You are on page 1of 22

Applied Mathematical Modelling 100 (2021) 33–54

Contents lists available at ScienceDirect

Applied Mathematical Modelling


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

Fast detection of free surface and surface tension modelling


via single-phase SPH
W.K. Sun b, L.W. Zhang a,∗, K.M. Liew b
a
Department of Engineering Mechanics, School of Naval Architecture, Ocean and Civil Engineering, Shanghai Jiao Tong University,
Shanghai, 200240, China
b
Department of Architecture and Civil Engineering, City University of Hong Kong, Kowloon, Hong Kong, China

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

Article history: Surface tension plays a crucial role in many practical issues ranging from the bubble and
Received 4 February 2021 droplet dynamics to the super-hydrophilic and hydrophobic bio-inspired designs. Current
Revised 21 June 2021
smoothed particle hydrodynamics models for surface tension are mostly multiphase-based
Accepted 25 June 2021
and computationally expensive. In this work, we propose a single-phase model of surface
Available online 7 July 2021
tension based on fast free-surface detections. We develop two free-surface detection algo-
Keywords: rithms: (1) a two-step detection combining the number of neighboring particles method
Free-surface detection and the arc-method, and (2) a central angle-based method, which is proved to be much
Surface tension model more efficient than the reference counterparts. The central angle method is incorporated
Smoothed particle hydrodynamics (SPH) into the single-phase model. We verify the single-phase model through several droplet
Meshfree method dynamics simulations including initially squared droplets, the oscillating droplet, binary
Droplet dynamics head-on coalescence of droplets, and the wetting problems. Compared with the multiphase
solver, this single-phase model is much simpler in the algorithm, more efficient in com-
putation, and meanwhile retains good accuracy and robustness. Although the problems
studied in this work are limited to 2D cases, the single-phase model lends itself naturally
to large-scale 3D scenarios. The free-surface detection algorithms can also be applied in
other meshfree methods.
© 2021 Published by Elsevier Inc.

1. Introduction

Surface tension plays an essential role in many natural and industrial scenarios. A recent report shows that surface
tension has a predominant effect on the growth of cloud condensation nuclei [1]. Surface tension is not just a unique
feature of fluids, and it has been proven that surface tension in the solid phase dominates the elasticity of soft materials in
the case of small-scale indentation [2]. Nature is a master of using surface tension [3], e.g., Xia and Lei reported how water
striders (see Fig. 1) walk on the water surface by exploiting surface tension [4], which can be applied in many bio-inspired
designs [5].
Surface tension also has a substantial impact in engineering applications, e.g., metal casting, polymer fabrication, inkjet
printing [6], and food processing [7]. The most noticeable scenarios involving surface tension are bubbles and droplets. There
are many works on cavitation bubbles [8], collapsing bubbles [9], bubble bursting jets [10], droplet coalescence [11] (also
see Fig. 1), and droplet impact [12] and its rebounds [13]. Since the time of Thomas Young in the early nineteenth cen-


Corresponding author.
E-mail address: lwzhang@sjtu.edu.cn (L.W. Zhang).

https://doi.org/10.1016/j.apm.2021.06.029
0307-904X/© 2021 Published by Elsevier Inc.
W.K. Sun, L.W. Zhang and K.M. Liew Applied Mathematical Modelling 100 (2021) 33–54

Fig. 1. Observations and experiments of the surface tension effect: (a) top view of a water strider, (b) surface tension in vicinity of the water strider’s leg
and its molecular causes, (c) experimental setup of the droplet coalescence, (d) shape evolution of two coalescing droplets.

tury, models and experiments for surface tension have been evolving with mathematics and engineering technologies. Now
surface tension can be studied with advanced experimental equipment, e.g., the close-up high-speed video imaging [12,14],
and analytical tools, e.g., complex functions and Fourier series [15]. With the advances in numerical tools and computer
capacity, modeling and simulation have become a new dominating force for investigating surface tension and its effects on
fluid dynamics.
SPH is a Lagrangian mesh-free method [16], first introduced independently by Lucy [17] and Gingold [18], which is flex-
ible to incorporate extra physical effects, e.g., surface tension and heat transfer. In SPH, just by assigning particles with
different color values, different phases can be modeled in a very natural way. Due to the conceptual simplicity, it has been
successfully applied to many fluid dynamics problems, such as free-surface flow [19,20], multiphase flow [21], fragmentation
[22], bubble rising [23], and underwater explosion [24]. More recently it has also been applied to model wetting phenom-
ena [25], binary droplet deformation [26], and dynamic behavior of metal droplet impact on the dry smooth wall [27].
Three fundamental issues have to be addressed for surface tension modeling using SPH: the dynamics model, the interface
detection, and the curvature representation.
As for the dynamic effect of surface tension, there are two major approaches, based on either micromechanisms and
molecular potential forces [28,29] or macroscale modeling [30–37]. Among the macroscale models, Brackbill proposed a
continuous surface force (CSF) model [38], which uses a body force to represent the equivalent effect of the surface force in
classical theory. This feature allows it to be readily added into the governing equations of fluid dynamics; it thus is adopted
in many simulations involving the surface tension effect.
In the early works of surface tension modeling with SPH, Morris used the color function to represent the interface and
discussed three SPH algorithms based on the CSF model [30]. But the problem is constrained to low-density ratio scenarios
only. Hu and Adams modified the SPH formulation by considering the distinct contribution of different phases [31,39]. Al-
though losing some simplicity and generality, it was proved to be capable of modeling the surface tension effect in a large
density ratio with good accuracy. Instead of using the color gradient representation for normal vector and curvature, Adami
et al. adopted a weighted color function and a reproducing divergence formulation for curvature based on earlier mathemat-
ical work on curvature formula [40], and this method was proved to surpass the former work in accuracy and robustness
[32]. Besides, a rigorous variational formulation was demonstrated, and the concept of the surface stress tensor is proposed,
which shows good accuracy and robustness in multiphase flow problems [41]. Zhang, based on Lagrangian interpolation
in local coordinate systems, reconstructed the interface and represented the curvature, which exhibits good accuracy in
head-on binary collision examples [35]. Based on Morris’s model [30], wetting problems were also successfully simulated
via the SPH framework [42]. Szewc et al. systematically studied the SPH application in multi-phase flows and compared
different surface tension models [43]. These works do display diverse ideas in surface modeling with SPH, particularly in

34
W.K. Sun, L.W. Zhang and K.M. Liew Applied Mathematical Modelling 100 (2021) 33–54

curvature representation based on CSF; however, they are multiphase solvers, which are relatively expensive compared with
the single-phase solver [44].
The performance of a single-phase solver relies on the accuracy of free-surface detection. Many neighborhood-based
approaches have been developed for this problem in particle methods. By dealing with the sparseness of the neighborhood
for a free-surface particle with different philosophies, the free-surface detection methods can be classified as the filter-based
methods and the geometry exposure methods. The former one filters the free-surface particles by setting an appropriate
threshold of properties such as the number of neighboring particles (NNP) [45], particle number density [46], the magnitude
of the normal vector, divergence [47], centroid coordinate [48], and radius of circumcircle obtained by Delaunay triangulation
[49]. The latter one detects the free-surface particles by checking the completeness of a geometrical target such as the
arc-method [50] or scanning a zone for neighboring particles [44,51] or target points such as the spoke method [50] and
back-point method [50]. Besides, the coupling of two or more methods can also be used to further boost the performance of
free-surface detection [52,53]. However, few of these methods can balance the accuracy and efficiency issues, and meanwhile
keep simplicity in algorithm implementation.
In this paper, we propose two novel algorithms for fast and accurate interface detection: (1) a two-step detection com-
bining the NNP method and the arc-method, which is denoted as the NNP+Arc method in the following sections; (2) the
central angle method. The latter one is further applied in the developed single-phase CSF model and implemented by the
SPH method. In Section 2, we briefly introduce the SPH formulation and some numerical treatments. The proposed free-
surface detection algorithms are described in Section 3. Section 4 is dedicated to formulating the single-phase surface ten-
sion model. Test cases for verification and application are demonstrated in Section 5, and we close our paper with some
critical remarks in Section 6.

