You are on page 1of 17

Applied Thermal Engineering 219 (2023) 119360

Contents lists available at ScienceDirect

Applied Thermal Engineering


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

Research Paper

Impact and boiling characteristics of a droplet on heated surfaces: A 3D


lattice Boltzmann study
Yunjie Xu a, Linlin Tian a, Chunling Zhu a, b, Ning Zhao a, b, *
a
College of Aerospace Engineering, Nanjing University of Aeronautics and Astronautics, Nanjing, Jiangsu 210016, China
b
State Key Laboratory of Mechanics and Control of Mechanical Structures, Nanjing University of Aeronautics and Astronautics, Nanjing, Jiangsu 210016, China

A R T I C L E I N F O A B S T R A C T

Keywords: Droplet impingement and boiling on heated surfaces play an important role in many industrial applications. Due
Droplet impact to its complexity in nature, it is challenging to simulate the process of droplet impact on heated surfaces
Droplet boiling involving phase change. In the present work, a three-dimensional numerical simulation based on lattice Boltz­
Lattice Boltzmann method
mann method (LBM) is applied to analyze the dynamic and thermodynamic behaviors of droplet impact on
Bubble nucleation
heated surfaces. The process of vapor bubble nucleation, growth, coalescence, and even burst at the interface of
Bouncing
the droplet can be successfully captured, and four typical regimes are numerically reproduced, which include
film evaporation, nucleate boiling, transition boiling, and film boiling. Due to the idealized smooth surface
applied here, a constant contact angle (CCA) model is observed during the process of droplet evaporation, and
the time evolution of normalized droplet volume and spreading factor in the evaporation stage are consistent
with the sessile droplet evaporation. When the bubble nucleation occurs in the droplet, the droplet interface is
seriously distorted, and the heat transfer performance exhibits a strong dependence on vapor bubble dynamics.
In addition, according to the numerical results, three different rebound types, including burst rebound, partial
rebound, and complete rebound, are observed and analyzed in detail.

1. Introduction During the process of droplet impingement on heated surfaces, the


heat transfer performance and impact dynamics are severely affected by
Droplet impact on heated surfaces has attracted more and more surface temperature. With the increase in surface temperature, four
attention in many industrial applications, such as spray cooling of distinct regimes can be identified, namely film evaporation, nucleate
electronic devices [1], metal quenching [2], internal combustion en­ boiling, transition boiling, and film boiling, which are similar with that
gines [3], etc. When the surface temperature is higher than the liquid in pool boiling [7–10]. In the film evaporation regime, due to the low
saturation point, the droplet undergoes boiling, and due to the syner­ surface temperature, the heat flux is insufficient to trigger bubble
gistic effects of different heat transfer mechanisms, including sensible nucleation. Although the contact is maintained until the droplet
heating of liquid droplets, evaporation at the drop-vapor interface, and completely evaporates, there is no vapor bubble nucleation within the
nucleate boiling behavior at the droplet-solid interface, a good heat droplet, which leads to poor heat transfer performance. In the nucleate
transfer performance can be obtained [4]. However, when the surface boiling regime, vapor bubbles begin to nucleate, grow, coalesce, and
temperature increases to a critical value, due to the occurrence of violent even burst at the interface of the droplet. Due to the vapor bubble
boiling beneath the droplet, a layer of micro-scale vapor film is gener­ generation and enhanced convection, the heat transfer performance
ated, and the physical contact between the droplet and the heated sur­ becomes efficient [11]. Transition boiling spans the region between the
face disappears [5]. At this time, the heat transfer model is limited to critical heat flux point when the evaporation time of the droplet is
conduction and radiation across the vapor layer, and the heat transfer shortest to the Leidenfrost point when there is a vapor layer between the
performance is therefore significantly reduced [4,6]. The critical tem­ droplet and the surface due to the vigorous boiling [7]. However, due to
perature limit for such a phenomenon to occur is the Leidenfrost point, the small size of the droplet and the transient nature of the phenomenon,
and it is important to avoid the Leidenfrost phenomenon in many ap­ it is challenging to accurately determine these two points, which is not
plications requiring cooling or refrigeration [6]. conducive to the investigation of transition boiling [7,12]. In the film

* Corresponding author at: College of Aerospace Engineering, Nanjing University of Aeronautics and Astronautics, Nanjing, Jiangsu 210016, China.
E-mail address: zhaoam@nuaa.edu.cn (N. Zhao).

https://doi.org/10.1016/j.applthermaleng.2022.119360
Received 18 April 2022; Received in revised form 18 August 2022; Accepted 20 September 2022
Available online 26 September 2022
1359-4311/© 2022 Elsevier Ltd. All rights reserved.
Y. Xu et al. Applied Thermal Engineering 219 (2023) 119360

boiling regime, a thin vapor layer will rapidly form, and the droplet will terms of numerical stability and Galilean invariance [34,35]. More
rebound off the surface without any distortion leading to the substantial recently, the CLBM and MRT-LBM were unified in a general model
deterioration of heat transfer. Typically, the Leidenfrost point is taken as framework [26,36]. Combined with the pseudopotential model, the first
the onset of the film boiling regime and increases with increasing impact multiphase CLBM was proposed by Lycett-Brown and Luo in 2014 [35]
velocity or Weber number (We) [13,14]. and then was further developed to simulate high-density ratios and high
Despite these useful experimental findings, owing to its extreme Weber numbers [37,38]. To improve the implementation, a non-
complexity in nature, droplet boiling is still far from completely un­ orthogonal moment set and consistent forcing scheme were applied in
derstood. One critical limitation of experimental investigations is that it Ref. [39,40], and the thermal effects were considered in Ref. [41]. For
is difficult to visualize liquid–solid interfaces and volumetric space the boiling simulations, the multiphase CLBM has also been used to
within the droplet, as well as to obtain detailed quantitative information simulate two-dimensional (2D) forced-convection boiling [42], and
[12,15]. Fortunately, with the rapid development of computational three-dimensional (3D) pool boiling [10].
technology, numerical simulations of the boiling process can overcome In the present work, the CLBM coupled with pseudopotential
the visualization challenges and become an effective method to inves­ multiphase model is employed to simulate the density and velocity
tigate the boiling mechanism [16]. Li et al. [17] numerically investi­ fields, and the temperature field is solved by a 3D thermal MRT lattice
gated the self-propelled motion of Leidenfrost droplets on ratchet Boltzmann method which was proposed in our previous work [43]. This
surfaces via a hybrid thermal lattice Boltzmann model [9]. They thermal MRT lattice Boltzmann method can effectively eliminate the
demonstrated that this self-propelled motion of Leidenfrost droplets error terms in the recovered thermal equation and was well-validated
results from the asymmetry of the ratchets and the vapor flows below the through Laplace’s law, D2 law, and existing experimental data [43]. It
droplets. Wang et al. [18] adopted the volume of fluid method coupled should be noted that in our previous work, the slope or evaporation
with a vaporization model proposed by Hardt and Wondra [19] to constant was not considered in the D2 law cases, and recently a revised
investigate the interface oscillation of the droplet in the Leidenfrost analysis of this law can be found in Ref. [44]. In our previous study [43],
state, and they demonstrated that the interface oscillation has an the droplet impact on a heated surface with pillars was numerically
important impact on the flow in the vapor layer and the heat transfer investigated. However, the size of pillars is comparable to that of
between the droplet and heated surface. Xu et al. [20] investigated the droplets, which is inconsistent with many experiments and industrial
droplet evaporation behaviors on an over-heated surface by the Gong- applications. Therefore, in the present work, the surface structures are
Cheng phase-change lattice Boltzmann model [21], and the effects of ignored due to their insignificant dimension compared with droplets,
wall superheat and wettability were analyzed. It should be noted that, at and the impact and boiling characteristics of a droplet on a smooth
present, most of the previous numerical studies focus on the Leidenfrost heated surface are numerically investigated. To trigger the bubble
phenomenon or droplet evaporation, and there are few numerical nucleation, a small disturbance is added to the temperature of the bot­
studies on more complex phenomena, namely, droplet impact in tom of the substrate. The rest of the present paper is structured as fol­
nucleate boiling regime or transition boiling regime. lows. The lattice Boltzmann model and the specific simulation setup are
In most macroscopic multiphase methods [22–24], an initial vapor given in Section 2. Subsequently, the dynamic behaviors and heat
phase is required to act as an artificial input for the boiling simulations. transfer performance are analyzed in detail in Section 3. Finally, a brief
Therefore, these methods are limited to film boiling simulations and conclusion is provided in Section 4.
cannot provide any physical insights into bubble nucleation [10,16].
Different from conventional macroscopic numerical methods, as a 2. Model description
mesoscopic lattice Boltzmann method, the pseudopotential model in­
troduces the interaction force among the neighboring fluid particles to In this work, the flow field is solved by a three-dimensional cascaded
realize the phase separation [25], and the interface between different lattice Boltzmann method (CLBM) coupled with the pseudopotential
phases can arise, deform, and migrate automatically, without resorting model, the temperature field is solved by an improved thermal lattice
to any additional techniques to capture interfaces [26,27]. Furthermore, Boltzmann method that can recover the target equation without error
due to its computational efficiency, easy implementation of wetting terms, and the coupling between two fields is established via a non-ideal
boundaries, and ability to capture bubble nucleation, the pseudopo­ equation of state.
tential model makes it possible to indicate the boiling mechanism
[10,16,28–30] and hence is employed in the present work.
2.1. Cascaded lattice Boltzmann method
In lattice Boltzmann methods, the classical single-relaxation-time
(SRT) collision operator has been frequently used owing to its
The raw and central moments of the discrete distribution functions
simplicity, but it may suffer from instability under low-viscosity condi­
can be expressed as follows:
tions [31]. The multi-relaxation-time (MRT) operator, in which the
⃒ ⃒
collision is performed in the raw moment space, is an improved collision ⃒ ⃒ ( )n
kmnp =< fi ⃒emix eniy epiz >, kmnp =< fi ⃒(eix − ux )m eiy − uy (eiz − uz )p > (1)
operator and has been extensively shown to be capable of enhancing
numerical stability [32,33]. In 2006, the first cascaded operator was where m, n, and p are integers and ux , uy , anduz are the velocity
proposed by Geier et al. [34], in which the collision is performed in the components in the x, y, and z directions respectively. To improve the
central moment space. Compared with the SRT-LBM, the numerical implementation of the 3D CLBM, a non-orthogonal central moment set
stability can be improved by setting reasonable relaxation ratios in both for the D3Q19 lattice is adopted, which is defined as [10]:
the MRT-LBM and the cascaded LBM (CLBM) [33–35]. Some studies also
showed that the cascaded operator outperforms the MRT operator in

|Ti >= [k000 , k100 , k010 , k001 , k110 , k101 , k011 , k200 , k020 , k002 , k120 , k102 , k210 , k201 , k012 , k021 , k220 , k202 , k022 ]T , (2)

2
Y. Xu et al. Applied Thermal Engineering 219 (2023) 119360

where |> denotes a 19-dimensional column vector. According to Eq.


