Ro = –5 at
during the solidification [34]. τ = 2.2. This indicates that when changing from the convex sur-

7
T.V. Vu International Journal of Heat and Mass Transfer 181 (2021) 121881

Fig. 7. Effect of the inner-to-outer radius ratio Ri /Ro . (a) Comparison between Ri /Ro = 0.4 and Ri /Ro = 0.7 at τ = 2.5 for the droplet shape with the normalized temperature
and normalized velocity fields. (b) Solidified droplet shape for various Ri /Ro . (c) Temporal variation of the average height Hs of the solidifying interface for various Ri /Ro . (d)
Variations with respect to Ri /Ro of the outer solidification time τ o , the inner solidification time τ i and the average space-time velocity Usl of the solidifying interface. (e)
Variations with respect to Ri /Ro of the tip angle α t and the additional height Hsf – H0 of the droplet. The horizontal arrow with its magnitude in (a) shows the reference
vector. The lines in (d) and (e) show the data fitting. The parameters are provided in the text.

face to the concave surface, the compound droplet terminates its liquid seems to flow into the solid phase for ρ sl = 1.1 as a re-
solidification process around the inner core earlier (Fig. 5c). This is sult of volume shrinkage. The other parameters for the calculation
confirmed by the circles shown in Fig. 5d. A line matching these shown in Fig. 6 are St = 0.1, Ri /Ro = 0.5, Rw /Ro = 10, φ gr = 0°
circles gives τ i = –1.6635Ro /Rw + 5.765. The maximum error of and θ o = 100°. Because of this effect of volume change upon so-
this fit is 1.1%. Even though an increase in Ro /Rw leads to an in- lidification, the final solidified product, i.e., the form of the solid-
crease in the solidified compound droplet height (Fig. 5b and c), ified droplet, behaves differently when the density ratio changes
it has no effect on the additional height of the droplet, as shown between 0.8 and 1.2. Volume expansion (ρ sl < 1.0) produces a so-
by the circles in Fig. 5e (the solid line remaining horizontal from lidified droplet with a conical top while volume shrinkage (ρ sl >
left to right). Consequently, the tip angle is also independent of the 1.0) induces a cavity, as illustrated in Fig. 6b. Such cavity formation
shape of the contact surface (Fig. 5b and the squares in Fig. 5e). It is frequently observed in casting [37]. In consequence, the solidi-
is found that the variation of Ro /Rw from –1/5 to 1/5 yields a tip fied droplet becomes smaller by raising the density ratio (Fig. 6b
angle α t = 65° ± 0.5° (the dash line in Fig. 5e). This finding is and c) [34,37]. However, volume change introduces a very minor
consistent with the experimental observation reported by Ju et al. effect on the form of the inner core.
[31] in which the cone on top of the frozen water droplets was al- The comparison in Fig. 6a also indicates that at τ = 3.6, the
most constant for different convex surfaces. In addition, the values solid-liquid interface in the case of ρ sl = 0.8 has moved further
of the tip angle is in agreement with experiments of Marin et al. than that for ρ sl = 1.1. It illustrates that volume shrinkage reduces
[15] and again confirms that the formation of the cone at the drop the solidification rate while volume expansion increases the rate of
top is almost independent of the cold surface morphology as re- solidification, as confirmed in Fig. 6d, where increasing the value
ported by Ismail and Waghmare [48]. of ρ sl leads to a decrease in Usl (the triangles). It is found that
Usl /Uc = – 0.1421ρ sl + 0.3454 (the dash-dot line). Because of a de-
crease in Usl when increasing ρ sl , the time τ i necessary to finish
4.3. Effect of the density ratio ρ sl
solidification around the inner core increases with ρ sl , as shown in
Fig. 6d (the circles). A linear fit provides τ i = 4.4667ρ sl + 1.5528,
It has been demonstrated that the liquid moves away from the
yielding the maximum error of 1.3% (the solid line in Fig. 6d).
solidifying interface when the volume expands while it flows to-
In contrast to the variation of τ i with respect to ρ sl , increas-
wards the phase-change interface when the volume shrinks [25].
ing ρ sl leads to a decrease in the time τ o necessary to complete
Accordingly, at τ = 3.6 (Fig. 6a) there is a flow from the solid-
the entire process (the squares in Fig. 6d), i.e., τ o = 7.1061ρ sl 2 –
liquid interface while the liquid near the droplet top is going down
18.879ρ sl + 19.736 (with the maximum error of 0.2%) (the dash
for ρ sl = 0.8, i.e., volume expansion. In contrast, at this time the

