Download as pdf or txt
Download as pdf or txt
You are on page 1of 30

Journal of Computational Physics 501 (2024) 112783

Contents lists available at ScienceDirect

Journal of Computational Physics


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

A cell-based smoothed finite element model for the analysis of


turbulent flow using realizable k-ε model and mixed meshes
Mingyang Liu a, b, c, Chen Jiang b, Boo Cheong Khoo c, Huifen Zhu b, Guangjun Gao b, *
a
State Key Laboratory of Mechanical Transmission for Advanced Equipment, College of Mechanical Engineering, Chongqing University, Chongqing
400044, China
b
Key Laboratory of Traffic Safety on Track of Ministry of Education, School of Traffic & Transportation Engineering, Central South University,
Changsha 410075, China
c
Department of Mechanical Engineering, National University of Singapore, 9 Engineering Drive 1, Singapore 117575, Singapore

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

Keywords: Smoothed finite element method (S-FEM) has been widely employed in computational mechanics,
Smoothed finite element method (S-FEM) particularly in computational solid mechanics (CSM) and, to a lesser extent, in computational
Streamline Upwind Petrov-Galerkin (SUPG) fluid dynamics (CFD). The present work focuses on the development of S-FEM for solving three-
Stabilized Pressure Gradient Projection (SPGP)
dimensional incompressible turbulent flow problems using the realizable k-ε model. The
Incompressible flow
Streamline Upwind Petrov-Galerkin approach combined with Stabilized Pressure Gradient Pro­
Turbulent flow
jection (SUPG/SPGP) was used to restrain/control the spatial oscillation and instability in
conjunction with mixed meshes employed to discretize complex geometries. The presented tur­
bulent cell-based S-FEM (CS-FEM) was validated under several common flow types, including
mesh turbulence, turbulent channel flow, turbulent flow past a circular cylinder, flow separation
around an obstacle, flow in the dimpled channel, and flow in the upper human airway. In
particular, the presented work exhibited the latent capability of the CS-FEM in solving turbulent
flows near boundary regions and strong convection regions. Meanwhile, the results of the CS-FEM
were compared to those of the lowest equal-order finite element method (i.e. P1–P1 or Q1–Q1
elements) in simulating turbulent flows, and it was found that the CS-FEM had better agreement
with other published works.

1. Introduction

The Galerkin method is a common method in numerical analysis that applies basis functions to convert continuous operator
problems to discretize problems. Galerkin methods have continued to gain prominence, such as the element-free Galerkin method [1,
2], meshless local Petrov–Galerkin method [3,4], smoothed point interpolation method (S-PIM) [5,6], smoothed finite element method
(S-FEM) [7,8] and so on. Among them, S-FEM is an unconventional Galerkin method and provides a new paradigm shift compared to
the finite element method (FEM) [9]. The gradient smoothing approach originally proposed from Smoothed Particle Hydrodynamic
(SPH) [10] is applied in the framework of the FEM, which modifies the compatible variable field and forms the new method, S-FEM.
Regarding the S-FEM, taking the derivative of the shape function is not essential, which is beneficial for removing the restriction of
isoparametric mapping. S-FEM offers a promising solution to address the issue of overly-stiff problems and can reduce accuracy loss

* Corresponding author.
E-mail address: gjgao@csu.edu.cn (G. Gao).

https://doi.org/10.1016/j.jcp.2024.112783
Received 18 September 2022; Received in revised form 7 January 2024; Accepted 15 January 2024
Available online 17 January 2024
0021-9991/© 2024 Elsevier Inc. All rights reserved.
M. Liu et al. Journal of Computational Physics 501 (2024) 112783