(1), the raw moment space can be transferred from a discrete velocity
space via a transformation matrix M, and the central moment space can
be transferred from the raw moment space by a shift matrix N [10]:

|Ti >= M|fi >, |T i >= N|Ti >= NM|fi > . (3)
The explicit formulations for M and N, and their inverses are given in
Appendix A. A block-diagonal relaxation matrix S is adopted here to
independently relax normal stress differences and the trace of the
pressure tensor, which can be given as [10,40]:
⎧ ⎛ ⎞ ⎫
⎨ s+ s− s− ⎬
S = diag s0 ,s1 ,s1 ,s1 ,sv ,sv ,sv , ⎝ s− s+ s− ⎠,s3 ,s3 ,s3 ,s3 ,s3 ,s3 ,s4 ,s4 ,s4 ,
⎩ ⎭
s− s− s+
(4)
where s+ = (s2b + 2sv )/3 and s− = (s2b − sv )/3. The kinematic viscosity
v and bulk viscosity ξ are given by v = (1/sv − 0.5)c2s and ξ =
2/3(1/s2b − 0.5)c2s , respectively. In the present work, s0 = s1 = 1.0,s2b =
0.3, and s3 = s4 = 0.5 are adopted. The kinematic viscosity is 0.0025 and
Fig. 1. Comparison of the evolution of the spreading factor using three
0.2 for liquid and vapor respectively, and the bulk viscosity is fixed at
different resolutions at ΔT = 0.97Tc.
0.6296. Thus, the post-collision central moments become:
* eq
|T i >= |T i > − S(|T i > − |T i >) + (I − S/2)δt|Ci > (5) action force between a solid and a fluid is introduced, calculated as
follows [47]:
where |Ci > are the forcing terms. ∑
For CLBM, the continuous central moments of the Maxwell- Fads (x) = − Gads ψ (x) wi ψ (x)s(x + ei )ei , (11)
Boltzmann distribution should be adopted [10]. Therefore, the equilib­
where Gads is the fluid–solid interaction strength, and s(x) is an in­
rium central moments based on the central moment set in Eq. (2) can be
dicator function. The gravitational force is given as follows:
given as follows:
[ ]T Fg = ρg, (12)
(6)
eq
|T i >= ρ, 0, 0, 0, 0, 0, 0, ρc2s , ρc2s , ρc2s , 0, 0, 0, 0, 0, 0, ρc4s , ρc4s , ρc4s ,
where g is the gravitational acceleration (-0.00000127). Thus, the
The corresponding forcing terms in the central moment space can be total force in Eq. (8) becomes: F = Fint + Fads + Fg .
written as follows: In order to ensure the thermodynamic consistency, the force terms in
[ ]T
|Ci >= 0, Fx , Fy , Fz , 0, 0, 0, 0, 0, 0, Fx c2s , Fx c2s , Fy c2s , Fz c2s , Fy c2s , Fz c2s , 0, 0, 0 . the CLBM could be modified into [10,48,49]:

(7) |Ci >


[ ]T
The streaming step is still performed in the velocity space, i.e., = 0, Fx , Fy , Fz , 0, 0, 0, η, η, η, Fx c2s , Fx c2s , Fy c2s , Fz c2s , Fy c2s , Fz c2s , 0, 0, 0 ,
fi (x +ei , t +1) = fi* (x, t) where fi* = M− 1 N− 1 |Ti >, and the density ρ and (13)
actual velocity u can be obtained by:
2θ|Fint |2
∑ ∑ where η = ψ 2 (se− 1 − 0.5)δt
, in which θ is a parameter to adjust the me­
ρ= fi , ρu = fi ei + F/2. (8)
i i chanical stability condition and is fixed at 0.12 in the present work.

2.2. Pseudopotential model 2.3. A 3D thermal MRT lattice Boltzmann method

In the pseudopotential multiphase lattice Boltzmann model, the If the viscous heat dissipation is neglected, the temperature equation
interaction force between fluids is introduced to achieve the phase for the liquid–vapor phase change process can be rewritten as follows
separation, which can be written as follows: [50]:

Fint = − Gψ (x, t) w(|ei |2 )ψ (x + ei δt, t)ei , (9) ∂t T + ∇⋅(uT) = ∇⋅(k∇T) + Sϕ , (14)

where G represents the strength of the fluid–fluid interaction force, where k is the thermal diffusivity, and Sϕ = ρ1cv ∇⋅(λ∇T) − ∇
w(|ei |2 ) are the weights, setting by w(0) = 1, w(1) = 16, and w(2) = 12 1
in 1 ∂PEOS
⋅(k∇T) +T[1 − ρcv ( ∂T )ρ ]∇⋅u where λ is the heat conductivity and cv is the
the D3Q19 lattice, and ψ is the effective density, which can be expressed specific heat.
√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
as ψ =
2(PEOS − ρc2s )
to incorporate the realistic equation of state (EOS). In MRT lattice Boltzmann method, Eq. (14) can be expressed as
G
[51–53]:
To make the term inside the square root positive, G is set as − 1
[10,45,46]. In the present work, the Peng-Robinson EOS is adopted, [
gi (x + ei δt, t + δt) = gi (x, t) − Λij gj (x, t) − geq
] ′
(15)
j (x, t) + δt Si (x, t),
which can be given by:
where gi is the temperature distribution function, and gieq is the
ρRTaρ2 α(T)ρ2 equilibrium temperature distribution function. In order to eliminate the
PEOS = − , (10)
1 − bρ 1 + 2bρ − b2 ρ2
error term ∇⋅(Tuu), gieq = wi T(1 + eci 2⋅u) is adopted. Λij = (M−T 1 ΛMT )ij is the
s
[ ( √̅̅̅̅̅̅̅̅̅̅ )]2
where α(T)= 1+(0.37464+1.54226ω − 0.26992ω2 ) 1− T/Tc collision matrix, where Λ is a diagonal matrix, and MT is an orthogonal
transformation matrix [54,55]. Si ′ (x, t) is the thermal source term in the
with ω = 0.344. The values of a, b, and R are 1/49, 2/21, and 1,
discrete velocity space, which corresponds to Sϕ in Eq. (14). The explicit
respectively.
expression for MT and M−T 1 can be given by:
Taking the wettability of the surface into consideration, the inter­

3
Y. Xu et al. Applied Thermal Engineering 219 (2023) 119360

⎡ ⎤
1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1
⎢ − 30 − 11 − 11 − 11 − 11 − 11 − 11 8 8 8 8 8 8 8 8 8 8 8 8 ⎥
⎢ ⎥
⎢ 12 − 4 − 4 − 4 − 4 − 4 − 4 1 1 1 1 1 1 1 1 1 1 1 1 ⎥
⎢ ⎥
⎢ 0 1 − 1 0 0 0 0 1 1 − 1 − 1 1 − 1 1 − 1 0 0 0 0 ⎥
⎢ ⎥
⎢ 0 − 4 4 0 0 0 0 1 1 − 1 − 1 1 − 1 1 − 1 0 0 0 0 ⎥
⎢ ⎥
⎢ 0 0 0 1 − 1 0 0 1 − 1 1 − 1 0 0 0 0 1 1 − 1 − 1⎥
⎢ ⎥
⎢ 0 0 0 − 4 4 0 0 1 − 1 1 − 1 0 0 0 0 1 1 − 1 − 1⎥
⎢ ⎥
⎢ 0 0 0 0 0 1 − 1 0 0 0 0 1 1 − 1 − 1 1 − 1 1 − 1⎥
⎢ ⎥
⎢ 0 0 0 0 0 − 4 4 0 0 0 0 1 1 − 1 − 1 1 − 1 1 − 1⎥
⎢ ⎥
MT = ⎢
⎢ 0 2 2 − 1 − 1 − 1 − 1 1 1 1 1 1 1 1 1 − 2 − 2 − 2 − 2⎥

⎢ 0 − 4 − 4 2 2 2 2 1 1 1 1 1 1 1 1 − 2 − 2 − 2 − 2⎥
⎢ ⎥
⎢ 0 0 0 1 1 − 1 − 1 1 1 1 1 − 1 − 1 − 1 − 1 0 0 0 0 ⎥
⎢ ⎥
⎢ 0 0 0 − 2 − 2 2 2 1 1 1 1 − 1 − 1 − 1 − 1 0 0 0 0 ⎥
⎢ ⎥
⎢ 0 0 0 0 0 0 0 1 − 1 − 1 1 0 0 0 0 0 0 0 0 ⎥
⎢ ⎥
⎢ 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 − 1 − 1 1 ⎥
⎢ ⎥
⎢ 0 0 0 0 0 0 0 0 0 0 0 1 − 1 − 1 1 0 0 0 0 ⎥
⎢ ⎥
⎢ 0 0 0 0 0 0 0 1 1 − 1 − 1 − 1 1 − 1 1 0 0 0 0 ⎥
⎢ ⎥
⎣ 0 0 0 0 0 0 0 − 1 1 − 1 1 0 0 0 0 1 1 − 1 − 1⎦
0 0 0 0 0 0 0 0 0 0 0 1 1 − 1 − 1 − 1 1 − 1 1

and.