8
T.V. Vu International Journal of Heat and Mass Transfer 181 (2021) 121881

Fig. 8. Effect of growth angle φ gr . (a) Comparison between φ gr = 5° and φ gr = 15° at τ = 2.3 for the droplet shape with the normalized temperature and normalized velocity
fields. (b) Solidified droplet shape for various φ gr . (c) Temporal variation of the average height Hs of the solidifying interface for various φ gr . (d) Variations with respect to
φ gr of the outer solidification time τ o , the inner solidification time τ i and the average space-time velocity Usl of the solidifying interface. (e) Variations with respect to φ gr of
the tip angle α t and the additional height Hsf – H0 of the droplet. The horizontal arrow with its magnitude in (a) shows the reference vector. The lines in (d) and (e) show
the data fitting. The parameters are provided in the text.

line in Fig. 6d). Even though the rate of solidification decreases, corresponds to the increase in the arc length to be completed by
this decrease in τ o when changing from expansion to shrinkage the inner triple point (Fig. 7a). Also, because of a larger inner core,
is understandable because with the same volume of solid formed there is an opposite circulation in the inner gas phase as compared
the volume of the remaining liquid phase of the droplet increases to that in the liquid phase for Ri /Ro = 0.7, whereas such circula-
with a decrease in ρ sl from 1.2 to 0.8. Finally, the solidifying inter- tion is not available in the gaseous phase inside the inner core for
face takes longer to propagate to the droplet top. As a result, the Ri /Ro = 0.3 (Fig. 7a).
additional height Hsf – H0 of the compound droplet changes from As indicated in Fig. 7c, the average height profiles Hs of the so-
a positive value to a negative value when ρ sl increases from 0.8 lidification interface for different shell liquid thickness almost over-
to 1.2 (the circles in Fig. 6e). The second-order polynomial profile, lap during solidification around the inner core. However, they be-
(Hsf – H0 )/Ro = 2.652ρ sl 2 – 7.8508ρ sl + 5.1156, (the solid line in come different after that. Consequently, the whole process takes
Fig. 6e) is a good approximation, i.e., the maximum error equal to more time, i.e., τ o (the squares in Fig. 7c), when the size of
3.14%. Unlike τ o and the additional height, the tip angle increases the inner core increases. An approximation to this increase gives
with ρ sl (the squares in Fig. 6f). It is found that this increase in the τ o = 11.267(Ri /Ro )2 – 4.6536Ri /Ro + 7.9625 with the maximum
tip angle follows well α t (in°) = 173.11ρ sl – 88.372 (the dash line error of 0.65% (the dash line). The increase in the solidification
in Fig. 6f). The error is less than 4.2%. This finding is in line with time confirms that the mean rate of solidification decreases with
our earlier work [25,34]. a decrease in the thickness of the shell (the triangles with the fit
Usl /Uc = – 0.129Ri /Ro + 0.2801 represented by the dash-dot line).
This decrease is understandable because the decrease in the shell
4.4. Effect of the shell liquid thickness
liquid thickness corresponds to the decrease in the initial wetting
liquid area, and thus the effect of the cold surface becomes weaker.
Fig. 7 illustrates the effects of the shell liquid thickness, in
The decrease in the wetting area (i.e., the contact area with the
terms of Ri /Ro , on the solidification process of the droplet on the
curved surface) when increasing Ri /Ro can be referred to increasing
curved surface with Rw /Ro = 10. The other parameters are St = 0.1,
the wetting contact angle leading to the increase in the solidifica-
Ri /Ro = 0.3 – 0.8, φ gr = 0°, ρ sl = 0.9 and θ o = 100°. The thickness
tion time [49] as earlier mentioned.
of the shell increases with a decrease in Ri /Ro , as exemplified in
Because the variation of Ri /Ro contributes most to the inner
Fig. 7a. Accordingly, the inner solidification time increases with a
core and the shape of the inner core after complete solidification
decrease in the shell thickness (the circles in Fig. 7d). It is found
still remains like a spherical cap, the outer shape of the solidified
that τ i increases linearly with Ri /Ro , i.e., τ i = 17.629Ri /Ro – 3.2374
droplet is not affected by the variation of the initial shell liquid
(the solid line in Fig. 7d). This is understood as the increase in Ri /Ro