2. SPH formulation for surface tension modeling

Surface tension is a physical property of any fluid, which is an interaction between a liquid phase and another fluid
(gas or another liquid). How to model surface tension accurately largely depends on how to represent the free surface (or
interface between two fluids) appropriately, and meanwhile correctly model the dynamic law dictates this phenomenon.
SPH is a Lagrangian meshfree method, which is very suitable for this problem since the fluid is represented as particles
with physical properties. By assigning each particle with a different ‘color’ (or ‘type’), it easily differentiates an inner fluid
particle from a free-surface counterpart. Thus, it is more convenient to impose surface tension force on free-surface particles
and model surface tension compared with grid-based methods.

2.1. Governing equations

The fluid flow is governed by the Navier–Stokes equations

Dρ ∂ vi
= −ρ , (1)
Dt ∂ xi

Dv 1 
= −∇ p + f v + f s + f e , (2)
Dt ρ
where ρ , v, and p are the fluid density, velocity, and pressure respectively.fv , fs , and fe are the viscosity force, the body force
to represent surface tension in the CSF framework, and the external body force respectively. Here, a typical source of the
external body force fe is gravity, where fe = ρ g.
The fluid incompressibility is approximated as a weakly-compressible model by using an equation of state (EOS) in the
form [16]
 
p = p0 (ρ /ρ0 )γ − 1 + pb , (3)

where ρ 0 is the reference density, p0 and pb are the reference and background pressure respectively, and γ =7. Set p0 = 1,
ρ 0 = 10 0 0 kg/m3 , and pb =0. c0 > 10vmax can limit the variation of fluid density to a maximum of 1%, where vmax is the
maximal magnitude of velocity vector in the fluid.

2.2. Continuum surface force model

Based on the continuum surface force (CSF) model, the surface tension is defined as a body force fs and

f s = −σ κ nδ , (4)

where σ , κ , and n are the surface tension coefficient, curvature, and unit normal vector of the interface respectively, and
δ  is a delta function for limiting the regions of the body force to the vicinity of phase boundaries.

35
W.K. Sun, L.W. Zhang and K.M. Liew Applied Mathematical Modelling 100 (2021) 33–54

Fig. 2. SPH approximation of the reference particle by using the weighted summation over its neighbors. The dashed lines indicate the weight values for
the corresponding particles in the support domain.

2.3. The SPH solver

The governing equations are solved with the SPH method. In SPH, the fluid is represented by particles with certain
physical properties. The core philosophy of SPH is to approximate the property of a reference particle a located in xa (directly
called particle xa for simplicity) by using the smoothed (or weighted) summation over its neighbors. The weight function
is called the kernel, as shown in Fig. 2, and the neighbors are required to lie in the compact domain of particle xa . For
instance, the density of particle xa can be expressed as

ρa = ρbWab , (5)
b

where Wab is called the kernel function (or kernel), and it has a form

Wab = W (xa − xb , h ), (6)


where h is the smoothing length.
The SPH approximation of a general function in an integral form can be written as
 f
f (x ) ≈  f (x ) = (x )W (x − x , h)dx , (7)

where W(x − x , h) is the Gaussian-like kernel or weight function, and h is the smoothing length. A kernel in SPH should
satisfy three requirements, namely the unity property, the compact support domain, and the delta function property [16]

W ( x − x  , h )d x  = 1 , (8)

W (x − x , h ) = 0 for |x − x | > κ h, (9)

lim W (x − x , h ) = δ (x − x ). (10)
h→0

The kernel function we adopted in this work is the piecewise cubic spline kernel [20], and the discretized form can be
written as

N
mb
 f (xa ) = f (xb )W (xa − xb , h ). (11)
b=1
ρb
With Eq. (11), using the strategy of integration by parts whilst considering the kernel properties, divergence representa-
tions of a vector field f(x), and the gradient of a scalar function f(x) can be furnished in the following SPH discretized forms
respectively


N
mb
∇ · f (xa ) = [ f (xb ) − f (xa )] · ∇aWab , (12)
b=1
ρb


N
f ( xb ) f ( xa )
∇ f (xa ) = ρa mb + ∇aWab , (13)
b=1
ρb2 ρa2

36
W.K. Sun, L.W. Zhang and K.M. Liew Applied Mathematical Modelling 100 (2021) 33–54

where ∇ a Wab is the gradient of kernel function W(xa − xb ,h) with respect to the particle a.
Based on these formulations, the Navier-Stokes equations can be reformulated as follows:

d ρa  mb N
= ρa v ·∇ W , (14)
dt ρb ab a ab
b=1


dv  pb pa
N
ρa a = ρa mb + 2 ∇aWab + fav + fas + fae , (15)
dt ρb ρa
2
b=1

where vab = va − vb .
It should be noted that many reformulated versions of SPH have been proposed [54] since the proposed original works
[16,18], for instance, the δ -plus SPH scheme [55–58], and the multi-resolution parallel framework [59–61], but for simplicity
and generality, we adopt the standard SPH formulation in this work.

2.4. Artificial viscosity

The Monaghan type of artificial viscosity [16] is the most widely used artificial viscosity in the SPH literature. It not only
provides the necessary dissipation for converting kinetic energy into heat at the shock front but also prevents unphysical
penetration for particles approaching each other. The detailed formulation is as follows:

−α

cab φab +β
φab
2
vab · xab < 0

ab = ρ ab , (16)
0 vab · xab ≥ 0

where

h̄ab vab · xab


φab = , (17)
|xab |2 + ϕ 2
1
c̄ab = (ca + cb ), (18)
2

1
ρ̄ab = (ρa + ρb ), (19)
2

1
h̄ab = (ha + hb ), (20)
2
in which vab = va − vb and xab = xa − xb . In the abovementioned equations, α
and β
are constant parameters that are
typically set to 1 and 2 respectively, φ is the parameter used to prevent numerical divergence when two particles are
approaching each other and is set to 0.1h, and c̄ab , ρ̄ab , and h̄ab are the average sound velocity, density and smoothing
length, respectively.
In addition, another two types of artificial viscosity were also proposed. The first one proposed by Cleary [62] reads

ξ 4 μa μb vab · xab

ab = , (21)
ρa ρb (μa + μb ) |xab |2 + ϕ 2
where μ is the dynamic viscosity, ξ = 4.96333, and φ is the same as that in the Monaghan type. The second one by Cleary
and Monaghan [63] reads
νa + νb vab · xab

ab = 8 , (22)
(ρa + ρb ) |xab |2 + ϕ 2
where ν is the kinematic viscosity, and φ is also the same as the one in the Monaghan type.
It was reported that their differences in performance are very small [64], and we adopted the last one Eq. (22) in all
cases in this study.
With the artificial viscosity term, the discretized forms of Eqs. (15) can be rewritten as

dva  pb
N
pa
ρa = ρa mb + −
ab ∇aWab + fas + fae , (23)
dt ρb2 ρa2
b=1

where
ab is the artificial viscosity defined in Eq. (22), and the other variables are consistent with those in Eq. (15).