⎡ ⎤
⎢ 1 5 1 ⎥
⎢ − 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 ⎥
⎢ 19 399 21 ⎥
⎢ ⎥
⎢ 1 11 1 1 1 1 1 ⎥
⎢ − − − 0 0 0 0 − 0 0 0 0 0 0 0 0 ⎥
⎢ 19 2394 63 10 10 18 18 ⎥
⎢ ⎥
⎢ ⎥
⎢ 1 11 1 1 1 1 1 ⎥
⎢ − − − 0 0 0 0 − 0 0 0 0 0 0 0 0 ⎥
⎢ 19 2394 63 10 10 18 18 ⎥
⎢ ⎥
⎢ ⎥
⎢ 1 11 1 1 1 1 1 1 1 ⎥
⎢ − − 0 0 − 0 0 − − 0 0 0 0 0 0 ⎥
⎢ 19 2394 63 10 10 36 36 12 12 ⎥
⎢ ⎥
⎢ 1 11 1 1 1 1 1 1 1 ⎥
⎢ − − 0 0 − 0 0 − − 0 0 0 0 0 0 ⎥
⎢ ⎥
⎢ 19 2394 63 10 10 36 36 12 12 ⎥
⎢ ⎥
⎢ 1 11 1 1 1 1 1 1 1 ⎥
⎢ − − 0 0 0 0 − − − 0 0 0 0 0 0 ⎥
⎢ 19 2394 63 10 10 36 36 12 12 ⎥
⎢ ⎥
⎢ ⎥
⎢ 1 11 1 1 1 1 1 1 1 ⎥
⎢ − − 0 0 0 0 − − − 0 0 0 0 0 0 ⎥
⎢ 19 2394 63 10 10 36 36 12 12 ⎥
⎢ ⎥
⎢ 1 4 1 1 1 1 1 1 1 1 1 1 1 1 ⎥
⎢ 0 0 0 0 − 0 ⎥
⎢ ⎥
⎢ 19 1197 252 10 40 10 40 36 72 12 24 4 8 8 ⎥
⎢ ⎥
⎢ 1 4 1 1 1 1 1 1 1 1 1 1 1 1 ⎥
⎢ − − 0 0 − 0 0 0 ⎥
⎢ 19 1197 252 10 40 10 40 36 72 12 24 4 8 8 ⎥
⎢ ⎥
⎢ ⎥
⎢ 1 4 1 1 1 1 1 1 1 1 1 1 1 1 ⎥
MT− 1 =⎢ − − 0 0 − 0 0 − − 0 ⎥
⎢ 19 1197 252 10 40 10 40 36 72 12 24 4 8 8 ⎥
⎢ ⎥
⎢ 1 4 1 1 1 1 1 1 1 1 1 1 1 1 ⎥
⎢ − − − − 0 0 0 0 − 0 ⎥
⎢ ⎥
⎢ 19 1197 252 10 40 10 40 36 72 12 24 4 8 8 ⎥
⎢ ⎥
⎢ 1 4 1 1 1 1 1 1 1 1 1 1 1 1 ⎥
⎢ 0 0 − − 0 0 − 0 ⎥
⎢ 19 1197 252 10 40 10 40 36 72 12 24 4 8 8 ⎥
⎢ ⎥
⎢ ⎥
⎢ 1 4 1 1 1 1 1 1 1 1 1 1 1 1 ⎥
⎢ − − 0 0 − − 0 0 − 0 ⎥
⎢ 19 1197 252 10 40 10 40 36 72 12 24 4 8 8 ⎥
⎢ ⎥
⎢ 1 4 1 1 1 1 1 1 1 1 1 1 1 1⎥
⎢ 0 0 − − − − 0 0 − − 0 − ⎥
⎢ ⎥
⎢ 19 1197 252 10 40 10 40 36 72 12 24 4 8 8⎥
⎢ ⎥
⎢ 1 4 1 1 1 1 1 1 1 1 1 1 1 1⎥
⎢ − − 0 0 − − − − 0 0 0 − ⎥
⎢ 19 1197 252 10 40 10 40 36 72 12 24 4 8 8⎥
⎢ ⎥
⎢ ⎥
⎢ 1 4 1 1 1 1 1 1 1 1 1 1⎥
⎢ 0 0 − − 0 0 0 0 0 − ⎥
⎢ 19 1197 252 10 40 10 40 18 36 4 8 8⎥
⎢ ⎥
⎢ 1 4 1 1 1 1 1 1 1 1 1 1 ⎥
⎢ 0 0 − − − − 0 0 0 − 0 0 ⎥
⎢ ⎥
⎢ 19 1197 252 10 40 10 40 18 36 4 8 8 ⎥
⎢ ⎥
⎢ 1 4 1 1 1 1 1 1 1 1 1 1⎥
⎢ 0 0 − − − − 0 0 0 − 0 0 − − ⎥
⎢ 19 1197 252 10 40 10 40 18 36 4 8 8⎥
⎢ ⎥
⎢ ⎥
⎢ 1 4 1 1 1 1 1 1 1 1 1 1 ⎥
⎣ 0 0 − − − − − − 0 0 0 0 0 − ⎦
19 1197 252 10 40 10 40 18 36 4 8 8

4
Y. Xu et al. Applied Thermal Engineering 219 (2023) 119360

Through some algebraic operations, the temperature distribution


evolution Eq. (15) can be divided into the following two steps:

m*i (x, t) = mi (x, t) − Λi (mi (x, t) − meq


i (x, t) ) + δtSi (x, t), (16)

gi (x + ei δt, t + δt) = g*i (x, t), (17)

where m = MT g, m = MT g , Λ= [1,0.8,0.8,1,1,1,1,1,1,1,1,1,1,1,1,
eq eq

1,1,1,1] and S is the source term in the moment space, which can be chosen
[ ]T
as S = sϕ + 0.5δt ∂t sϕ , 0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0 . It
should be noted that the additional term 0.5δt ∂t sϕ is used to eliminate the
discrete effects of source terms [51].
In addition, some correction terms should be applied to m3 , m5 ,
and m7 to eliminate the error term ∂t (Tu) [43]. Therefore, the collision
step for i = 3, 5, 7 becomes:
( s3 )
m*3,new = m*3 + δt 1 − ∂t0 (Tux ),
2
( s ) ( )
m*5,new = m*5 + δt 1 − (18)
5
∂t0 Tuy ,
2
( s )
m*7,new = m*7 + δt 1 −
7
∂t0 (Tuz ).
2
According to the Chapman-Enskog analysis in Ref. [43], the
macroscopic temperature equation can be obtained:
∂t T + ∇⋅(uT) = ∇⋅(k∇T) + Sϕ , (19)
( )
1
k = c2s − 0.5 , i = 3, 5, 7
si

2.4. Simulation setup

To facilitate the description of fluid conditions, dimensionless


numbers are introduced here:

ρl U 20 D0 μl ν ρgD20
We = ̅, Pr = , Bo =
, Oh = √̅̅̅̅̅̅̅̅̅̅̅ , (20)
σ ρl σ D0 α σ

where D0 is the initial droplet diameter, U0 is the initial velocity of


the droplet (-0.0345), and σ is the surface tension (0.0914). Unless
otherwise stated, Ohnesorge number (Oh), Bond number (Bo), contact
angle, and We are set as 0.0023, 1.067, 73.6 , and 10, respectively, and

Fig. 2. (a) Time evolution of the spreading factor and transient heat flux. (b) the time t used in the present work is nondimensionalized by t0 , which
*
√̅̅̅̅̅̅3̅
Time evolution of the normalized droplet volume. Insert figure: droplet profiles ρl R0
corresponding to t* = 8.95, 12.03, 15.12, 18.21, 21.29, and 24.38. (ΔT =
is given as:t0 = σ . In addition, ΔT = Tb − Ts , where Tb is the tem­
0.01Tc). perature of the substrate bottom, is introduced to describe the wall su­
perheat. The code used for all cases in the present work is written in the
programming language c++.
For the grid independency checking, three sets of grids are adopted

Fig. 3. Sequential images of droplet impingement on the heated surface at ΔT = 0.03Tc. The droplet interface and solid wall are marked by green and gray,
respectively.

5
Y. Xu et al. Applied Thermal Engineering 219 (2023) 119360

fluid, the subscript s and f represent the solid and fluid, respectively, and
α represents the opposite direction of α. In the present work, the satu­
ration temperature is set to be Ts = 0.75Tc , and the corresponding liquid
and vapor densities are ρl = 7.6788 and ρv = 0.1086, respectively. The
Prandtl number and specific heat for liquid are Prl = 6.1 and Cvl = 4.2,
while for vapor Prv = 1.0 and Cvv = 2.1. The density and specific heat of
solid are set as ρs = 28.66 and Cvs = 0.8. In fact, at the beginning of the
heating stage, a short time is required to transfer the heat from the
bottom to the top of the substrate [57]. Therefore, it is more reasonable
to place the droplet above the substrate for a period without gravity
instead of directly releasing the droplet. However, during such a process,
the droplet may slightly evaporate and rise due to the heat transfer.
Considering these comprehensive effects, in the present work, the
droplet with a radius of 50 is first placed just above the substrate for
1000 time steps. Then, it is released with a falling velocity U0 . It should
be noted that all the parameters used in the present work are all based on
the lattice units, and the relationship between lattice units and physical
units are provided in Appendix B.