during the solution of distorted meshes. Compared with the FEM, S-FEM promises and delivers a more accurate solution in the field of
solid mechanics [7,11]. It is interesting to note that S-FEM also exhibits good accuracy and convergences on the solution for laminar
flow problems, as recently published [12–14]. Hence, it is not a surprise that S-FEM has emerged as an interesting method to solve the
Navier-Stokes (N-S) equations. The cell-based smoothed finite element method (CS-FEM), as one of the S-FEM variations [15,16], was
the first attempt to solve fluid flow problems among S-FEM variations [11]. Besides this, the CS-FEM provided a simple construction of
N-S equations via a smoothed weak form. Therefore, there is an increasing attention on the potential of CS-FEM for the solution of fluid
dynamics. However, most studies about S-FEM are limited to laminar flow problems. The prediction of irregular turbulent flow, which
is the main difficulty addressed in this work, includes wall treatment and strong convection.
For numerous fluid flow problems, turbulent flow is one typical type of fluid flow, which is significant in both theoretical
development and industrial applications. Turbulent flow is an omnipresent phenomenon, as can be seen in our daily life, such as
separation flow behind a high-speed train [17,18], internal flows in a turbulent channel [19,20], and the flow around buildings [21], to
name a few. Numerical simulation is a widely adopted approach for studying turbulent flow problems due to its distinct advantages,
including cost-effectiveness and short research periods. However, the prediction of turbulent flow is still technically very challenging.
Several numerical approaches have been proposed, for example, direct numerical simulation (DNS) [22], large eddy simulation (LES)
[23], detached eddy simulation (DES) [24], and Reynolds-averaged Navier-Stokes (RANS) equations [25]. RANS equations have
become widespread in engineering applications due to their more affordable computational cost and acceptable computational ac­
curacy. In the traditional RANS equations, the turbulence model can be classified based on the number of governing equations, such as
the one-equation model, two-equation model, etc. The k-ε model has been spotlighted and provided fairly satisfying results in solving
turbulent engineering problems [26,27]. Furthermore, the realizable k-ε model exhibited substantial improvements compared to the
standard k-ε model, mainly where the flow fields include strong streamlined curvature, vortices, and rotation [28]. Subsequently,
considerable research has been conducted on the application of the realizable k-ε model to solve various turbulent flow problems, such
as the flow dynamics of a backward-facing step problem [29], sand colliding on high-speed trains [30], snow depositing on high-speed
trains [31], and many others. In terms of turbulent flow, it has more complex flow characteristics and stronger convection charac­
teristics than laminar flow. Therefore, the performance of S-FEM in solving turbulent flow problems still needs to be further explored,
especially the three-dimensional turbulent flow problems. In this work, we have adopted the realizable k-ε model with the standard
wall functions for the CS-FEM and extended it to complex 3D turbulent flow problems.
To solve the RANS equations, common methods used for solving laminar flows governed by N-S equations are still valid, including
Finite Volume Method (FVM), Finite Difference method (FDM), discontinuous Galerkin (DG) methods, and Finite Element Method
(FEM). Although FVM is still the dominant method in fluid flow simulation, DG method is also a suitable option for turbulent sim­
ulations, as it provides local conservation and a stable discretization of the convective operator [32,33]. DG method is particularly
useful for implementing high-order schemes on complex geometries and unstructured meshes. However, it is important to note that the
DG method requires significantly more computational operations per computational mesh compared to FDM and FVM. FEM also has
an advantage in easily implementing unstructured mesh and high-order approximation. Specifically, FEM only necessitates element
information, whereas FVM also requires information about neighboring mesh elements. As a result, when dealing with unstructured
meshes that lack an explicit method for obtaining neighbor element information, implementing FVM poses a significant programming
challenge. Thus, FEM continues to gain interest and attention in this field.
However, Spatial oscillations have been routinely found when solving the convective term of the N-S equations presenting a
significant challenge for FEM solving fluid problems. In the past, several stabilization methods have been proposed and exhibited
viable performances in solving fluid problems, such as the Characteristic-Based Split (CBS) approach [34–36], Variational Multiscale
(VMS) framework [37,38], Taylor-Galerkin (TG) method [39,40], and Streamline Upwind Petrov-Galerkin (SUPG)[41,42]. The SUPG
scheme has been widely employed to solve a range of flow problems and has demonstrated its effectiveness in various domains,
including incompressible fluid flow and heat transfer problems [43], high Reynolds number compressible turbulent flow problems
[44], hypersonic flow problems [45] and so on. In addition to addressing spatial oscillations, the issue of pressure stability is also
deemed severely detrimental to the accurate solution of incompressible flow using FEM. Therefore, to address the pressure stability, it
is essential to satisfy the Ladyzhenskaya-Babuška-Breezi (LBB) or inf-sup condition. The Artificial Compressibility method [46,47],
Bubble Shape functions [48], Stabilized Pressure Gradient Projection (SPGP) method [49,50], and others have been applied previously
to address the instability problem. The SPGP method presents the latent ability to address pressure instability when solving fluid flow
problems. For example, Wang et al. [51] employed the SPGP scheme to solve incompressible fluid-structure interaction problems.
Furthermore, the SPGP scheme was also used to solve incompressible laminar flows [52] and incompressible fluid-structure interaction
problems [53]. In previous works [13,42], the CS-FEM scheme combined with the SUPG and SPGP (SUPG/SPGP) showed promising
performances in solving incompressible laminar flow. However, compared with laminar flow, the stronger convective characteristics
of turbulence pose a more significant challenge to numerical stability. For the numerical stability method, the behavior in dealing with
complex 3D turbulent flows still need to be discovered and clarified. Therefore, in the present work, we have utilized the CS-FEM
scheme in conjunction with the SUPG/SPGP method to evaluate the performance of S-FEM and the numerical stability method for
solving complex 3D turbulent flows.
The overall structure of this paper has been divided into five sections. Section 2 introduces the semi-implicit SUPG/SPGP approach
and applies the realizable k-ε model for RANS equations in the scheme. Next, the fundamentals of the CS-FEM and the spatial dis­
cretization forms of RANS equations are described in Section 3. Section 4 presents the validation of two-dimensional and three-
dimensional turbulent flow problems along with discussions centered on the performance of the CS-FEM in comparison to the Q1-
Q1 and P1-P1 mixed with Q1-Q1 FEM scheme. Conclusions are given in Section 5.