9
T.V. Vu International Journal of Heat and Mass Transfer 181 (2021) 121881

Fig. 9. Effect of growth angle St. (a) Comparison between St = 0.05 and St = 0.4 at τ = 2.3 for the droplet shape with the normalized temperature. (b) Solidified droplet
shape for various St. (c) Temporal variation of the average height Hs of the solidifying interface for various St. (d) Variations with respect to St of the outer solidification time
τ o , the inner solidification time τ i and the average space-time velocity Usl of the solidifying interface. (e) Variations with respect to St of the tip angle α t and the additional
height Hsf – H0 of the droplet. The lines in (d) and (e) show the data fitting. The parameters are provided in the text.

thickness (Fig. 7b). As a consequence, the additional height of the effects of the growth angle are also in accordance with experimen-
droplet and the tip angle at the droplet top are slightly affected by tal observations, e.g., an indium antimonide droplet with a growth
Ri /Ro (Fig. 7d). angle of 28° is higher and more conical than a germanium droplet
with a growth angle of 14° [28].
4.5. Effect of the growth angle However, the average growth rate is almost independent of the
change in φ gr (the triangles in Fig. 8d). Accordingly, the inner so-
Fig. 8a compares the droplet shape, the velocity and tempera- lidification time τ i is also independent of φ gr ranging between 0°
ture fields at τ = 2.3 between φ gr = 5° and φ gr = 15°. The other and 15° (the circles in Fig. 8d) even though the height of the inner
parameters include St = 0.1, Rw /Ro = 10, Ri /Ro = 0.5, ρ sl = 0.9 core after solidification is decreased (Fig. 8b). It can be understood
and θ o = 100°. At this time, the velocity for φ gr = 15° is stronger as illustrated in Fig. 8c, in which the variation of φ gr has a very
than that for φ gr = 5°. As a result, the compound droplet for minor effect on the evolution of the solidifying interface when the
φ gr = 15° is higher than that for φ gr = 5°. Consequently, the so- droplet performs solidification around the inner core.
lidified droplet after complete solidification becomes higher as the
growth angle increases, as shown in Fig. 8b. Because the initial 4.6. Effect of the Stefan number
height of the droplet at the start of the process remains unchanged
for all growth angles, the additional height of the droplet at the The effects of St on the solidification of the compound droplet
end of solidification increases with φ gr from 0° to 15° (the cir- on the curved surface is shown in Fig. 9 with Rw /Ro = 10,
cles in Fig. 8e). It is found that this increase follows well (Hsf – Ri /Ro = 0.5, ρ sl = 0.9, φ gr = 0° and θ o = 100°. With a low Ste-
H0 )/Ro = 1.5543φ gr + 0.1931, (the solid line in Fig. 8e and φ gr in fan number, e.g., St = 0.05 (the left side of Fig. 9a), at τ = 2.3
radian), where the maximum error is found to be 2.45%. As a re- only a portion of the liquid phase has been solidified. However,
sult of increasing the droplet height, the angle at the droplet top increasing the value of St to 0.4 (the right side of Fig. 9a), the
decreases with the increase in φ gr (the squares in Fig. 8e), i.e., liquid-solid interface is close to the top at this time. This is be-
α t = –1.2436φ gr + 1.1244 (the dash line in Fig. 8e, the maximum cause the increase in St decreases the latent heat of solidification
error of 1.6%). The increase in the height of the droplet also leads and thus increases the solidification rate, as illustrated by Usl var-
to an increase in the entire solidification time τ o (the squares in ied with St shown in Fig. 9d (the triangles). Usl is observed to in-
Fig. 8d), i.e., τ o = 7.62φ gr + 8.4333 (the dash line in Fig. 8d and crease linearly with St by Usl /Uc = 1.4412 St + 0.1 (the dash-dot
φ gr in radian). The increase in the droplet height and the decrease line in Fig. 9d). Because of increasing the rate of solidification, the
in the tip angle with respect to the increase in the growth angle is inner and outer solidification times τ i (the circles in Fig. 9d) and
consistent with the theoretical calculation of Satunkin [28]. These τ o (the squares in Fig. 9d) decrease rapidly with an increase in the