37
W.K. Sun, L.W. Zhang and K.M. Liew Applied Mathematical Modelling 100 (2021) 33–54

Fig. 3. The detection criteria of the arc-method: (a) the reference particle is detected as a free-surface particle since it has an uncovered arc, (b) the
reference particle is detected as an inner particle because it has a closed circle with a radius of x.

Algorithm 1
The NNP+Arc method for free-surface detection.

3. Free-surface detection

There are two novel free-surface detection algorithms we propose in this work. The first one is a two-step approach
combining the NNP method and the arc-method, which is denoted as the NNP+Arc method. The second one is called the
central angle method based on the neighborhood sparseness of the surface particles. Here, neighborhood sparseness indi-
cates the ratio of the gap area (the area of the largest sector formed by two neighboring particles and the reference particle)
to the whole area of the support domain for a reference particle. For brevity, the two methods are denoted as Method Ⅰ and
Method Ⅱ respectively.

3.1. Method Ⅰ

The NNP method is formulated as:

Na < α N̄, (24)

where Na is the number of neighboring particles (NNP) for particle a, N̄ is the average NNP over all particles and α is the
free-surface threshold parameter for the NNP method.
The arc-method is a geometry exposure-based method, the detection criteria of which are illustrated in Fig. 3. As implied
in its name, the major concerns in this method are the arcs generated by intersections of the circles with unit particle
spacing ࢞x (average spacing of initial configuration), whose centers are located at the reference particle and its neighbors
(or neighboring particles) in the support domain. If the Boolean addition of these arcs can form a closed circle, then the
reference particle is an inner particle; otherwise, it is a free-surface particle.
The algorithm for Method Ⅰ can be described as:
Algorithm 1.

38
W.K. Sun, L.W. Zhang and K.M. Liew Applied Mathematical Modelling 100 (2021) 33–54

Fig. 4. The detection procedures of the central angle method.

Fig. 5. The two-step curvature representation: Step 1 is the determination of the three local particles and Step 2 is the calculation of the radius of the
circumcircle formed by the three local particles.

3.2. Method Ⅱ

As coined in its name, central angles are the focuses of this method. We observed a fact that the maximal central angle
formed by a free-surface particle and its two neighbors tend to be much larger than its inner-particle counterparts. There-
fore, the central angle can be utilized as a measuring quantity for free-surface detection. The overall algorithm for Method
Ⅱ is depicted in Fig. 4, where M is the total number of neighboring particles for the reference particle under detection and ε
is the free-surface parameter in the central angle method. Step 1 is the evaluation of the phase angles between the position
vector and the coordinate vector for each neighboring particle. Step 2 is to sort these phase angles in ascending order. Step
3 is the determination of central angles by calculation the difference between each two neighboring sorted phase angles.
Step 4 is to check whether the maximum central angle of the reference particle is beyond the threshold value: if so, the
reference particle under detection is a free-surface particle; otherwise, it is an inner particle. In this work, ε is set to be
7/12. An explanation of the chosen ε and α in Method I, together with a possible 3D extension of the central angle method
is presented in Appendix A.
The detailed algorithm is shown as follows.
Algorithm 2.

4. Surface tension model

The CSF model is used for surface tension formulation. Here we propose a new curvature calculation algorithm. This
algorithm is based on the free-surface detection results and depends on the local free-surface particles only.

4.1. Curvature representation

The curvature representation method we propose here is based on the definition of curvature as follows:
1
κ= , (25)
Rcc
where Rcc is the radius of the circumcircle of three neighboring local free-surface particles, as shown in Fig. 5. The radius
of the circumcircle can be determined through two steps. Step 1 is the determination of two neighboring particles for
the reference particle by filtering with a selecting circle with a radius of Rs . Rs = 1.4 × dx is used in this work. Step 2 is

39
W.K. Sun, L.W. Zhang and K.M. Liew Applied Mathematical Modelling 100 (2021) 33–54

Algorithm 2
The central angle method for free-surface detection.

Fig. 6. Determination of normal vectors with three local particles.

the determination of the circumcircle with the three particles obtained, which can be achieved by calculating the center
position of the circumcircle first, as depicted in Fig. 5. Once the center position is determined, Rcc is represented by the
distance between the center position and any one of the three local particles.

4.2. Normal vector calculation

The normal vector is calculated approximately by the vector addition of unit relative position vectors between the refer-
ence particle and the neighboring particles.

n a = − ( r 1 + r 2 ), (26)

where r1 and r2 are the unit relative position vectors between the reference particle a and the neighboring particles b1 and
b2 respectively, as shown in Fig. 6.

40
W.K. Sun, L.W. Zhang and K.M. Liew Applied Mathematical Modelling 100 (2021) 33–54

Fig. 7. Surface tension models: the upper row is the definition of surface tension and the lower row is Young’s theory on the pressure difference due to
surface tension effect.

Fig. 8. The schematic diagram and formulation of contact line force FCFL at the vapor-liquid-solid triple line.

4.3. Implementation of the surface tension model

The surface tension model is formulated based on the equivalence of the total force in an infinitely small cylindrical shell
element of the fluid domain. On the one hand, the definition of surface tension coefficient σ is the in-plane traction force
imposed on the interface (or free-surface) per unit length, as shown in Fig. 7. Thus, the total force for an interfacial zone
with a length of L is
θ ∼
Fσ = 2σ Lsin = σ L θ , (27)
2
given that θ is small enough.
On the other hand, Thomas Young found that there is a pressure difference p between the inner and outer sides of the
interface due to the surface tension effect. Therefore, surface tension can be equivalently treated as the uniformly distributed
pressure force with the identical magnitude of p, as can be seen in Fig. 7. The total pressure force is

F p =  pLRθ . (28)
Eqs. (27) and (28) should be equal since the total force should be equivalent, therefore, we have
σ
p = = σ κ. (29)
R
Brackbill treated the surface tension as a body force imposed on each material particle. In this perspective,
F p σ κn
fs = − n=− , (30)
LRθ H H
where H is the thickness of the free-surface band. In this work, H is equal to the grid size (or initial particle spacing),
since the free surface is composed of a single layer of fluid particles in the single-phase framework.