2
M. Liu et al. Journal of Computational Physics 501 (2024) 112783

2. Formulations

2.1. The semi-implicit SUPG/SPGP scheme for the incompressible RANS equations with realizable k-ε model

The incompressible flow is governed by the mass conservation equation and the momentum equation. The transport equations for
the realizable k-ε model include the turbulent kinetic energy and turbulent dissipation rate equations. These governing equations can
be written as below:
∂ui
= 0, (1)
∂xi
( )
∂ui ∂ui ∂p ∂2 ui 2ρ ∂ ( )
ρ + uj =− + (μ + μt ) − kδij + ρfi , (2)
∂t ∂xj ∂xi ∂xj ∂xj 3 ∂xj
[( ) ]
∂k ∂ ∂ μt ∂k
ρ + ρ (ui k) = μ+ + Gk + Gb − ρε − YM , (3)
∂t ∂xi ∂xi σk ∂xi
[( ) ]
∂ε ∂ ∂ μt ∂ε ε2 ε
ρ + ρ (ui ε) = μ+ + ρC1 Sε − ρC2 √̅̅̅̅̅ + Cε1 Cε3 Gb , (4)
∂t ∂xi ∂xi σ ε ∂xi k + νε k

where ui is the velocity components of the fluid, xi is the Cartesian coordinate, ρ is the fluid density, t is the computing time, p is the
pressure, μ is the laminar viscosity of the fluid, μt is the turbulent viscosity of the fluid, fi is the external body force (including gravity)
and i = 1, 2, 3. Here σ k = 1.0,σε = 1.2, C2 = 1.9, and Cε1 = 1.44 are model coefficients. k is the turbulent kinetic energy, and ε is the
turbulent dissipation rate. Gk and Gb are the generated turbulent kinetic energy associated with velocity gradients and buoyancy,
respectively. YM represents the effect of compressibility on turbulence. For incompressible flow without consideration of heat transfer,
Gb and YM are set as zero. Cε3 can be taken as constant depending on the buoyancy term as follows:
{
1 for Gb ≥ 0
Cε3 = (5)
0 for Gb < 0

And Gk is defined as below:

Gk = μt S2 . (6)
In the above equation, the modulus of the mean strain rate tensor is defined as
( )
√̅̅̅̅̅̅̅̅̅̅̅̅ 1 ∂ui ∂uj
S ≡ 2Sij Sij , Sij = + . (7)
2 ∂xj ∂xi

The turbulent viscosity can be calculated by:

k2
μt = ρCμ fμ , (8)
ε

where Cμ is computed from:


1
Cμ = ∗. (9)
A0 + As kUε

Here,
√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
U∗ = Sij Sij + Ω ̃ ij Ω̃ ij , (10)

and
̃ ij = Ωij − 2εijk ωk , Ωij = Ωij − εijk ωk .
Ω (11)

Here, εijk ωk is the rotating term for a rotating reference frame. This term is neglected because the rotating motion is not considered in
this work. The model constants A0 and As are defined as:
√̅̅̅
A0 = 4.04, As = 6cosφ, (12)

where
1 (√̅̅̅ ) Sij Sjk Ski ̃ √̅̅̅̅̅̅̅̅̅
φ = cos− 1 6W , W = 3
, S = Sij Sij . (13)
3 ̃
S