10
T.V. Vu International Journal of Heat and Mass Transfer 181 (2021) 121881

Stefan number, i.e., τ i = 0.6617St0.936 (the solid line in Fig. 9d) and Acknowledgments
τ o = 1.1082St0.898 (the dash line in Fig. 9d). The maximum errors
for these fittings are 3.8% (for τ i ) and 6.7% (for τ o ). This research is funded by Vietnam National Foundation for Sci-
Another way for interpretation of increasing the St number is ence and Technology Development (NAFOSTED) under grant num-
to increase the supercooling effect by decreasing the temperature ber 107.03-2019.307.
of the curved surface (Eq. (10)). Accordingly, the decrease of the
curved surface temperature results in the decrease of the solidi- References
fication time. This tendency is completely consistent with exper-
imental measurements of, e.g., Ju et al. [31] or Zhang et al. [33]. [1] C. Augustin, W. Hungerbach, Production of hollow spheres (HS) and hollow
sphere structures (HSS), Mater. Lett. 63 (2009) 1109–1112.
Therefore, the temporal change in the average height of the solid- [2] A.A. Maan, K. Schroën, R. Boom, Spontaneous droplet formation techniques for
ification front becomes more lateral when decreasing the value of monodisperse emulsions preparation – Perspectives for food applications, J.
the Stefan number, as exemplified in Fig. 9c. Food Eng. 107 (2011) 334–346.
[3] V. Lapkovskis, V. Mironovs, Excitation of lightweight steel hollow spheres
In contrast, the variation of St in the range of 0.025 – 1.8 has a
by means of pulsed electromagnetic field, Procedia Comput. Sci. 104 (2017)
minor effect on the shape of the compound droplet after complete 408–412.
solidification, as indicated in Fig. 9b. Consequently, the tip angle α t [4] M.L. Fabiilli, J.A. Lee, O.D. Kripfgans, P.L. Carson, J.B. Fowlkes, Delivery of water–
and the additional height Hsf – H0 are slightly affected by St. This is soluble drugs using acoustically triggered perfluorocarbon double emulsions,
Pharm. Res. 27 (2010) 2753–2765.
consistent with the finding in simple droplet experiments in which [5] K.D. Bhagat, T.V. Vu, J.C. Wells, H. Takakura, Y. Kawano, F. Ogawa, Production of
the tip angle is almost independent of the cooling rate [15]. hollow germanium alloy quasi-spheres through a coaxial nozzle, Jpn. J. Appl.
Phys. 58 (2019) 068001.
[6] T.V. Vu, H. Takakura, J.C. Wells, T. Minemoto, Production of hollow spheres of
5. Conclusion eutectic tin-lead solder through a coaxial nozzle, J. Solid Mech. Mater. Eng. 4
(2010) 1530–1538.
We have presented the numerical results, obtained using the [7] M. Cao, F. Jiang, C. Wang, H. Cui, C. Guo, Y. Chang, Z. Wang, Preparation, mi-
crostructure and mechanical property of double-layered metal-ceramic hollow
front-tracking method, of a compound droplet that solidifies on a spheres, Mater. Sci. Eng., A 780 (2020) 139188.
cold curved surface. The compound droplet contains an inner core [8] I.B. Bazhlekov, P.J. Shopov, Z.D. Zapryanov, Unsteady motion of a type-A com-