Fig. 4. Time evolution of the transient heat flux. 3. Results and discussion

in which the grid sizes are 121 × 121 × 241, 151 × 151 × 301, and 3.1. Droplet impact behaviors
181 × 181 × 361 lu3 , respectively. In Fig. 1, it can be found that under
the same configuration, the time evolution of the spreading factor is In this section, the effects of wall superheat on the droplet dynamics
almost same when the grid size is more than 151 × 151 × 301 lu3 which and thermodynamics are investigated in detail. For the cases of low
means the grid size of 151 × 151 × 301 lu3 is sufficient. Therefore, superheat such as ΔT = 0.01Tc, due to the low heat flux, no nucleation
considering the compromise between accuracy and computational cost, occurs within the droplet, and the whole process of droplet impingement
on heated surfaces can be divided into two stages based on the droplet
the grid size of 151 × 151 × 301 lu3 is adopted in the present work.
dynamics and thermodynamics, namely, the initial stage referring to the
As previous analyzed, 3D computational domain with the di­
droplet impingement phase and evaporation stage referring to the
mensions of 151 × 151 × 301 lu3 is employed in the present work. A
evaporation phase. The transition from the initial stage to the evapo­
droplet with R0 = 50 lu is initially placed above the solid substrate. The
ration stage occurs approximately at t* = 8.5. It should be noted that the
height of the substrate is fixed at 10 lu. The no-slip boundary condition is
present work does not consider the contact angle hysteresis. Therefore,
employed for the fluid–solid boundary, and the symmetry boundary
as shown in the insert figure of Fig. 2(b), a constant contact angle (CCA)
conditions are applied on the sides of the domain. The top boundary is
model in which the contact angle remains almost constant during the
set as a free outflow boundary, and the bottom of the substrate is
droplet evaporation can be observed, while the fundamental and final
imposed at a constant temperature. Similar to previous studies [9,10], to
stages referring to the pinning and shrinking of contact line cannot be
trigger the vapor bubble nucleation, small temperature fluctuations are
observed here [7,58]. Fig. 2(a) shows the time evolution of the
added to the temperature of the substrate bottom. A fluid–solid conju­
spreading factor and transient heat flux, which are defined as ξ = Dl /D0
gate thermal boundary scheme is applied at the fluid–solid interface to ∫L ∫L [ ( )]
ensure the continuity of heat transfer across the fluid–solid boundary and q = L12 0 0 − λ ∂∂Tz z=0 dxdy, respectively [43,59]. It can be found
[56]: that in the initial stage, the transient heat flux and spreading factor reach
the maximum at almost the same time, and then both decrease with
( ) 1 − γ *( ) 2γ *
gα xf , t + δt = g xf , t + g (xs , t) time. Therefore, it can be inferred that the heat flux increases with the
1+γ α 1+γ α length of the three-phase contact line. While, in the evaporation stage,
(21)
gα (xs , t + δt ) =
γ− 1 *
g (xs , t) +
2 *(
g xf , t
) the evolution of normalized droplet volume and spreading factor are
1+γ α 1+γ α both consistent with the sessile droplet evaporation reported in the
where γ = (ρcv )s /(ρcv )f is the thermal mass ratio of the solid and previous study [58], which means the effects of droplet evaporation

Fig. 5. Sequential images of droplet impingement on the heated surface at ΔT = 0.15Tc.

6
Y. Xu et al. Applied Thermal Engineering 219 (2023) 119360

Fig. 6. (a) Sequential images of droplet impingement on the heated surface at ΔT = 0.25Tc. (b) The temperature field on the midsection of the x direction at t * =
2.16.

a heated surface at ΔT = 0.03Tc. Different from the cases of ΔT =


0.01Tc, the vapor bubble nucleates in the center of the droplet (as
shown in Fig. 3 t * = 8.18), and due to the evaporation at the interface of
the droplet, the three-phase contact line moves inward driving the
coalescence of the growing vapor bubble, which finally leads to the
bubble burst (as shown in Fig. 3 t * = 11.26). The evaporation time,
which is defined as the interval from the moment when the droplet
contacts the surface to the droplet completely evaporates into vapor, is
reduced by 22 % compared with ΔT = 0.01Tc. In Fig. 4, the time evo­
lution of heat flux for ΔT = 0.01Tc and ΔT = 0.03Tc are presented. It
can be found that although the transient heat flux is very close when t * is
about 7, with the vapor bubbles nucleating in the droplet, the transient
heat flux of ΔT = 0.03Tc is then much greater than that of ΔT = 0.01Tc.
Furthermore, for ΔT = 0.03Tc, the vapor bubble bursts at the interface
of the droplet (as shown in Fig. 3 t* = 11.26 13.58), and meanwhile, the
transient heat flux decreases (as shown in Fig. 4), which also convinc­
ingly demonstrates the dependence of the heat transfer on the bubble
dynamics. It should be noted that although the nucleation process is
initiated within the droplet, the contact between the liquid and surface is
Fig. 7. Time evolution of the transient heat flux. Insert figure: the corre­ maintained until all liquid evaporates. Therefore, in some previous
sponding droplet dynamics at ΔT = 0.25Tc . studies, this case is also named as contact boiling [4,43,60].
Liang and Mudawar [7] defined the nucleate boiling regime as the
become dominant in this stage. In addition, as shown in Fig. 2 (a), the regime that extends from the onset of boiling to the critical heat flux
reduction of droplet spreading factors shows a slowing speed, which (CHF) point when the lifetime of the droplet is shortest. When ΔT =
implies the accelerated motion of the three-phase contact line. Since no 0.01Tc and ΔT = 0.02Tc , no obvious vapor bubble nucleation can be
vapor bubble nucleation occurs and the droplet evaporates continu­ found within the droplet during the whole impact process. However, as
ously, this phenomenon can be classified into film evaporation regime shown in Fig. 3, when ΔT = 0.03Tc , the vapor bubble nucleation,
following the definition in Ref. [7]. growth, coalescence, and burst can be observed. Thus, following the
With an increase in wall superheat, the vapor bubble may nucleate, definition of Liang and Mudawar [7], the cases of ΔT = 0.03Tc can be
grow, and coalesce within the droplet. These complex vapor bubble taken as the incipience of the nucleate boiling regime in the present
behaviors distort the droplet interface and exert a great influence on work. In fact, due to the randomness of the disturbance imposed on the
heat transfer. Fig. 3 shows the whole process of droplet impingement on temperature of the substrate bottom, it is difficult to accurately

7
Y. Xu et al. Applied Thermal Engineering 219 (2023) 119360

Fig. 8. Sequential images of droplet impingement on the heated surface at (a) ΔT = 0.67Tc, (b) ΔT = 0.80Tc.

Fig. 9. Time evolution of the spreading factor. Fig. 10. Time evolution of the underneath pressure at ΔT = 0.80Tc and ΔT =
0.67Tc.

determine the CHF point, which is also the major challenge in the ex­
droplet surface is severely distorted by the evaporation and nucleation
periments [7]. However, because the evaporation time of droplets at the
process. Unlike pool boiling, the vapor bubbles within the droplet do not
CHF point is the least, with the increase in surface temperature, the
detach from the heated surface but reside on the heated substrate [7,15].
evaporation time of droplets will first decrease and then increase near
In addition, due to the fact that the droplet is much thinner compared to
the CHF point, which makes it possible to determine the approximate
bulk liquid, the vapor bubbles may get ruptured by reaching the free
position of the CHF point. According to the numerical results, when
surface of the droplet. Under the action of the vapor bubble burst, the
ΔT = 0.1Tc , 0.15Tc , and 0.2Tc , the evaporation time of the droplets are
central part of the droplet rebounds from the surface (as shown in Fig. 6
t * = 16.54, 14.50, and 15.55, respectively. Therefore, according to the
t * = 2.16 3.39). Then, under the action of gravity, the droplet will
previous analysis, it can be inferred that the CHF occurs near ΔT =
0.15Tc . Fig. 5 shows the whole process of droplet impingement on a contact the surface again. The temperature field at t * = 2.16 is presented
heated surface at ΔT = 0.15Tc . The contact is maintained until the in Fig. 6 (b), in which the black lines represent the droplet and the
droplet completely evaporates, and due to the intense boiling heat substrate, and the dimensionless temperature is defined as T* = T/Tc. It
transfer, the droplet interface is seriously distorted by the vapor bubble can be clearly observed that the temperature is much smaller in the
behaviors, which greatly reduces the evaporation time of the droplet. three-phase contact region, which indicates a better heat transfer in this
As the wall superheat increases further, the outcomes of droplet region. Different from film boiling, at this time, the wall superheat is
impingement on the heated surface will present lift-off behavior. Fig. 6 insufficient to generate a vapor layer with sufficient strength to make the
(a) shows the process of droplet impingement on the heated surface at droplet rebound. Benefiting from the vapor bubble burst, the upward
ΔT = 0.25Tc . It can be observed that with the increase of wall superheat, momentum of the droplet is increased, and the droplet consequently
the time for vapor bubble nucleation is greatly advanced, and the rebounds from the heated surface. Since this droplet rebound is

8
Y. Xu et al. Applied Thermal Engineering 219 (2023) 119360

Fig. 11. Sequential images of droplet impingement on the heated surface at ΔT = 0.75Tc.

increases with the increase of wall superheat, and the transient heat flux
is seriously affected by bubble dynamics. To be more specific, the
transient heat flux may decrease with the decrease of the length of the
three-phase contact line. Even when the droplet is completely out of
contact with the surface, namely, the length of the three-phase contact
line becomes zero (as shown in the insert figure in Fig. 7), the heat flux is
reduced by two orders of magnitude.
When the wall superheat is relatively high, the droplet may rebound
away from the surface due to the vigorous boiling at the droplet bottom,
without vapor bubble generation within the droplet. However, due to
the difference in wall superheat, this rebound behavior of droplets can
be further divided into two ways, which are defined as partial rebound
and complete rebound, depending on whether the adhesion phenome­
non occurs. Although the droplet dynamics are different in those two
rebound ways, the gain in the upward momentum is attributed to the
same reason, namely, the vigorous boiling beneath the droplet. The
processes of those two rebound ways are shown in Fig. 8. It can be found
that after impacting the heated surface, the droplet spreads on the
Fig. 12. Instantaneous velocity fields of a droplet impacting on a heated sur­ heated surface until it reaches the maximum spreading diameter, and
face at We = 10, ΔT = 0.75Tc, and t* = 19.29. then begins to shrink under the action of the surface tension. It should be
pointed out that during the spreading and shrinking of the droplet, a
vapor layer always exists and there is no contact between the droplet
characterized by the vapor bubble burst, it is called burst rebound. and the heated surface. Different from the complete rebound, the con­
Fig. 7 shows the time evolution of the transient heat flux at ΔT = tact between the bottom of the droplet and the top of the substrate oc­
0.1Tc , 0.15Tc , and 0.25Tc . For the cases considered here, the vapor curs at t * = 2.62 in the partial rebound. A similar phenomenon was also
bubbles nucleate in the droplet, and thus the droplet interface is seri­ observed in the previous experiments [60]. The time evolution of the
ously distorted which leads the transient heat flux to fluctuate violently. spreading factor ξ is illustrated in Fig. 9. It can be found that the droplet
From Fig. 7, it can be inferred that the peak value of transient heat flux spreading factor is almost the same at ΔT = 0.67Tc and ΔT = 0.80Tc