4.4. Contact line forces

The surface tension effect takes the form of contact line force at the vapor-liquid-solid triple line. As can be seen in
Fig. 8, the contact line force FCFL at the vapor-liquid-solid triple line can be derived from the surface tension coefficient, and
the dynamic and equilibrium contact angles.

41
W.K. Sun, L.W. Zhang and K.M. Liew Applied Mathematical Modelling 100 (2021) 33–54

Fig. 9. Computational performances of different free-surface detection methods for a fluid domain with a sinusoidal free-surface. Left panel: square lattice
configuration. Right panel: hexagonal lattice configuration.

Fig. 10. Accurate free-surface detections of two-particle systems with different lattice configurations. Upper panel: square lattice configuration. Lower
panel: hexagonal lattice configuration. Red particles are those detected as belonging to the free surface.

Similar to the continuum force theory by Brackbill, the contact line force can also be treated as a body force imposed on
each material particles as follows,
FCF L L σ (cosθeq − cosθ )
fCF L = τ= τ, (31)
L H H (H )2
where τ is the unit tangent vector along with the vapor-solid interface.

5. Results and discussion

5.1. Free-surface detections

We test three cases to verify the developed central angle method and the NNP+Arc method. The arc-method is taken
as the reference method with no error, and the computational cost is normalized to unity. The reasons are twofold: first,
the arc-method is a totally geometry based method (no tunable parameters involved), and second, the detection result
is unique (either surface particle or inner particle). The first case is the detection of a fluid domain of square lattice (a
frozen configuration with no SPH simulation of the dynamic situation involved, the same for hexagonal lattice below) with
a sinusoidal free surface. As can be seen in Fig. 9, all the methods yield accurate computational results except for the
umbrella-method [51], which has an error of slightly less than 10 %. While the costs for these methods are quite different.
The cost of the NNP method is negligible compared to the remaining four methods, among which the NNP+Arc method is
the most efficient one, followed by the central angle method.
For the second case, we test a fluid domain of hexagonal lattice with the same sinusoidal free surface as the first case.
This time the central angle method matches with the umbrella method in computational cost while the NNP method leads
the efficiency again, followed by the NNP+Arc method. On the other side, only the NNP method suffers from serious accu-
racy issues for this case.
It can be observed from these two simple test scenarios that the NNP+Arc method has the best balance between accuracy
and efficiency followed by the central angle method. The accurate test results for these two cases are displayed in Fig. 10.
As can be seen, one single layer of free-surface particles is detected, which is consistent with the definition of free-surface.

42
W.K. Sun, L.W. Zhang and K.M. Liew Applied Mathematical Modelling 100 (2021) 33–54

Fig. 11. Computational performances of different free-surface detection methods for a fluid domain from a result sample of the dam break test.

Fig. 12. The free-surface detection results of different methods for the dam break sample.

The third case is the detection of the fluid domain from a result sample of the dam break test via the SPH method.
As can be seen in Fig. 11, the NNP method has a noticeable error when the free-surface parameter is set slightly larger,
regardless of its leading efficiency. The NNP+Arc method once again dominates the comprehensive performance of accuracy
and efficiency. The central angle method is much more accurate than the umbrella method although its efficiency a little
lagged behind this time.
From all these test cases, we can conclude that the NNP+Arc method is the best player among the five methods due to
its successful combination of the accuracy advantage of the arc-method and the efficiency strength of the NNP method. It is
also worth noting that the central angle method is much efficient than the arc-method while maintaining nearly the same
accuracy.
The test results of the dam break sample are shown in Fig. 12. As can be seen, a few "underwater" red particles are
detected as free surface particles in the arc-method because they are apart with comparably large distances, and they form
gaps with large enough volumes. This can be quite reasonable when relating to the small underwater cavity bubbles in the
natural world. As can also be noticed, the NNP method tends to overestimate the number of free-surface particles due to
the improper set of the free-surface parameter. This implies that there should be a lot of trials to determine the appropriate
free-surface parameter. However, with the arc-method conducted in a second step in the NNP+Arc method, the results are
much more accurate than those achieved purely by the NNP method. Besides, the central angle method and the umbrella
method can work independently and exhibit a satisfying cost-accuracy performance.

43
W.K. Sun, L.W. Zhang and K.M. Liew Applied Mathematical Modelling 100 (2021) 33–54

Fig. 13. Shape evolution of the initially squared droplet.

Fig. 14. Free surface and curvature distribution of the initially squared droplet at 0.2 s.

5.2. Initially squared droplet

The central angle method is incorporated into the single-phase surface tension model and implemented by the SPH
solver. The deformation process of an initially squared droplet is investigated to verify the single-phase surface tension
model we propose. The fluid density is 10 0 0 kg/m3 and the dynamic viscosity is 0.1 Pa • s. The side length of the droplet
is 1 cm and the surface tension coefficient is 0.025 N/m in this test. A total of 41 × 41 particles are used in this test. As
can be observed in Fig. 13, the droplet deforms under the surface tension effect, and eventually converges to a circular
droplet. This shape evolution process is consistent with the results of the multiphase surface tension solver [32] as well as
the single-phase solver [44].
The free surface and curvature distribution at 0.2 s are displayed in Fig. 14. As can be observed, the shape of the free
surface is highly consistent with the theoretical result. The curvature distribution results also agree with the theoretical
counterpart well except for a discrepancy at the sharp corners. This discrepancy is caused by the local nature of the proposed
curvature calculation algorithm, which, however, imposes negligible influence on the final results of simulations.
The average pressure difference between the inner and outer sides of the free surface is quantitatively assessed, as shown
in Fig. 15. As can be seen, the average pressure difference first experiences an oscillation, which eventually vanishes due to
the numerical dissipation term in this model. The average pressure difference converges to a constant value, which agrees
with the theoretical one as formulated in Eq. (29) in Section 4.3. The average pressure difference error between the numer-
ical and theoretical results is less than 5%. It can also be observed that the single-phase surface tension model shows good
convergence property as the numerical results approach the theoretical counterparts with the increase of spatial resolution
from L/20 to L/40. The pressure distribution is also presented in Fig. 15. As can be noticed, the pressure is nearly constant
since the curvature of the droplet surface is eventually converging to a constant value. The discrepancy at the corner spot is
possibly due to the sensitivity of the curvature representation method in this work since we use local particles for curvature

44
W.K. Sun, L.W. Zhang and K.M. Liew Applied Mathematical Modelling 100 (2021) 33–54

Fig. 15. The pressure along the free surface. Left panel: the average pressure evolution of free-surface particles. Right panel: pressure profiles in the
steady-state solution at t=0.5s.

Fig. 16. Comparison between the single-phase and multiphase surface tension models in droplet shapes in a stable state. The black dashed lines indicate
the theoretical results of the droplet boundary.

Table 1
Comparison between the single-phase and multiphase surface tension models in computational cost.