of gas, and the curved surface of radius Rw can be concave (Rw pound multiphase drop at moderate reynolds numbers, J. Colloid Interf. Sci.
169 (1995) 1–12.
> 0) or convex (Rw < 0). The time necessary to complete the en- [9] S. Tasoglu, G. Kaynak, A.J. Szeri, U. Demirci, M. Muradoglu, Impact of a com-
tire solidification process, i.e., the outer solidification time τ o , in- pound droplet on a flat surface: a model for single cell epitaxy, Phys. Fluid. 22
creases when the curved surface changes from a convex shape to (2010) 082103.
[10] Y.( ) Wei, M.-J.(  ) Thoraval, Maximum spreading of an impacting
a concave one. In contrast, such a variation from a more convex
air-in-liquid compound drop, Phys. Fluid. 33 (2021) 061703.
surface to a more concave surface causes a decrease in the inner [11] M.J. Neeson, R.F. Tabor, F. Grieser, R.R. Dagastine, D.Y. Chan, Compound sessile
solidification time τ i , i.e., time for completing solidification around drops, Soft Matter 8 (2012) 11042–11050.
the inner core. However, the tip angle, i.e., the angle defined at the [12] S. Shklyaev, A.O. Ivantsov, M. Díaz-Maldonado, U.M. Córdova-Figueroa, Dynam-
ics of a Janus drop in an external flow, Phys. Fluid. 25 (2013) 082105.
droplet top, and the additional height, i.e., the difference between [13] N. Blanken, M.S. Saleem, M.-.J. Thoraval, C. Antonini, Impact of compound
the initial height and the final height of the droplet, after complete drops: a perspective, Curr. Opin. Colloid Interface Sci. 51 (2021) 101389.
solidification are independent of the surface radius. In contrast to [14] O.R. Enríquez, Á.G. Marín, K.G. Winkels, J.H. Snoeijer, Freezing singularities in
water drops, Phys. Fluid. 24 (2012) 091102–091102–2.
the curved surface, a decrease in the solid-to-liquid density ratio [15] A.G. Marin, O.R. Enriquez, P. Brunet, P. Colinet, J.H. Snoeijer, Universality of
ρ sl or an increase in the growth angle φ gr leads to a more coni- tip singularity formation in freezing water drops, Phys. Rev. Lett. 113 (2014)
cal solidified droplet. Further investigations on the Stefan number 054301.
[16] H. Peng, Q. Wang, T. Wang, L. Li, Z. Xia, J. Du, B. Zheng, H. Zhou, L. Ye, Study on
St (from 0.025 to 1.6), the shell liquid thickness (in terms of Ri /Ro dynamics and freezing behaviors of water droplet on superhydrophobic alu-
from 0.3 to 0.8) and the shape of the outer droplet (in terms of minum surface, Appl. Phys. A 126 (2020) 811.
the outer wetting angle θ o from 60° to 130°) reveal that (i) τ i and [17] Y. Pan, K. Shi, X. Duan, G.F. Naterer, Experimental investigation of water
droplet impact and freezing on micropatterned stainless steel surfaces with
τ o increase with increasing θ o or Ri /Ro or decreasing St; (ii) the varying wettabilities, Int. J. Heat Mass Transf. 129 (2019) 953–964.
tip angle induced by volume expansion is slightly affected by the [18] J.H. Snoeijer, P. Brunet, Pointy ice-drops: how water freezes into a singular
variations in these parameters; (iii) the additional height increases shape, Am. J. Phys. 80 (2012) 764–771.