9
Y. Xu et al. Applied Thermal Engineering 219 (2023) 119360

spreading factor. To be more quantitative, the time evolution of the


pressure underneath the droplet at ΔT = 0.67Tc and ΔT = 0.80Tc is
plotted in Fig. 10. It can be found that when t * < 2.0, the pressure at the
bottom of the droplet is almost the same, while when t * is about 2.0, the
underneath pressure becomes larger than ΔT = 0.67Tc at ΔT = 0.80Tc,
which results in the different dynamic behavior as shown in Fig. 8. Then,
with the droplet bouncing off, the underneath pressure decreases
sharply at ΔT = 0.80Tc, while due to the adhesion at the droplet bot­
tom, the pressure at ΔT = 0.67Tc does not show a significant decrease
until t * is about 2.93 when the droplet rebounds away as shown in Fig. 8
(a). In addition, in the present work, the numerical results show that
ΔT = 0.75Tc is a critical point when the transition of rebound way
occurs.
As analyzed before, a high wall superheat is required for the droplet
to rebound completely. However, during the whole process of droplet
impact, the droplet rebound way may change. A typical process of the
transition of rebound types is shown in Fig. 11. It can be found that the
droplet leaves the heated surface in a complete rebound way for the first
time (as shown in Fig. 11 t * = 0 8.64), while the second rebound is in a
Fig. 13. Time evolution of the droplet height at ΔT = 0.75Tc.
partial rebound way. These different impact results can be attributed to
the following points. Firstly, after the first impact, the temperature of the
droplet increases considerably, which will reduce the boiling intensity at
except in the final stage when the droplet bounces off the surface, which the bottom of the droplet [61]; Secondly, in the second impact, although
may be attributed to the presence of a high-pressure vapor layer be­ the impact velocity of the droplet largely depends on the maximum
tween the droplet and the heated surface. For ΔT = 0.67Tc , due to the rebound height of the droplet, due to the buffer effect of the vapor layer,
insufficient superheat of the surface, the generated high-pressure vapor it will be much less than that of the first impact [62,63]; Finally, in the
layer is not stable enough to overcome the downward movement of the second impact, the shape of the droplet seems to be more similar to an
droplet bottom caused by the propagation of the interface wave. ellipsoid than a standard sphere at the exact moment of droplet impact
Therefore, as shown in Fig. 8 (a) t * = 2.62, the bottom of the droplet will on the heated surface. Therefore, the outcomes of the second droplet
adhere to the solid surface, which then leads to the deviation in the impingement become different. In the third rebound, the deformation of

Fig. 14. Sequential images of droplet impingement on the heated surface at ΔT = 1.10Tc.

10
Y. Xu et al. Applied Thermal Engineering 219 (2023) 119360

Fig. 15. Time evolution of the spreading factor at ΔT = 0.935Tc and ΔT =


1.10Tc.

Fig. 17. The variation of the maximum spreading factor ξmax with the (a)
nondimensional wall superheat, and (b) We.

the droplet becomes intense. Fig. 12 shows the instantaneous velocity


fields of a droplet impacting on a heated surface at t* = 19.29. It can be
found that the bottom of the droplet is pushed outward by the growing
vapor bubble, while the upper part keeps the downward movement
trend. Therefore, with the further growth of the vapor bubble, the bot­
tom of the droplet will leave the heated surface, and due to the incon­
sistent velocity distribution inside the droplet, the original pyramid
shape of the droplet at t* = 19.60 evolves to an inverted pyramid shape
at t* = 19.60. Then, after several successive impacts, the droplet ach­
ieves the Leidenfrost state, and the rebound way changes back to the
complete rebound. In fact, during the continuous rebound of the droplet,
the impact velocity of droplets is greatly reduced, and the size of
droplets continues to decrease due to evaporation, which leads to the
decreases of We. As previous studies pointed out [7,13,14], the Lei­
denfrost point decreases with the decrease of We, which explains why
the droplets eventually achieve the Leidenfrost state. Numerical results
also show that even if the first rebound of droplets is in a partial rebound
way, with continuous evaporation, the size and impact velocity of the
droplet will both be greatly reduced, and the droplets still will achieve
the Leidenfrost state. In addition, when ΔT ≥ 0.935Tc , a thin layer of
Fig. 16. Time evolution of the spreading factor at various (a) wall superheat at
vapor between the droplet and the surface will always exist during the
We = 10, and (b) We at ΔT = 1.10Tc.
whole process of droplet impingement. Therefore, according to the

11
Y. Xu et al. Applied Thermal Engineering 219 (2023) 119360

definition in the previous studies [4,60], ΔT = 0.935Tc can be taken as


the Leidenfrost point, and so far, the transition boundaries between
regimes are all identified.
In the recent work [64], a trampolining behavior of the droplet was
observed where the droplet bounces to higher heights with each
consecutive contact with the heated surface. Due to the visualization
challenges, it is impossible to directly observe the vapor bubble dy­
namics in the droplet. Therefore, Agrawal and Dash [64] proposed a
hypothesis that the upward rebound momentum of the droplet that re­
sults in trampolining is caused by the generation and escape of bubbles
at the liquid-substrate interface prior to droplet takeoff, based on the
observation of a flatting of the droplet base. However, one of the ad­
vantages of numerical simulation is the observation of the interaction
between droplets, vapor bubbles, and heated surfaces. The value of
droplet height which is nondimensionalized by R0 at different time is
plotted in Fig. 13. According to the insert figures, it can be found that
due to the intense deformation of the droplet, the marked part in Fig. 13
represents only one rebound process, although there are two peaks. It
should be noted that in our numerical simulations, the droplets impact
the heated surface with We = 10, while We < 1 in the experiments [64].
Therefore, during the initial serval rebounds, the effects of droplet
impact velocity cannot be ignored, which leads to no increase in the
rebound height of the droplet in the second and third rebound, although
vapor bubble forms and escapes. Combined with Fig. 13 and Fig. 11, it
can be found that the vapor bubble formation and escape occur during
t* = 22.06 22.84, and the rebound height of the droplet increases
significantly. Furthermore, after t* = 26.70, no bubble formation and
escape can be observed, and the rebound height of the droplet contin­
uously decreases. Therefore, it can be indicated that the vapor bubble
formation and escape can increase the rebound height, which confirms
the previous hypothesis [64].
A typical impingement process for the droplet in the Leidenfrost state
(ΔT = 1.10Tc) is shown in Fig. 14. During the whole process of droplet
impingement, no vapor bubble formation or adhesion occurs. Combined
with the insert figures, it can be found that a vapor layer exists between
the droplet and the solid surface even when the droplet has been
significantly deformed. In addition, due to the buffer effect of this vapor
layer [62,63], the movement and deformation of the droplet in subse­
quent impingements tend to be gentle. Fig. 15 illustrates the time evo­
lution of spreading factor at ΔT = 0.935Tc and ΔT = 1.10Tc. When
ΔT = 0.935Tc and ΔT = 1.10Tc, the difference of droplet spreading
Fig. 18. Residence time variation with the (a) nondimensional superheat
factor is very small. Therefore, it can be inferred that for the droplet in
ΔT * = ΔT/Tc , and (b) We.
the Leidenfrost state, the influence of wall superheat on the droplet
spreading factor can be almost ignored.

3.2. Droplet spreading factor

The droplet spreading factor is an important parameter in the study


of boiling heat transfer, because it is a direct measure of the heat transfer
area [4]. Similar to previous studies [4,63,65], the discussions of the
spreading factor and the following residence time and maximum
spreading time are all confined to the first impact in the complete
rebound regime. Fig. 16 (a) illustrates the temporal evolution of the
droplet spreading factor at different wall superheat. It can be found that
the spreading factor is insensitive to wall superheat and the evolution of
the spreading factor is basically the same under different wall superheat.
This can be attributed to the presence of a vapor layer between the
droplet and the heated surface. It should be noted that because a higher
wall superheat is required for the droplet to be levitated by the vapor
generated by the vigorous boiling with the increase of droplet impact
velocity [4,7,66], to ensure that the droplet can completely rebound
from the heated surface, the wall superheat here is taken as 1.10Tc in
Fig. 16 (b) which is much larger than the critical values 0.75Tc at We =
Fig. 19. The variation of the maximum spreading time with the non-
dimensional superheat ΔT * = ΔT/Tc . 10. As shown in Fig. 16 (b), with the increase of We, the spreading factor
increases greatly due to higher impact energy, and the time when ξ

12
Y. Xu et al. Applied Thermal Engineering 219 (2023) 119360

reaches its maximum is slightly advanced. √̅̅̅ √̅̅̅̅̅̅̅3̅


In addition, the dependence of the maximum spreading factor on the π ρR
tξm = √̅̅̅ (26)
wall superheat and We is also investigated. As shown in Fig. 17 (a), the 2 2 σ
maximum spreading factor is almost the same with the increase of wall Fig. 19 shows the variation of maximum spreading time with
superheat, which is consistent with experiments [63,65]. This phe­ nondimensional wall superheat. It can be found the maximum spreading
nomenon can be attributed to the presence of a thin layer of vapor be­ time is insensitive to the wall superheat which is consistent with the
tween the droplet and the heated surface. Fig. 17 (b) illustrates the previous study [63]. However, from the numerical data, the prefactor in
variation of the maximum spreading factor with We. The wall superheat √̅̅̅ √̅̅̅
Eq. (26) is deduced as 0.688 which is slightly larger than π/2 2 in
is taken as 1.10Tc , too. In previous studies [8,65,67,68], the maximum
Ref. [63].
spread factor was widely predicted by:
ξm = Cwea , (22) 4. Conclusions