Surface tension model Computational cost for 50 0 0 0 steps (s) Normalized computational cost

Single-phase 120 1
Multiphase 3480 28

calculation. Due to this local sensitivity, the discrepancy is overestimated compared with the true value. However, this local
discrepancy does not have much effect on the overall computational results, as shown in Figs. 13 and 15.
The comparisons between the single-phase and multiphase surface tension models in droplet shapes and the computa-
tional cost are presented in Fig. 16 and Table 1. As can be seen in Fig. 16, the numerical results of the single-phase model
are in good consistency with both the multiphase numerical results and the theoretical counterparts. It is worth noting that
although both the single-phase and the multiphase model can yield satisfying results, the computational cost of the former
one is significantly reduced compared with the latter one, which can be observed in Table 1. The factor is possibly caused
by two contributions: first, the computational complexity is proximately scaling as O(N2 ), thus, 4 times of the total particles
lead to about 16 times of the computational cost; second, there are a considerable amount of extra ghost particles for the
implementation of boundary conditions in the multiphase method, which leads to a further increase in computational cost.

5.3. Oscillating droplet

For this problem, the computational domain is a circular fluid region with a radius R of 0.564 cm. Different spatial
resolutions are tested, and the fluid domain consists of the 40 0, 90 0, and 160 0 particles respectively, which are input from
the steady-state results of the initially square droplet tests. For the fluid, we set the density to be 10 0 0 kg/m3 , and the
dynamic viscosity is 0.1 Pa • s. To compare with the theoretical oscillation period, surface tension coefficients of 0.025, 0.1,
0.4, and 0.9 N/m are studied respectively. We set the initial velocity of this fluid domain as

2
 
vx = v0 rx0 1 − ryr0 exp − rr0
 2
  , (32)
vy = −v0 ry0 1 − rxr0 exp − rr0
45
W.K. Sun, L.W. Zhang and K.M. Liew Applied Mathematical Modelling 100 (2021) 33–54

Fig. 17. Shape evolution of the oscillating droplet. σ =0.025 N/m.

Fig. 18. Left panel: the time evolution of the normalized mass center position of the upper right quarter of the oscillating droplet. x is the particle
spacing, L=0.01 cm, and σ =0.025 N/m. Right panel: the relationship between oscillation period and surface tension coefficient.

where v0 =1 cm/s, r0 =0.141 cm, and r is the distance between the point (x, y) to the center of the droplet. This velocity
field is a divergence-free field. The distribution of velocity at the initial state and the shape evolution process are shown in
Fig. 17. As can be seen, the circular droplet oscillates due to the initial kinetic energy and eventually converges to a steady
circular shape as the kinetic energy is completely dissipated. This process can also be observed in the quantitative results
shown in the left panel of Fig. 18. In addition, good convergence of the results can be observed in the left panel of Fig. 18.,
which shows the robustness of the proposed method. The theoretical oscillation period is as follows [65]:

R3 ρ
τ = 2π , (33)

where τ is the oscillation period, R is the radius of the circular droplet, ρ is the fluid density, and σ is the surface tension
coefficient. As can be seen in the right panel of Fig. 18, the simulation results of the oscillation period coincide with the
theoretical ones, which displays the accuracy of the proposed surface tension model.

5.4. Head-on binary coalescence

To further verify the single-phase surface tension model in more complex scenarios, we carry out two head-on binary
coalescence tests. The radius ratio of the two droplets is 1:1 for the first case, and it is 1:2 for the second one. In the first
case, both of the droplets have a radius of 0.564 cm, and in the second case, the larger droplet has a radius of 0.564 cm. The
input data is also from the initially-squared droplet tests. The characteristic length is L=1 cm. The surface tension coefficient
is 0.005 N/m, the mass density is 10 0 0 kg/m3 , and the dynamic viscosity is 0.1 Pa • s. The two droplets move towards each
other at a relative velocity of 0.04 m/s. The Reynolds number is 4, and the Weber number is 3.2. The coalescence processes

46
W.K. Sun, L.W. Zhang and K.M. Liew Applied Mathematical Modelling 100 (2021) 33–54

Fig. 19. Coalescence of two droplets with a radius ratio of 1:1. The upper panel is the multi-phase results and the lower panel is from the single-phase
model.

Fig. 20. Coalescence of two droplets with a radius ratio of 1:2.

from both the multi- and single-phase methods for the first case are shown in Fig. 19. There are three stages, and during the
initial stage of coalescence, the droplets will go through a process of interface instability, which leads to interface reduction
and droplet merging. They form an unstable circular droplet at the end of this stage. At the second stage, the circular
droplet continues contracting in the collision axis, and form an oval shape. Finally, after a short period of oscillation, the
oval droplet eventually evolves into a new larger circular droplet. As can be observed in Fig. 19, some liquid particles are
trapped in the second phase in the multi-phase simulation, which is spurious and inaccurate. In contrast, the single-phase
method proposed in this work can naturally avoid this issue with even less computational cost. The oscillation can also
be observed in the left panel of Fig. 21. This shape evolution process is consistent with the results of the Galerkin finite
element method [66] as well as the single-phase solver [44].
The results for the second case are presented in Fig. 20. As can be noticed, there is an obvious flow-in phenomenon
between the two different-sized droplets, and the flow-in direction is from the smaller one to the bigger one. After a short
period, the two droplets merge into one bigger droplet. The oscillation process is negligible, however, for this case. This
smoothed oscillation is probably due to the large inertia of the bigger droplet, and the dissipation effect.
The mass center position evolution of the upper half parts of the droplets is presented in Fig. 21. It can be clearly seen
that both the simulation results of radius ratio 1:1 and 1:2 match with the theoretical ones (at stable state) remarkably
well.

47
W.K. Sun, L.W. Zhang and K.M. Liew Applied Mathematical Modelling 100 (2021) 33–54

Fig. 21. Evolution of mass center position in y direction of the upper half parts of the droplets in head-on binary coalescence tests. Left panel: the radius
ratio is 1:1. Right panel: the radius ratio is 1:2.

Fig. 22. Shape evolution of an initially squared droplet on a solid wall with θ eq =120°.

In addition, we also experimented on droplet coalescence, which has been displayed in Section 1 of this paper. The
experimental setup of the droplet collision is shown in Fig. 1. As can be seen, the equipment includes a light source, a
transparent panel (not necessarily made of glass), and a camera that can cast high-resolution videos. The oil in the exper-
iment has a density of 931Kg/m3 . In this work, we used a video mode of 4K and 30 fps, and the experimental results in
Fig. 1 are obtained from the video by selecting images of certain moments.
As can be observed, the shape evolution of the coalescing droplets with a radius ratio of 1:2 also coincides with the
experimental results (shown in Fig. 1 in Section 1). This further validates the accuracy and robustness of the single-phase

48
W.K. Sun, L.W. Zhang and K.M. Liew Applied Mathematical Modelling 100 (2021) 33–54

Fig. 23. Shape evolution of an initially squared droplet on a solid wall with θ eq =150°.

model we propose in this work. The discrepancy between simulation and experimental results in the time elapsed during
shape evolution may be caused by the fact that extra surface tension is imposed on the upper surface of the droplets in the
experiment while in 2D simulation the droplets are merely subject to in-plane surface tension. 3D simulation may address
this issue since it concerns the identical boundary condition as the experiment counterpart.