with the decrease in θ o , and is slightly affected by Ri /Ro or St. [19] X. Zhang, X. Liu, J. Min, X. Wu, Shape variation and unique tip formation of a
sessile water droplet during freezing, Appl. Therm. Eng. 147 (2019) 927–934.
As far as we are aware, the present numerical simulations pro- [20] M. Tembely, A. Dolatabadi, A comprehensive model for predicting droplet
vide a more complete picture of the solidification process of com- freezing features on a cold substrate, J. Fluid Mech. 859 (2019) 566–585.
pound droplets on cold surfaces. However, further investigations [21] A. Virozub, I.G. Rasin, S. Brandon, Revisiting the constant growth angle: esti-
mation and verification via rigorous thermal modeling, J. Cryst. Growth 310
on, e.g., compound droplets with various inner wetting angles, (2008) 5416–5422.
droplets under forced convection or droplets with two or more [22] H. Shetabivash, A. Dolatabadi, M. Paraschivoiu, A multiple level-set approach
inner cores should be performed. Fully three-dimensional simula- for modelling containerless freezing process, J. Comput. Phys. (2020) 109527.
[23] X. Zhang, X. Liu, X. Wu, J. Min, Simulation and experiment on supercooled ses-
tions are also challenging for such problems with the presence of sile water droplet freezing with special attention to supercooling and volume
nearby droplets. In addition, a theoretical analysis should be con- expansion effects, Int. J. Heat Mass Transf. 127 (2018) 975–985.
ducted to provide more complete prediction tools of investigations [24] C. Zhang, H. Zhang, W. Fang, Y. Zhao, C. Yang, Axisymmetric lattice Boltzmann
model for simulating the freezing process of a sessile water droplet with vol-
on the problem. ume change, Phys. Rev. E 101 (2020) 023314.
[25] T.V. Vu, G. Tryggvason, S. Homma, J.C. Wells, Numerical investigations of drop
Declaration of Competing Interest solidification on a cold plate in the presence of volume change, Int. J. Multiph.
Flow 76 (2015) 73–85.
[26] T.V. Vu, K.V. Dao, B.D. Pham, Numerical simulation of the freezing process of a
The author declares that he has no known competing financial water drop attached to a cold plate, J. Mech. Sci. Technol. 32 (2018) 2119–2126.
interests or personal relationships that could have appeared to in- [27] G. Tryggvason, B. Bunner, A. Esmaeeli, D. Juric, N. Al-Rawahi, W. Tauber, J. Han,
S. Nas, Y.-.J. Jan, A front-tracking method for the computations of multiphase
fluence the work reported in this paper.
flow, J. Comput. Phys. 169 (2001) 708–759.
[28] G.A. Satunkin, Determination of growth angles, wetting angles, interfacial ten-
CRediT authorship contribution statement sions and capillary constant values of melts, J. Cryst. Growth 255 (2003)
170–189.
[29] H. Itoh, H. Okamura, C. Nakamura, T. Abe, M. Nakayama, R. Komatsu, Growth
Truong V. Vu: Conceptualization, Methodology, Validation, In- of spherical Si crystals on porous Si3N4 substrate that repels Si melt, J. Cryst.
vestigation, Writing – review & editing. Growth 401 (2014) 748–752.