where C and a can be obtained through fitting experimental or nu­


In the present work, a lattice Boltzmann model is employed to
merical data. Taking the fact that ξm tends to be one at small We into
investigate the droplet dynamics and heat transfer performance during
consideration, Castanet et al. [69] modified Eq. (22) into a form of:
the process of droplet impingement on the heated surface. Four different
ξm = 1 + CWe0.5 . (23) regimes can be numerically reproduced, including film evaporation,
nucleate boiling, transition boiling, and film boiling. Different from the
They also validated Eq. (23) based on the data of Tran et al. [14] in experimental research, the vapor bubble nucleation, growth, coales­
which ξm We0.4 was demonstrated. As shown in Fig. 17, our numerical cence, and even burst near the droplet interface can all be captured,
results agree well with the fit curve given in ξm = 0.17*We0.5 + 1.0, which can help us confirm the hypothesis made in previous experiments
which is consistent with the studies of Castanet et al. [69]. [64]. In addition, according to the numerical results, we can get the
following conclusions:
3.3. Residence time and maximum spreading time
(1) When the wall superheat is low, the heat flux is insufficient to
The droplet residence time tr is defined as the time interval from its trigger bubble nucleation, and the process of droplet impact can
first impact to the first rebound from the hearted surface, which controls be divided into the initial stage and evaporation stage referring to
the amount of thermal energy transported from the surface to the the droplet impingement phase and evaporation phase, respec­
droplet [4,7]. Therefore, the droplet residence time is an important tively. Due to the idealized smooth surface adopted here, a CCA
parameter to analyze the dynamic and thermodynamic behavior of the evaporation model is observed in which the apparent contact
droplet impact on heated surfaces and has been widely studied. In some angle is almost constant. In the evaporation stage, the evolution
experimental investigations [63,70], it has been reported that the resi­ of normalized droplet volume and spreading factor are both
dence time can be approximated by the period of a freely oscillating consistent with the sessile droplet evaporation, and the reduction
droplet given in form of: of droplet spreading factors shows a slowing speed, suggesting
√̅̅̅̅̅̅̅̅̅ the accelerated motion of the three-phase contact line.
π ρl R30 (2) The bubble nucleation can effectively enhance heat transfer and
tr = √̅̅̅ (24)
2 σ shorten the evaporation time of droplets, although it makes the
where D0 , ρl , and σ represents the droplet diameter, liquid density, interface of the droplet distorted. In the region of bubble nucle­
and surface tension respectively. However, some researchers have also ation in the droplet, the peak value of heat flux increases with the
pointed out that the Eq. (24) does not agree well with their experimental increase of wall superheat and the transient heat flux shows a
data, and to improve the agreement with the data, a more general form strong dependence on bubble dynamics.
is recommended [4]: (3) With the increase of wall superheat, three different rebound
√̅̅̅̅̅̅̅̅̅ types, including burst rebound, partial rebound, and complete
ρl R30 rebound, are observed in the numerical results. The burst
tr = C (25) rebound is caused by the bubble burst near the periphery of
σ
droplets, while the other two rebound types are caused by
Fig. 18 illustrates the effects of the wall superheat and We on the vigorous boiling beneath the droplet. The main difference be­
√̅̅̅̅̅̅3̅
nondimensional residence time which is nondimensionalized by
ρl R0 tween partial rebound and complete rebound is whether the
σ .
adhesion phenomenon occurs. During the whole process of
Similarly, in order to ensure that the droplet is always in the complete
droplet impact, the droplet rebound type may change, and for the
rebound regime, in Fig. 18 (b), the wall superheat is set to 1.1 0Tc . From
second and subsequent impact, the droplet may exhibit different
Fig. 18, it can be inferred that the residence time of droplets is inde­
impact behaviors due to its preheating, low impact velocity, and
pendent of the wall superheat and We. In addition, our results show
√̅̅̅ different shape. With the decrease of actual impact velocity and
good agreement with the theoretical value π/ 2. continuous evaporation, the droplet will eventually achieve the
Recently, based on the experimental results, Illis et al. [63] proposed Leidenfrost state after several impacts.
a theoretical model for predicting the maximum spreading time when (4) For the complete rebound, the maximum spreading factor is
the droplet spreads to its maximum spreading radius: insensitive to the wall superheat while it exhibits a strong
dependence on the We in the formulation of ξm = 0.17*We0.5 +
1.0. While for the time scale, the droplet residence time is

13
Y. Xu et al. Applied Thermal Engineering 219 (2023) 119360

insensitive to both the wall superheat and We and shows good Data availability
√̅̅̅̅̅̅
agreement with the theoretical value √π̅̅2 ρRσ . Furthermore, the
3

No data was used for the research described in the article.


maximum spreading time is also insensitive to the wall superheat.
Acknowledgments
Declaration of Competing Interest
This study was supported by the National Natural Science Founda­
The authors declare that they have no known competing financial tion of China (NSFC Grant No. 11832012).
interests or personal relationships that could have appeared to influence
the work reported in this paper.

Appendix A:. Transformation matrix and shift matrix

The explicit expression of the transformation matrix M in Eq. (3) can be given by:
⎡ ⎤
1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1
⎢0 1 − 1 0 0 0 0 1 − 1 1 − 1 1 − 1 1 − 1 0 0 0 0 ⎥
⎢ ⎥
⎢0 0 0 1 − 1 0 0 1 1 − 1 − 1 0 0 0 0 1 − 1 1 − 1⎥
⎢ ⎥
⎢0 0 0 0 0 1 − 1 0 0 0 0 1 1 − 1 − 1 1 1 − 1 − 1⎥
⎢ ⎥
⎢0 0 0 0 0 0 0 1 − 1 − − 1 1 0 0 0 0 0 0 0 0 ⎥
⎢ ⎥
⎢0 0 0 0 0 0 0 0 0 0 0 1 − 1 − 1 1 0 0 0 0 ⎥
⎢ ⎥
⎢0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 − 1 − 1 1 ⎥
⎢ ⎥
⎢0 1 1 0 0 0 0 1 1 1 1 1 1 1 1 0 0 0 0 ⎥
⎢ ⎥
⎢0 0 0 1 1 0 0 1 1 1 1 0 0 0 0 1 1 1 1 ⎥
⎢ ⎥
M=⎢
⎢0 0 0 0 0 1 1 0 0 0 0 1 1 1 1 1 1 1 1 ⎥
⎥,
⎢0 0 0 0 0 0 0 1 − 1 1 − 1 0 0 0 0 0 0 0 0 ⎥
⎢ ⎥
⎢0 0 0 0 0 0 0 0 0 0 0 1 − 1 1 − 1 0 0 0 0 ⎥
⎢ ⎥
⎢0 0 0 0 0 0 0 1 1 − 1 − 1 0 0 0 0 0 0 0 0 ⎥
⎢ ⎥
⎢0 0 0 0 0 0 0 0 0 0 0 1 1 − 1 − 1 0 0 0 0 ⎥
⎢ ⎥
⎢0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 − 1 1 − 1⎥
⎢ ⎥
⎢0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 1 − 1 − 1⎥
⎢ ⎥
⎢0 0 0 0 0 0 0 1 1 1 1 0 0 0 0 0 0 0 0 ⎥
⎢ ⎥
⎣0 0 0 0 0 0 0 0 0 0 0 1 1 1 1 0 0 0 0 ⎦
0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 1 1 1
and its inverse can be expressed as:
⎡ ⎤
4 0 0 0 0 0 0 − 4 − 4 − 4 0 0 0 0 0 0 4 4 4
⎢0 2 0 0 0 0 0 2 0 0 − 2 − 2 0 0 0 0 − 2 − 2 0 ⎥
⎢ ⎥
⎢0 − 2 0 0 0 0 0 2 0 0 2 2 0 0 0 0 − 2 − 2 0 ⎥
⎢ ⎥
⎢0 0 2 0 0 0 0 0 2 0 0 0 − 2 0 − 2 0 − 2 0 − 2⎥
⎢ ⎥
⎢0 0 − 2 0 0 0 0 0 2 0 0 0 2 0 2 0 − 2 0 − 2⎥
⎢ ⎥
⎢0 0 0 2 0 0 0 0 0 2 0 0 0 − 2 0 − 2 0 − 2 − 2⎥
⎢ ⎥
⎢0 0 0 − 2 0 0 0 0 0 2 0 0 0 2 0 2 0 − 2 − 2⎥
⎢ ⎥
⎢0 0 0 0 1 0 0 0 0 0 1 0 1 0 0 0 1 0 0 ⎥
⎢ ⎥
⎢0 0 0 0 − 1 0 0 0 0 0 − 1 0 1 0 0 0 1 0 0 ⎥
1 ⎢ ⎥
M− 1 = ⎢ 0 0 0 0 − 1 0 0 0 0 0 1 0 − 1 0 0 0 1 0 0 ⎥
4⎢⎢0 0

⎢ 0 0 1 0 0 0 0 0 − 1 0 − 1 0 0 0 1 0 0 ⎥

⎢0 0 0 0 0 1 0 0 0 0 0 1 0 1 0 0 0 1 0 ⎥
⎢ ⎥
⎢0 0 0 0 0 − 1 0 0 0 0 0 − 1 0 1 0 0 0 1 0 ⎥
⎢ ⎥
⎢0 0 0 0 0 − 1 0 0 0 0 0 1 0 − 1 0 0 0 1 0 ⎥
⎢ ⎥
⎢0 0 0 0 0 1 0 0 0 0 0 − 1 0 − 1 0 0 0 1 0 ⎥
⎢ ⎥
⎢0 0 0 0 0 0 1 0 0 0 0 0 0 0 1 1 0 0 1 ⎥
⎢ ⎥
⎢0 0 0 0 0 0 − 1 0 0 0 0 0 0 0 − 1 1 0 0 1 ⎥
⎢ ⎥
⎣0 0 0 0 0 0 − 1 0 0 0 0 0 0 0 1 − 1 0 0 1 ⎦
0 0 0 0 0 0 1 0 0 0 0 0 0 0 − 1 − 1 0 0 1
The explicit formulations of the shift matrix and its inverse are given as follows:

14
Y. Xu et al. Applied Thermal Engineering 219 (2023) 119360

⎡ ⎤
⎢ ⎥
⎢ ⎥
⎢ ⎥
⎢ ⎥
⎢ 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0⎥
⎢ ⎥
⎢ ⎥
⎢ − ux 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0⎥
⎢ ⎥
⎢ − uy 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0⎥
⎢ ⎥
⎢ ⎥
⎢ − uz 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0⎥
⎢ ⎥
⎢ ux uy − uy − ux 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0⎥
⎢ ⎥
⎢ ⎥
⎢ ux uz − uz 0 − ux 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0⎥
⎢ ⎥
⎢ uu 0 − uz − uy 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0⎥
⎢ y z ⎥
⎢ 2 ⎥
⎢ ux − 2ux 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0⎥
⎢ ⎥
⎢ u2 0 − 2uy 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0⎥
⎢ y ⎥
⎢ 2 ⎥
⎢ 0⎥
N = ⎢ uz 0 0 − 2uz 0 0 0 0 0 1 0 0 0 0 0 0 0 0 ⎥,
⎢ ⎥
⎢ − ux u2y u2y 2ux uy 0 − 2uy 0 0 0 − ux 0 1 0 0 0 0 0 0 0 0⎥
⎢ ⎥
⎢ − u u2 u2z 0 2ux uz 0 − 2uz 0 0 0 − ux 0 1 0 0 0 0 0 0 0⎥
⎢ x z ⎥
⎢ 2 ⎥
⎢ − ux uy 2ux uy u2x 0 − 2ux 0 0 − uy 0 0 0 0 1 0 0 0 0 0 0⎥
⎢ ⎥
⎢ 2 ⎥
⎢ − ux uz 2ux uz 0 u2x 0 − 2ux 0 − uz 0 0 0 0 0 1 0 0 0 0 0⎥
⎢ ⎥
⎢ − uy u2 0 u2z 2uy uz 0 0 − 2uz 0 0 − uy 0 0 0 0 1 0 0 0 0⎥
⎢ z ⎥
⎢ 2 ⎥
⎢ − uy uz 0 2uy uz u2y 0 0 − 2uy 0 − uz 0 0 0 0 0 0 1 0 0 0⎥
⎢ ⎥
⎢ u2 u2 − 2ux u2y − 2u2x uy 0 4ux uz 0 0 u2y u2x 0 − 2ux 0 − 2uy 0 0 0 1 0 0⎥
⎢ x y ⎥
⎢ ⎥
⎢ u2 u2 − 2ux u2z 0 − 2u2x uz 0 4ux uz 0 u2z 0 u2x 0 − 2ux 0 − 2uz 0 0 0 1 0⎥
⎢ x z ⎥
⎢ 2 2 ⎥
⎢ uy uz 0 − 2uy u2z − 2u2y uz 0 0 4uy uz 0 u2z u2y 0 0 0 0 − 2uy − 2uz 0 0 1⎥
⎢ ⎥
⎢ ⎥
⎢ ⎥
⎣ ⎦

and.
⎡ ⎤
⎢ ⎥
⎢ ⎥
⎢ ⎥
⎢ ⎥
⎢ 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0⎥
⎢ ⎥
⎢ ⎥
⎢ ux 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0⎥
⎢ ⎥
⎢ uy 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0⎥
⎢ ⎥
⎢ ⎥
⎢ uz 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0⎥
⎢ ⎥
⎢ ux uy uy ux 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0⎥
⎢ ⎥
⎢ ⎥
⎢ ux uz uz 0 ux 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0⎥
⎢ ⎥
⎢u u 0 uz uy 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0⎥
⎢ y z ⎥
⎢ 2 ⎥
⎢ ux 2ux 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0⎥
⎢ ⎥
⎢ u2 0 2uy 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0⎥
⎢ y ⎥
⎢ 2 ⎥
⎢ 0⎥
N − 1
= ⎢ uz 0 0 2uz 0 0 0 0 0 1 0 0 0 0 0 0 0 0 ⎥.
⎢ ⎥
⎢ ux u2y u2y 2ux uy 0 2uy 0 0 0 ux 0 1 0 0 0 0 0 0 0 0⎥
⎢ ⎥
⎢ u u2 u2z 0 2ux uz 0 2uz 0 0 0 ux 0 1 0 0 0 0 0 0 0⎥
⎢ x z ⎥
⎢ 2 ⎥
⎢ ux uy 2ux uy u2x 0 2ux 0 0 uy 0 0 0 0 1 0 0 0 0 0 0⎥
⎢ ⎥
⎢ 2 ⎥
⎢ ux uz 2ux uz 0 u2x 0 2ux 0 uz 0 0 0 0 0 1 0 0 0 0 0⎥
⎢ ⎥
⎢ uy u2 0 u2z 2uy uz 0 0 2uz 0 0 uy 0 0 0 0 1 0 0 0 0⎥
⎢ z ⎥
⎢ 2 ⎥
⎢ uy uz 0 2uy uz u2y 0 0 2uy 0 uz 0 0 0 0 0 0 1 0 0 0⎥
⎢ ⎥
⎢ u2 u2 2ux u2y 2u2x uy 0 4ux uy 0 0 u2y u2x 0 2ux 0 2uy 0 0 0 1 0 0⎥
⎢ x y ⎥
⎢ ⎥
⎢ u2 u2 2ux u2z 0 2u2x uz 0 4ux uz 0 u2z 0 u2x 0 2ux 0 2uz 0 0 0 1 0⎥
⎢ x z ⎥
⎢ 2 2 ⎥
⎢ uy uz 0 2uy u2z 2u2y uz 0 0 4uy uz 0 u2z u2y 0 0 0 0 2uy 2uz 0 0 1⎥
⎢ ⎥
⎢ ⎥
⎢ ⎥
⎣ ⎦

Appendix B:. The relationship between LB units and physical units.

The conversion of the thermos-physical properties used in this study from lattice units to physical units is shown in Table 1.

15
Y. Xu et al. Applied Thermal Engineering 219 (2023) 119360

Table 1
Conversion between lattice units and physical units.
Parameters Symbols Value of physical units Value of lattice units

Density of water/vapor ρl /ρv 849.92 /10.04 kg/m 3 7.6788/0.1086


Critical temperature Tc 647.3 k 0.03646
Specific heat of water/vapor Cvl /cvv 4.2/2.1 kJ/(kg⋅ k) 4.2/2.1
Thermal diffusivity of water/vapor kl /kv 0.68 × 00− 7 /8.2 × 00− 5 m2 /s 0.00041/0.2
Thermal conductivity of water/vapor λl /λv 0.74/0.73 W/(m ⋅k) 0.013/0.046