5.5. Initially squared droplet on a solid wall

To further verify our proposed free-surface detection methods and the single-phase surface tension model, we also test
an initially squared droplet on a solid wall. The liquid phase is water. The side length of the droplet is 1 mm, and the
surface tension coefficient is 0.07275 N/m in this test. Referring to the recent publication [25], the equivalent initial radius
R0 is 0.7979 mm (if the droplet is initially semicircular as the one in [25]), and the Eötvos number E ö = ρL gR0 2 /σ = 0.086.
It was proven that the stable droplet has a shape as in the case when gravity is absent when E ö
1 [25]. Therefore, in
this case, simulation results can be verified with the analytical solution (formulated by assuming that the stable droplet is
a spherical cap cut from the wall):
H = Req (1 − cos(θeq ) ),
W = 2Req sin(θeq ), (34)
where

π
Req = R0 . (35)
2(θeq − sin(θeq )cos(θeq ) )
A total of 40 × 40 particles are used in this test. As can be observed in Figs. 22 and 23, the droplet deforms with the
driving forces of surface tension at sharp corners, and it gradually evolves into a stable state with the equilibrium contact
angle close to the theoretical value of 120° and 150° respectively.
The contact length and the droplet height are compared in the left panel of Fig. 24. As can be observed, good agreement
is achieved between the numerical results and the theoretical counterparts when the equilibrium contact angle θ eq ranges
from 100° to 150°. However, when θ eq < 90° tensile instability kills the simulation, and for θ eq greater than 150° large error

49
W.K. Sun, L.W. Zhang and K.M. Liew Applied Mathematical Modelling 100 (2021) 33–54

Fig. 24. Left panel: the contact length and droplet height with varying contact angles. Right panel: time history of the kinetic energy of the initially
squared droplet on a solid wall.

appears. For the unstable regime, certain numerical techniques may be applied to address the problem, e.g., the δ -plus SPH
scheme [55–58].
The kinetic energy history of the droplet is presented in the right panel of Fig. 24. As can be observed, the kinetic energy
of the particle system increases to a peak value with the acceleration of particles due to the surface tension effect and the
contact line forces. After that, the kinetic energy decreases fast to a negligible value as a result of the dissipation and the
disappearance of contact line forces. The decrease in kinetic energy is in more than 5 orders of magnitude, which shows
the temporal convergence of this test. This convergence characteristic implies satisfying robustness of the numerical model
with θ eq ranging from 100° to 150°.

6. Conclusion

In this study, we propose two free-surface detection methods, which are verified via the square and hexagonal lattice
domain, and the arbitrary particle distribution cases. The first free-surface detection method is a two-step coupling method
called the NNP+Arc method, which shows the best balance between accuracy and efficiency in the studied cases. The second
one is called the central angle method, which works alone and outperforms its counterpart the umbrella method in accuracy,
and is much more efficient than the arc-method. These two methods can be applied in surface tension models, multiphase-
flow solvers, and those numerical techniques dependent on free-surface detections.
In addition, we develop a novel single-phase surface tension model via SPH framework by incorporating the central
angle method for free-surface detection, which is verified to be accurate and robust through several droplet dynamics cases.
Retaining simplicity in the algorithm, this model also uses much fewer particles than the multiphase solvers, thus can save
a lot of computational resources and power, which reveals its potential in applications, e.g., dynamics of multiple droplets,
bubble initiation and explosion, wetting problems, and super-hydrophobic designs. The test cases in this work are limited
to low Reynolds number and low Weber number cases. For other scenarios, more investigation work will be carried out and
reported in our future research works.

Declaration of Competing Interest

The authors declare that they have no known competing financial interests or personal relationships that could have
appeared to influence the work reported in this paper.

CRediT authorship contribution statement

W.K. Sun: Conceptualization, Methodology, Data curation, Formal analysis, Validation, Visualization, Writing – original
draft, Writing – review & editing. L.W. Zhang: Project administration, Resources, Supervision, Conceptualization, Methodol-
ogy, Writing – review & editing. K.M. Liew: Project administration, Resources, Supervision, Conceptualization, Methodology,
Writing – review & editing.

Credit authorship contribution statement

W.K. Sun: Performed the research work. L.W. Zhang and K.M. Liew: Defined the research direction and supervised the
research work. All authors analyzed the results and wrote the manuscript.

50
W.K. Sun, L.W. Zhang and K.M. Liew Applied Mathematical Modelling 100 (2021) 33–54

Acknowledgments

The authors acknowledge the supports provided by the National Natural Science Foundation of China (Grant No.
11872245) and the Research Grants Council of the Hong Kong Special Administrative Region, China (Project No. 9042644,
CityU 11205518).

Appendix A

For the parameter α , since we will apply a two-step method combining the NNP method and arc-method, we should
make sure both real surface particles and spurious ones are detected, which means the detected surface particles should
be greater than the real amount. Thus, from Fig. a. 1, we can notice that the applicable range of α is 0.8-0.99, however,
0.87-0.95 will be better considering issues of accuracy (NNP tends to make the mistake of detecting an inner particle as a
surface one) and computational cost. Finally, we take 0.88, but other values in 0.87-0.95 can also work out.
For the parameter ε , we studied two simple but typical scenarios, as shown in Fig. a. 2. In both cases, we detect whether
the particle at point O is a surface particle using the arc-method. For a square lattice, since the arcs BD and BE still connect
at B even with the missing particle at B, which makes particle at O an inner particle. For this case, we can obtain ε =ε 1 .
Similarly, we derive ε =ε 2 for the hexagonal lattice scenario. In the central angle method we proposed, we take ε in the
range of ε 1 ∼ε 2 since we notice that any ε less than 1/2 in the square case leads to an inner particle at O while any ε
greater than 2/3 in the hexagonal case leads to a surface particle at O, and particle clusters usually take a form between the
two scenarios in real incompressible fluid simulations. Particularly, we take ε =(ε 1 +ε 2 )/2 in this study.
From numerical tests, on the other hand, the accuracy of the chosen ε can also be verified as shown in Fig. a. 3. As can
be observed, ε 1 gives overestimated results while ε 2 shows the opposite trend. Taking the average of both, however, leads
to results with considerable accuracy compared with the arc-method. Therefore, ε =7/12 is recommended and adopted in
this work.

Fig. A.1. The number of surface particles detected in the dam break sample with varying α .

Fig. A.2. Two standard cases for determination of the parameter ε .

51
W.K. Sun, L.W. Zhang and K.M. Liew Applied Mathematical Modelling 100 (2021) 33–54

Fig. A.3. The number of surface particles detected in the dam break sample with varying ε .

Fig. A.4. A possible 3D extension of the central angle method.

The central angle method and surface tension model we proposed are not directly applicable for 3D simulation. How-
ever, both the ideas of surface detection and surface tension modelling can be introduced into 3D cases with some special
considerations. For the surface detection algorithm, for example, we can study central angles in a specific context. As shown
in Fig. a. 4, if we map neighboring particles of the reference particle on a sphere (see the left panel of Fig. a. 4) and conduct
a triangulation, then we can study the central angle in a specific local context (see the right panel of Fig. a. 4). Presumably,
the local maximal central angle β 1 (before the missing particle lost) should be smaller than the one β 2 (after the missing
particle lost). And there can be a criterion for 3D cases similar to the 2D case. For surface tension modelling, especially,
the curvature calculation, a certain local approximation can also be conducted. However, detailed algorithms can be distinct
from the 2D versions and rigorous study should be carried out before achieving any convincing results and drawing any
solid conclusions, which will be investigated and reported in our future work.