11
T.V. Vu International Journal of Heat and Mass Transfer 181 (2021) 121881

[30] Z. Jin, H. Zhang, Z. Yang, The impact and freezing processes of a water droplet [40] M. Song, D. Hu, X. Zheng, L. Wang, Z. Yu, W. An, R. Na, C. Li, N. Li, Z. Lu,
on different cold cylindrical surfaces, Int. J. Heat Mass Transf. 113 (2017) Enhancing droplet deposition on wired and curved superhydrophobic leaves,
318–323. ACS Nano 13 (2019) 7966–7974.
[31] J. Ju, Z. Jin, H. Zhang, Z. Yang, J. Zhang, The impact and freezing processes of [41] H.-.R. Liu, C.-.Y. Zhang, P. Gao, X.-.Y. Lu, H. Ding, On the maximal spreading of
a water droplet on different cold spherical surfaces, Exp. Therm. Fluid Sci. 96 impacting compound drops, J. Fluid Mech. 854 (2018).
(2018) 430–440. [42] Z. Jin, X. Cheng, Z. Yang, Experimental investigation of the successive freezing
[32] D. Wu, P. Wang, P. Wu, Q. Yang, F. Liu, Y. Han, F. Xu, L. Wang, Determination processes of water droplets on an ice surface, Int. J. Heat Mass Transf. 107
of contact angle of droplet on convex and concave spherical surfaces, Chem. (2017) 906–915.
Phys. 457 (2015) 63–69. [43] V.N. Duy, T.V. Vu, A numerical study of a liquid drop solidifying on a vertical
[33] H. Zhang, Z. Jin, M. Jiao, Z. Yang, Experimental investigation of the impact and cold wall, Int. J. Heat Mass Transf. 127 (2018) 302–312.
freezing processes of a water droplet on different cold concave surfaces, Int. J. [44] T.V. Vu, J.C. Wells, Numerical simulations of solidification around two tandem-
Therm. Sci. 132 (2018) 498–508. ly-arranged circular cylinders under forced convection, Int. J. Multiph. Flow 89
[34] N.X. Ho, T.V. Vu, B.D. Pham, A numerical study of a liquid compound drop (2017) 331–344.
solidifying on a horizontal surface, Int. J. Heat Mass Transf. 165 (2021) 120713. [45] A. Esmaeeli, G. Tryggvason, Computations of film boiling. Part I: numerical
[35] J.H. Nadler, T.H.Jr. Sanders, J.K. Cochran, Aluminum hollow sphere processing, method, Int. J. Heat Mass Transf. 47 (2004) 5451–5461.
Mater. Sci. Forum 331–337 (20 0 0) 495–50 0. [46] T.-.V. Vu, T.V. Vu, C.T. Nguyen, P.H. Pham, Deformation and breakup of a dou-
[36] T.V. Vu, Q.H. Luu, Containerless solidification of a droplet under forced convec- ble-core compound droplet in an axisymmetric channel, Int. J. Heat Mass
tion, Int. J. Heat Mass Transf. 143 (2019) 118498. Transf. 135 (2019) 796–810.
[37] M. Raessi, J. Mostaghimi, Three-dimensional modelling of density varia- [47] D.M. Anderson, M.G. Worster, S.H. Davis, The case for a dynamic contact angle
tion due to phase change in complex free surface flows, Numer. Heat in containerless solidification, J. Cryst. Growth 163 (1996) 329–338.
Tr. B–Fund 47 (2005) 507–531 http://www.tandfonline.com/doi/pdf/10.1080/ [48] M.F. Ismail, P.R. Waghmare, Universality in freezing of an asymmetric drop,
10407790590928964. (accessed June 19, 2015). Appl. Phys. Lett. 109 (2016) 234105.
[38] Q. Liu, M. Orme, High precision solder droplet printing technology and the [49] H. Zhang, Y. Zhao, R. Lv, C. Yang, Freezing of sessile water droplet for various
state-of-the-art, J. Mater. Process. Technol. 115 (2001) 271–283. contact angles, Int. J. Therm. Sci. 101 (2016) 59–67.
[39] P. Gao, J.J. Feng, Spreading and breakup of a compound drop on a partially
wetting substrate, J. Fluid Mech. 682 (2011) 415–433.

12

You might also like