References [28] Q. Li, Y. Yu, P. Zhou, H.J. Yan, Enhancement of boiling heat transfer using
hydrophilic-hydrophobic mixed surfaces: a lattice Boltzmann study, Appl. Therm.
Eng. 132 (2018) 490–499.
[1] R.H. Chen, L.C. Chow, J.E. Navedo, Effects of spray characteristics on critical heat
[29] X. Ma, P. Cheng, Dry spot dynamics and wet area fractions in pool boiling on
flux in subcooled water spray cooling, Int. J. Heat Mass Transfer 45 (2002)
micro-pillar and micro-cavity hydrophilic heaters: a 3D lattice Boltzmann phase-
4033–4043.
change study, Int. J. Heat Mass Transfer 141 (2019) 407–418.
[2] M. Pasandideh-Fard, S.D. Aziz, S. Chandra, J. Mostaghimi, Cooling effectiveness of
[30] W.X. Li, Q. Li, Y. Yu, K.H. Luo, Nucleate boiling enhancement by structured
a water drop impinging on a hot surface, Int. J. Heat Fluid Flow 22 (2001)
surfaces with distributed wettability-modified regions: a lattice Boltzmann study,
201–210.
Appl. Therm. Eng. 194 (2021), 117130.
[3] M.R.O. Panão, A.L.N. Moreira, Flow characteristics of spray impingement in PFI
[31] P. Lallemand, L.S. Luo, Theory of the lattice Boltzmann method: dispersion,
injection systems, Exp. Fluids 39 (2) (2005) 364–374.
dissipation, isotropy, Galilean invariance, and stability, Phys. Rev. E 61 (2000)
[4] J. Park, D.E. Kim, Droplet dynamics on superheated surfaces with circular
6546.
micropillars, Int. J. Heat Mass Transfer 142 (2019), 118459.
[32] L.S. Luo, W. Liao, X. Chen, Y. Peng, W. Zhang, Numerics of the lattice Boltzmann
[5] J.G. Leidenfrost, On the fixation of water in diverse fire, Int. J. Heat Mass Transfer
method: effects of collision models on the lattice Boltzmann simulations, Phys Rev
9 (11) (1966) 1153–1166.
E Stat Nonlin Soft Matter Phys 83 (5 Pt 2) (2011), 056710.
[6] X. Zhong, Y. Zhang, Y. Hou, H. Feng, L. Sun, Unique dynamics of water-ethanol
[33] Y. Wu, N. Gui, X. Yang, J. Tu, S. Jiang, Improved stability strategies for pseudo-
binary droplets impacting onto a superheated surface with nanotubes, Int. J. Heat
potential models of lattice Boltzmann simulation of multiphase flow, Int. J. Heat
Mass Transfer 164 (2021), 120571.
Mass Transfer 125 (2018) 66–81.
[7] G. Liang, I. Mudawar, Review of drop impact on heated walls, Int. J. Heat Mass
[34] M. Geier, A. Greiner, J.G. Korvink, Cascaded digital lattice Boltzmann automata for
Transfer 106 (2017) 103–126.
high Reynolds number flow, Phys. Rev. E 73 (2006), 066705.
[8] G. Liang, S. Shen, Y. Guo, J. Zhang, Boiling from liquid drops impact on a heated
[35] D. Lycett-Brown, K.H. Luo, Multiphase cascaded lattice Boltzmann method,
wall, Int. J. Heat Mass Transfer 100 (2016) 48–57.
Comput. Math. with Appl. 67 (2) (2014) 350–362.
[9] Q. Li, Q.J. Kang, M.M. Francois, Y.L. He, K.H. Luo, Lattice Boltzmann modeling of
[36] G. Wang, L. Fei, K.H. Luo, Unified lattice Boltzmann method with improved
boiling heat transfer: the boiling curve and the effects of wettability, Int. J. Heat
schemes for multiphase flow simulation: application to droplet dynamics under
Mass Transfer 85 (2015) 787–796.
realistic conditions, Phys. Rev. E 105 (4–2) (2022), 045314.
[10] L. Fei, J. Yang, Y. Chen, H. Mo, K.H. Luo, Mesoscopic simulation of three-
[37] D. Lycett-Brown, K.H. Luo, Improved forcing scheme in pseudopotential lattice
dimensional pool boiling based on a phase-change cascaded lattice Boltzmann
Boltzmann methods for multiphase flow at arbitrarily high density ratios, Phys.
method, Phys. Fluids 32 (10) (2020) 103312.
Rev. E 91 (2) (2015), 023305.
[11] D. Nie, G. Guan, Study on boiling heat transfer in a shear flow through the lattice
[38] D. Lycett-Brown, K.H. Luo, Cascaded lattice Boltzmann method with improved
Boltzmann method, Phys. Fluids 33 (4) (2021) 043314.
forcing scheme for large-density-ratio multiphase flow at high Reynolds and Weber
[12] M. Zhang, J. Zhu, Z. Tao, L.u. Qiu, A quantitative phase diagram of droplet
numbers, Phys. Rev. E 94 (5–1) (2016), 053313.
impingement boiling, Int. J. Heat Mass Transfer 177 (2021) 121535.
[39] G. Wang, J. Gao, K.H. Luo, Droplet impacting a superhydrophobic mesh array:
[13] S.C. Yao, K.Y. Cai, The dynamics and leidenfrost temperature of drops impacting
effect of liquid properties, Physical Review Fluids 5 (12) (2020).
on a hot surface at small angles, Exp. Therm. Fluid Sci. 1 (4) (1988) 363–371.
[40] L. Fei, K.H. Luo, Q. Li, Three-dimensional cascaded lattice Boltzmann method:
[14] T. Tran, H.J.J. Staat, A. Prosperetti, C. Sun, D. Lohse, Drop Impact on Superheated
improved implementation and consistent forcing scheme, Phys. Rev. E 97 (5–1)
Surfaces, Phys. Rev. Lett. 108 (3) (2012), 036101.
(2018), 053309.
[15] N. Saneie, V. Kulkarni, B. Treska, K. Fezzaa, N. Patankar, S. Anand, Microbubble
[41] L. Fei, K.H. Luo, Cascaded lattice Boltzmann method for thermal flows on standard
dynamics and heat transfer in boiling droplets, Int. J. Heat Mass Transfer 176
lattices, Int. J. Therm. Sci. 132 (2018) 368–377.
(2021), 121413.
[42] S. Saito, A. De Rosis, L. Fei, K.H. Luo, K.-I. Ebihara, A. Kaneko, Y. Abe, Lattice
[16] Q. Li, Y. Yu, Z.X. Wen, How does boiling occur in lattice Boltzmann simulations,
Boltzmann modeling and simulation of forced-convection boiling on a cylinder,
Phys. Fluids 32 (9) (2020) 093306.
Phys. Fluids 33 (2) (2021) 023307.
[17] Q. Li, Q.J. Kang, M.M. Francois, A.J. Hu, Lattice Boltzmann modeling of self-
[43] Y. Xu, L. Tian, C. Zhu, N. Zhao, Numerical investigation of droplet impact on
propelled Leidenfrost droplets on ratchet surfaces, Soft Matter 12 (1) (2016)
heated surfaces with pillars, Phys. Fluids 34 (2) (2022) 023305.
302–312.
[44] L. Fei, F. Qin, G. Wang, K.H. Luo, D. Derome, J. Carmeliet, Droplet evaporation in
[18] L. Wang, S. Rong, S. Shen, T. Wang, Z. Che, Interface oscillation of droplets upon
finite-size systems: theoretical analysis and mesoscopic modeling, Phys. Rev. E 105
impact on a heated surface in the Leidenfrost state, Int. J. Heat Mass Transfer 148
(2–2) (2022), 025101.
(2020), 119116.
[45] S. Wu, H. Dai, H.e. Wang, C. Shen, X. Liu, Role of condensation on boiling heat
[19] S. Hardt, F. Wondra, Evaporation model for interfacial flows based on a
transfer in a confined chamber, Appl. Therm. Eng. 185 (2021) 116309.
continuum-field representation of the source terms, J. Comput. Phys. 227 (11)
[46] Z. Deng, X. Liu, S. Wu, C. Zhang, Pool boiling heat transfer enhancement by bi-
(2008) 5871–5895.
conductive surfaces, Int. J. Therm. Sci. 167 (2021), 107041.
[20] Z. Xu, J. Li, Z. Yao, J. Li, Effects of superheat degree and wettability on droplet
[47] Q. Li, K.H. Luo, Q.J. Kang, Q. Chen, Contact angles in the pseudopotential lattice
evaporation time near Leidenfrost point through Lattice Boltzmann simulation, Int.
Boltzmann modeling of wetting, Phys. Rev. E 90 (5–1) (2014), 053301.
J. Therm. Sci. 167 (2021), 107017.
[48] Q. Li, K.H. Luo, X.J. Li, Forcing scheme in pseudopotential lattice Boltzmann model
[21] S. Gong, P. Cheng, A lattice Boltzmann method for simulation of liquid–vapor
for multiphase flows, Phys. Rev. E 86 (1 Pt 2) (2012), 016709.
phase-change heat transfer, Int. J. Heat Mass Transfer 55 (17–18) (2012)
[49] Q. Li, K.H. Luo, X.J. Li, Lattice Boltzmann modeling of multiphase flows at large
4923–4927.
density ratio with an improved pseudopotential model, Phys. Rev. E 87 (5) (2013),
[22] H. Chen, Y. Sun, L. Li, X. Wang, Bubble dynamics and heat transfer performance on
053301.
micro-pillars structured surfaces with various pillars heights, Int. J. Heat Mass
[50] A. Markus, G. Hazi, Simulation of evaporation by an extension of the
Transfer 163 (2020), 120502.
pseudopotential lattice Boltzmann method: a quantitative analysis, Phys. Rev. E 83
[23] N.K. Singh, B. Premachandran, Numerical investigation of film boiling on a
(4 Pt 2) (2011), 046705.
horizontal wavy wall, Int. J. Heat Mass Transfer 150 (2020), 119371.
[51] Q. Li, P. Zhou, H.J. Yan, Improved thermal lattice Boltzmann model for simulation
[24] Y. Chen, K. Ling, H. Ding, Y. Wang, S. Jin, W. Tao, 3-D numerical study of
of liquid-vapor phase change, Phys. Rev. E 96 (6–1) (2017), 063303.
subcooled flow boiling in a horizontal rectangular mini-channel by VOSET, Int. J.
[52] E.O. Fogliatto, A. Clausse, F.E. Teruel, Assessment of a double-MRT
Heat Mass Transfer 183 (2022), 122218.
pseudopotential lattice Boltzmann model for multiphase flow and heat transfer
[25] X. Shan, H. Chen, Lattice Boltzmann model for simulating flows with multiple
simulations, Int. J. Therm. Sci. 159 (2021), 106536.
phases and components, Phys. Rev. E 47 (3) (1993) 1815–1819.
[53] E.O. Fogliatto, A. Clausse, F.E. Teruel, Development of a double-MRT
[26] K.H. Luo, L. Fei, G. Wang, A unified lattice Boltzmann model and application to
pseudopotential model for tridimensional boiling simulation, Int. J. Therm. Sci.
multiphase flows, Philos Trans A Math Phys Eng Sci 379 (2208) (2021) 20200397.
179 (2022), 107637.
[27] Q. Li, K.H. Luo, Q.J. Kang, Y.L. He, Q. Chen, Q. Liu, Lattice Boltzmann methods for
[54] D. Zhang, K. Papadikis, S. Gu, Three-dimensional multi-relaxation time lattice-
multiphase flow and phase-change heat transfer, Prog. Energy Combust. Sci. 52
Boltzmann model for the drop impact on a dry surface at large density ratio, Int. J.
(2016) 62–105.
Multiphase Flow 64 (2014) 11–18.

16
Y. Xu et al. Applied Thermal Engineering 219 (2023) 119360

[55] Y. Xu, L. Tian, C. Zhu, N. Zhao, Reduction in the contact time of droplet impact on [63] S. Illias, S. Hussain, Y.A. Rahim, M.A. Idris, M.E. Baharudin, K.A. Ismail, M.H. Ani,
superhydrophobic surface with protrusions, Phys. Fluids 33 (7) (2021) 073306. Prediction of maximum spreading time of water droplet during impact onto hot
[56] L. Li, C. Chen, R. Mei, J.F. Klausner, Conjugate heat and mass transfer in the lattice surface beyond the Leidenfrost temperature, Case Stud. Therm. Eng. 28 (2021),
Boltzmann equation method, Phys. Rev. E 89 (4) (2014), 043308. 101396.
[57] S. Dou, L. Hao, Numerical study of droplet evaporation on heated flat and micro- [64] P. Agrawal, S. Dash, Droplet trampolining on heated surfaces in the transitional
pillared hydrophobic surfaces by using the lattice Boltzmann method, Chem. Eng. boiling regime, Int. J. Heat Mass Transfer 190 (2022), 122811.
Sci. 229 (2021), 116032. [65] T. Tran, H.J.J. Staat, A. Susarrey-Arce, T.C. Foertsch, A. van Houselt, H.J.G.
[58] C. Zhang, H. Zhang, X. Zhang, C. Yang, P. Cheng, Evaporation of a sessile droplet E. Gardeniers, A. Prosperetti, D. Lohse, C. Sun, Droplet impact on superheated
on flat surfaces: an axisymmetric lattice Boltzmann model with consideration of micro-structured surfaces, Soft Matter 9 (12) (2013) 3272.
contact angle hysteresis, Int. J. Heat Mass Transfer 178 (2021), 121577. [66] V. Bertola, An impact regime map for water drops impacting on heated surfaces,
[59] X. Chang, H. Huang, Y. Cheng, X. Lu, Lattice Boltzmann study of pool boiling heat Int. J. Heat Mass Transfer 85 (2015) 430–437.
transfer enhancement on structured surfaces, Int. J. Heat Mass Transfer 139 (2019) [67] A.L. Biance, F. Chevy, C. Clanet, G. Lagubeau, D. QuÉRÉ, On the elasticity of an
588–599. inertial liquid shock, J. Fluid Mech. 554 (2006) 47–66.
[60] S.C. Park, M.H. Kim, D.I. Yu, H.S. Ahn, Geometrical parametric study of drop [68] H. Lastakowski, F. Boyer, A.L. Biance, C. Pirat, C. Ybert, Bridging local to global
impingement onto heated surface with micro-pillar arrays, Int. J. Heat Mass dynamics of drop impact onto solid substrates, J. Fluid Mech. 747 (2014) 103–118.
Transfer 168 (2021), 120891. [69] G. Castanet, O. Caballina, F. Lemoine, Drop spreading at the impact in the
[61] Z. Hu, X. Wu, F. Chu, X. Zhang, Z. Yuan, Off-centered droplet impact on single- Leidenfrost boiling, Phys. Fluids 27 (6) (2015) 063302.
ridge superhydrophobic surfaces, Exp. Therm Fluid Sci. 120 (2021), 110245. [70] L. Rayleigh, On the capillary phenomena of jets, Proc. R. Soc. London 29 (1879)
[62] E. Kompinsky, E. Sher, Droplet First and Second Consecutive Impacts and Droplet- 71–97.
Droplet Collision on a Hot Surface in the Film Boiling Region, J. Heat Transfer 137
(5) (2015), 051503.

17

You might also like