References

[1] J. Ovadnevaite, A. Zuend, A. Laaksonen, K.J. Sanchez, G. Roberts, D. Ceburnis, S. Decesari, M. Rinaldi, N. Hodas, M.C. Facchini, J.H. Seinfeld, C O.D.,
Surface tension prevails over solute effect in organic-influenced cloud droplet activation, Nature 546 (2017) 637–641.
[2] R.W. Style, C. Hyland, R. Boltyanskiy, J.S. Wettlaufer, E.R. Dufresne, Surface tension and contact with soft elastic solids, Nat. Commun. 4 (2013) 2728.
[3] T. Darmanin, F. Guittard, Superhydrophobic and superoleophobic properties in nature, Mater. Today 18 (2015) 273–285.
[4] F. Xia, L. Jiang, Bio-inspired, smart, multiscale interfacial materials, Adv. Mater. 20 (2008) 2842–2858.

52
W.K. Sun, L.W. Zhang and K.M. Liew Applied Mathematical Modelling 100 (2021) 33–54

[5] B.L. Feng, S. Li, Y. Li, H. Li, L. Zhang, Y.S. Jin Zhai, B. Liu, L. Jiang, D. Zhu, Super-hydrophobic surfaces: from natural to artificial, Adv. Mater. 14 (2004)
1857–1860.
[6] H.-J. Chang, M.H. Tsai, W.-S. Hwang, The simulation of micro droplet behavior of molten lead-free solder in inkjet printing process and its experimental
validation, Appl. Math. Model. 36 (2012) 3067–3079.
[7] T. Karbowiak, F. Debeaufort, A. Voilley, Importance of surface tension characterization for food, pharmaceutical and packaging products: a review, Crit.
Rev. Food Sci. Nutr. 46 (2006) 391–407.
[8] S.R. Gonzalez-Avila, E. Klaseboer, B.C. Khoo, C.-D. Ohl, Cavitation bubble dynamics in a liquid gap of variable height, J. Fluid Mech. 682 (2011) 241–260.
[9] P. Cui, A.M. Zhang, S. Wang, B.C. Khoo, Ice breaking by a collapsing bubble, J. Fluid Mech. 841 (2018) 287–309.
[10] F.J. Blanco-Rodríguez, J.M. Gordillo, On the sea spray aerosol originated from bubble bursting jets, J. Fluid Mech. 886 (2020).
[11] J. Zhao, S. Chen, Y. Liu, Spontaneous wetting transition of droplet coalescence on immersed micropillared surfaces, Appl. Math. Model. 63 (2018)
390–404.
[12] A.B. Aljedaani, C. Wang, A. Jetly, S.T. Thoroddsen, Experiments on the breakup of drop-impact crowns by Marangoni holes, J. Fluid Mech. 844 (2018)
162–186.
[13] K. Lippera, M. Morozov, M. Benzaquen, S. Michelin, Collisions and rebounds of chemically active droplets, J. Fluid Mech. 886 (2020).
[14] M.S. Agrawal, H.S. Gaikwad, P.K. Mondal, G. Biswas, Analysis and experiments on the spreading dynamics of a viscoelastic drop, Appl. Math. Model.
75 (2019) 201–209.
[15] S.A. Dyachenko, On the dynamics of a free surface of an ideal fluid in a bounded domain in the presence of surface tension, J. Fluid Mech. 860 (2018)
408–418.
[16] J.J. Monaghan, Smoothed particle hydrodynamics, Annu. Rev. Astron. Astrophys. 1992 (1992) 543–574.
[17] L.B. Lucy, Numerical approach to testing the fission hypothesis, Astron. J. 82 (1977) 1013–1024.
[18] R.A. Gingold, J.J. Monaghan, Smoothed particle hydrodynamics: theory and application to non-spherical stars, Mon. Not. Roy. Astron. Soc. 181 (1977)
375–389.
[19] P.W. Cleary, Y. Serizawa, A coupled discrete droplet and SPH model for predicting spray impingement onto surfaces and into fluid pools, Appl. Math.
Model. 69 (2019) 301–329.
[20] J.J. Monaghan, Simulating free surface flows with SPH, J. Comput. Phys. 110 (1994) 399–406.
[21] J.P. Morris, P.J. Fox, Y. Zhu, Modeling low Reynolds number incompressible flows using SPH, J. Comput. Phys. 136 (1997) 214–226.
[22] J.-Y. Chen, F.-S. Lien, Simulations for soil explosion and its effects on structures using SPH method, Int. J. Impact Eng. 112 (2018) 41–51.
[23] A. Zhang, P. Sun, F. Ming, An SPH modeling of bubble rising and coalescing in three dimensions, Comput. Methods Appl. Mech. Eng. 294 (2015)
189–209.
[24] P. Wang, A.M. Zhang, F. Ming, P. Sun, H. Cheng, A novel non-reflecting boundary condition for fluid dynamics solved by smoothed particle hydrody-
namics, J. Fluid Mech. 860 (2018) 81–114.
[25] M. Olejnik, J. Pozorski, A robust method for wetting phenomena within smoothed particle hydrodynamics, Flow Turbul. Combust. 104 (2020) 115–137.
[26] S. Natsui, K. Tonya, H. Nogami, T. Kikuchi, R.O. Suzuki, SPH simulations of binary droplet deformation considering the Fowkes theory, Chem. Eng. Sci.
(2021) 229.
[27] T. Ma, D. Chen, H. Sun, D. Ma, A. Xu, P. Wang, Dynamic behavior of metal droplet impact on dry smooth wall: SPH simulation and splash criteria, Eur.
J. Mech. B-Fluids 88 (2021) 123–134.
[28] M. Becker, M. Teschner, Weakly compressible SPH for free surface flows, Eurographics/ACM SIGGRAPH Symposium on Computer Animation, 2007.
[29] A. Tartakovsky, P. Meakin, Modeling of surface tension and contact angles with smoothed particle hydrodynamics, Phys. Rev. E 72 (2005) 026301.
[30] J.P. Morris, Simulating surface tension with smoothed particle hydrodynamics, Int. J. Numer. Methods Fluids 33 (20 0 0) 333–353.
[31] X.Y. Hu, N.A. Adams, A multi-phase SPH method for macroscopic and mesoscopic flows, J. Comput. Phys. 213 (2006) 844–861.
[32] S. Adami, X.Y. Hu, N.A. Adams, A new surface-tension formulation for multi-phase SPH using a reproducing divergence approximation, J. Comput. Phys.
229 (2010) 5011–5021.
[33] M.B. Liu, G.R. Liu, Meshfree particle simulation of micro channel flows with surface tension, Comput. Mech. 35 (2004) 332–341.
[34] S.J. Bjorn Andersson, Andreas A. Mark, Fredrik F. Edelvik, Lars L. Davidson, Modeling surface tension in SPH by interface reconstruction using radial
basis functions, 5th international SPHERIC workshop, 2010.
[35] M. Zhang, Simulation of surface tension in 2D and 3D with smoothed particle hydrodynamics method, J. Comput. Phys. 229 (2010) 7238–7259.
[36] T. Breinlinger, P. Polfer, A. Hashibon, T. Kraft, Surface tension and wetting effects with smoothed particle hydrodynamics, J. Comput. Phys. 243 (2013)
14–27.
[37] N. Grenier, D.Le Touzé, A. Colagrossi, M. Antuono, A SPH multiphase formulation with a surface tension model applied to oil-water separation, 4th
SPHERIC Workshop, Nantes, France, 4th SPHERIC Workshop Local Organization Committee, 2009.
[38] J.U. Brackbill, D.B. Kothe, C. Zemach, A continuum method for modeling surface tension, J. Comput. Phys. 100 (1992) 335–354.
[39] X.Y. Hu, N.A. Adams, An incompressible multi-phase SPH method, J. Comput. Phys. 227 (2007) 264–278.
[40] R. Goldman, Curvature formulas for implicit curves and surfaces, Compu. Aided Geom. D. 22 (2005) 632–658.
[41] N. Grenier, M. Antuono, A. Colagrossi, D.Le Touzé, B. Alessandrini, An Hamiltonian interface SPH formulation for multi-fluid and free surface flows, J.
Comput. Phys. 228 (2009) 8380–8393.
[42] F. Yeganehdoust, M. Yaghoubi, H. Emdad, M. Ordoubadi, Numerical study of multiphase droplet dynamics and contact angles by smoothed particle
hydrodynamics, Appl. Math. Model. 40 (2016) 8493–8512.
[43] K. Szewc, A. Taniere, J. Pozorski, J.-P. Minier, A study on application of smoothed particle hydrodynamics to multi-phase flows, Int. J. Nonlinear Sci.
Numer. Simul. 13 (2012) 383–395.
[44] M. Ordoubadi, M. Yaghoubi, F. Yeganehdoust, Surface tension simulation of free surface flows using smoothed particle hydrodynamics, Sci. Iran. 24
(2017) 2019–2033.
[45] M. Tanaka, T. Masunaga, Stabilization and smoothing of pressure in MPS method by Quasi-Compressibility, J. Comput. Phys. 229 (2010) 4279–4290.
[46] S. Koshizuka, Y. Oka, MPS for fragmentation of incompressive fluid, Nucl. Sci. Eng. 123 (1996) 421–434.
[47] E.S. Lee, C. Moulinec, R. Xu, D. Violeau, D. Laurence, P. Stansby, Comparisons of weakly compressible and truly incompressible algorithms for the SPH
mesh free particle method, J. Comput. Phys. 227 (2008) 8417–8436.
[48] M.M. Tsukamoto, L.-Y. Cheng, F.K. Motezuki, Fluid interface detection technique based on neighborhood particles centroid deviation (NPCD) for particle
methods, Int. J. Numer. Methods Fluids 82 (2016) 148–168.
[49] Y. Lin, G.R. Liu, G. Wang, A particle-based free surface detection method and its application to the surface tension effects simulation in smoothed
particle hydrodynamics (SPH), J. Comput. Phys. 383 (2019) 196–206.
[50] G.A. Dilts, Moving least-squares particle hydrodynamics II: conservation and boundaries, Int. J. Numer. Methods Eng. 48 (20 0 0) 1503–1524.
[51] S. Marrone, A. Colagrossi, D.L. Touzé, G. Graziani, Fast free-surface detection and level-set function definition in SPH solvers, J. Comput. Phys. 229
(2010) 3652–3663.
[52] G. Duan, B. Chen, X. Zhang, Y. Wang, A multiphase MPS solver for modeling multi-fluid interaction with free surface and its application in oil spill,
Comput. Methods Appl. Mech. Eng. 320 (2017) 133–161.
[53] P.-P. Wang, Z.-F. Meng, A.M. Zhang, F.-R. Ming, P.-N. Sun, Improved particle shifting technology and optimized free-surface detection method for
free-surface flows in smoothed particle hydrodynamics, Comput. Methods Appl. Mech. Eng. (2019) 357.
[54] T. Ye, D. Pan, C. Huang, M. Liu, Smoothed particle hydrodynamics (SPH) for complex fluid flows: recent developments in methodology and applications,
Phys. Fluids 31 (2019) 011301.

53
W.K. Sun, L.W. Zhang and K.M. Liew Applied Mathematical Modelling 100 (2021) 33–54

[55] S. Marrone, M. Antuono, A. Colagrossi, G. Colicchio, D. Le Touzé, G. Graziani, δ -SPH model for simulating violent impact flows, Comput. Methods Appl.
Mech. Eng. 200 (2011) 1526–1542.
[56] P.N. Sun, A. Colagrossi, S. Marrone, A.M. Zhang, The δ plus-SPH model: Simple procedures for a further improvement of the SPH scheme, Comput.
Methods Appl. Mech. Eng. 315 (2017) 25–49.
[57] P.N. Sun, A. Colagrossi, S. Marrone, M. Antuono, A.M. Zhang, A consistent approach to particle shifting in the δ -Plus-SPH model, Comput. Methods
Appl. Mech. Eng. 348 (2019) 912–934.
[58] P.N. Sun, A. Colagrossi, D.Le L. Touzé, A.M. Zhang, Extension of the δ -Plus-SPH model for simulating Vortex-Induced-Vibration problems, J. Fluids
Struct. 90 (2019) 19–42.
[59] L. Fu, L. Han, X.Y. Hu, N.A. Adams, An isotropic unstructured mesh generation method based on a fluid relaxation analogy, Comput. Methods Appl.
Mech. Eng. 350 (2019) 396–431.
[60] Z. Ji, L. Fu, X. Hu, N. Adams, A consistent parallel isotropic unstructured mesh generation method based on multi-phase SPH, Comput. Methods Appl.
Mech. Eng. 363 (2020) 112881.
[61] Z. Ji, L. Fu, X.Y. Hu, N.A. Adams, A new multi-resolution parallel framework for SPH, Comput. Methods Appl. Mech. Eng. 346 (2019) 1156–1178.
[62] P.W. Cleary, Modelling confined multi-material heat and mass flows using SPH, Appl. Math. Model. 22 (1998) 981–993.
[63] P.W. Cleary, J.J. Monaghan, Conduction modelling using smoothed particle hydrodynamics, J. Comput. Phys. 148 (1999) 227–264.
[64] R. Issa, Numerical Assessment of the Smoothed Particle Hydrodynamics Gridless Method for Incompressible Flows and its Extension to Turbulent
Flows, The University of Manchester (United Kingdom, 2005.
[65] H. Lamb, Hydrodynamics, University Press, 1924.
[66] F. Mashayek, N. Ashgriz, W.J. Minkowycz, B. Shotorban, Coalescence collision of liquid drops, Int. J. Heat Mass Transf. 46 (2003) 77–89.

54

You might also like