3
M. Liu et al. Journal of Computational Physics 501 (2024) 112783

The model coefficients C1 can be computed from


η
C1 = max[0.43, ], (14)
η+5

k
η = |S| . (15)
ε
The positivity of k and ε is an important problem in the solution system. The negative or small values of turbulent dissipation rate
would lead to negative or overly large values for turbulent viscosity. The improper sign will have disastrous effects on the solution.
Therefore, both turbulent kinetic energy and turbulent dissipation rate are limited to prevent them from taking negative or overly
small values [54]. If k is tiny or negative, it will be replaced by:
k = kmax /conk , (16)

where kmax is the maximum value observed in the computational domain, and conk is a user-defined constant. For turbulent dissipation
rate ε, it will be replaced by:
/
ε = ρCu k2 μconε . (17)

Here, conε is a user-defined constant. It can limit the lower bound of turbulent viscosity to μconε if the value of ε is negative or tiny. If
turbulent kinetic energy and turbulent dissipation rate are limited, and the turbulent viscosity can be calculated:

(kmax /conk )2
μt = ρCμ fμ / = μfμ conε , (18)
ρCu (kmax /conk )2 μconε

This work mainly used the Neumann and Dirichlet boundary conditions for field variables. For the turbulent kinetic energy, the
Neumann boundary condition was used for the no-slip wall, as stated in Section 2.2. The Dirichlet boundary condition was used for the
velocity inlet and pressure outlet.
The fractional step method [55–57] provided a new form of governing equations. The equal-order interpolation of velocity and
pressure can be achieved in the fractional step method. Meanwhile, it is beneficial to restrain/limit the spatial oscillation of pressure by
solving the symmetric system of linear equations arising from the incompressible condition. The fractional steps of governing equa­
tions in one compute cycle are presented below.

Step 1. Predict the intermediate velocity u∗i


( )
∂ ∂uni 2ρ ∂ ( n ) ∂un ∂pn
ρu∗i − Δt(μ + μt ) + k δij + ρΔtunj i = ρuni − γΔt + Δtρfi . (19)
∂xj ∂xj 3 ∂xj ∂xj ∂xi

Step 2. Predict the pressure pn+1 by solving the following Poisson equation,
Δt ∂ ∂ ( ) ∂ ( )
− pn+1 + γpn δij = − u∗i . (20)
ρ ∂xi ∂xi ∂xi

Step 3. Correct the predicted mean velocity u∗i to get the mean velocity un+1
i in the next time step.
Δt ∂ ( )
un+1
i = u∗i + − pn+1 + γpn δij . (21)
ρ ∂xj

Step 4: Obtain the turbulent kinetic energy kn+1 ,


( ) 2 n
kn+1 − kn ∂kn μ ∂k
ρ + ρuni − μ+ t = Gnk − ρεn . (22)
Δt ∂xi σk ∂xi ∂xi

Step 5: Obtain the turbulent dissipation rate εn+1 ,


( ) 2 n
εn+1 − εn ∂εn μ ∂ε εn ⋅ εn
ρ + ρuni − μ+ t = ρC1 Sεn − ρC2 n √̅̅̅̅̅̅n̅, (23)
Δt ∂xi σε ∂xi ∂xi k + νε

where ν is the kinematic viscosity of the fluid.

4
M. Liu et al. Journal of Computational Physics 501 (2024) 112783

For the above equations, in order to apply the smallest pressure split in the second-order splitting in space, γ is set as 1. However, if
the second-order splitting error is applied, the direct fractional step method will cause the instability problem with the small time step
[58]. The SPGP scheme could address the instability problem in the fractional step method via pressure gradient projection. Mean­
while, the SPGP scheme also performs well in dealing with pressure stability [49]. The SUPG scheme effectively deals with spatial
oscillation issues by introducing the new derivative term, which includes the numerical dissipation in the streamline into a trial
function [39]. To synthesize the advantages of SUPG and SPGP, the SUPG scheme, in conjunction with the SPGP scheme, was used to
handle instability issues in solving turbulent flow using the CS-FEM.
The semi-SUPG/SPGP is combined with the fractional step method and can be derived as follows.
An auxiliary equation for the new variable qi is added to enhance the pressure stability:
∂p
qi − = 0. (24)
∂xi
A new parameter ϕ = ϕt /Δt is proposed [49], in which ϕt is a user-defined parameter. The effect of different values of ϕt has been
discussed and is available in the previous work [13]. In order to add the new variable to the governing equations, the parameter ϕ is
multiplied by the derivative of Eq. (24), which is then substituted into Eq. (1). The resulting form of the new governing equations is
written in the following forms:

∂ui ∂qi ∂2 p
+ϕ − ϕ = 0, (25)
∂xi ∂xi ∂xi ∂xi
( )
∂ui ∂ui ∂p ∂2 ui 2ρ ∂ ( )
ρ + unj =− + (μ + μt ) − kδij + ρfi . (26)
∂t ∂xj ∂xi ∂xj ∂xj 3 ∂xj
The first-order forward difference is used for temporal discretization. The governing equations with the SPGP scheme are written as
follows:

∂un+1 ∂qn+1 ∂2 pn+1


i
+ϕ i − ϕ = 0, (27)
∂xi ∂xi ∂xi ∂xi

un+1 − uni ∂un ∂2 uni 2ρ ∂ ( n ) ∂pn+1 ∂pn ∂pn


ρ i
+ ρunj i − (μ + μt ) + k δij = − + − + ρg i , (28)
Δt ∂xj ∂xi ∂xi 3 ∂xj ∂xi ∂xi ∂xi

∂pn+1
qn+1
i − = 0. (29)
∂xi
According to the fractional step method, the semi-implicit form Eq. (28) can be split as below:

u∗i − uni ∂un ∂2 uni 2ρ ∂ ( n ) ∂pn


ρ + ρunj i − (μ + μt ) + k δij = − + ρg i , (30)
Δt ∂xj ∂xj ∂xj 3 ∂xj ∂xi

un+1 − u∗i ∂pn+1 ∂pn


ρ i
=− + . (31)
Δt ∂xi ∂xi
By taking the divergence for Eq. (31) and substituting Eq. (27) into the new equation, we obtain the new pressure Poisson equation
with the added auxiliary variable:
( Δt) ∂2 pn+1 ∂u∗i Δt ∂2 pn ∂qn+1
ϕ+ = + +ϕ i . (32)
ρ ∂xi ∂xi ∂xi ρ ∂xi ∂xi ∂xi
Next, rearranging the above equations and introducing the SUPG stabilization term, a semi-implicit coupling form of the SUPG/
SPGP in conjunction with the fractional step method, can be presented below.

Step 1: Get the predicted velocity u∗i ,


( )
u∗i − uni ∂un ∂2 uni 2ρ ∂ ( n ) ∂ pn ∂ ∂un ∂pn
ρ + ρunj i − (μ + μt ) + k δij = − + τunk ρunj i + + ρgi . (33)
Δt ∂xj ∂xj ∂xj 3 ∂xj ∂xi ∂xk ∂xj ∂xi
⏟̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅⏞⏞̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅⏟
SUPGstabilizationterm

Step 2: Calculate the pressure pn+1 by solving the following Poisson equation,
( )( ∗ )
∂ ∂ ( n+1 ) ρ ∂uj Δt ∂2 pn ∂qnj
p = + +ϕ . (34)
∂xj ∂xj ρϕ + Δt ∂xj ρ ∂xj ∂xj ∂xj

5
M. Liu et al. Journal of Computational Physics 501 (2024) 112783

Step 3: Correct the predicted velocity u∗i to get the velocity un+1
i for the next time step,

un+1 − u∗i ∂pn+1 ∂pn ∂2 (pn+1 − pn )


ρ i
=− + + τunj . (35)
Δt ∂xi ∂xi ∂xj ∂xi
⏟̅̅̅̅̅̅̅̅̅̅̅̅⏞⏞̅̅̅̅̅̅̅̅̅̅̅̅⏟
SUPGstabilizationterm

Step 4: Renew the auxiliary variable qn+1


i ,

∂pn+1
qn+1
i − = 0. (36)
∂xi

Step 5: Then, obtain the turbulent kinetic energy kn+1 ,


( ) 2 n ( )
kn+1 − kn ∂kn μ ∂k ∂ ∂kn
ρ + ρuni − μ+ t = ρτunj uni + Gnk − ρεn (37)
Δt ∂xi σk ∂xi ∂xi ∂xj ∂xi
⏟̅̅̅̅̅̅̅̅̅̅̅̅̅⏞⏞̅̅̅̅̅̅̅̅̅̅̅̅̅⏟
SUPGstabilizationterm

Step 6: Calculate the turbulent dissipation rate εn+1 ,


( ) 2 n ( )
εn+1 − εn ∂εn μ ∂ε ∂ ∂εn εn ⋅ εn
ρ + ρuni − μ+ t = ρτunj uni + ρC1 Sεn − ρC2 n √̅̅̅̅̅̅n̅. (38)
Δt ∂xi σε ∂xi ∂xi ∂xj ∂xi k + νε
⏟̅̅̅̅̅̅̅̅̅̅̅̅̅⏞⏞̅̅̅̅̅̅̅̅̅̅̅̅̅⏟
SUPGstabilizationterm

Here, τ is the newly added stability coefficient. It may be noted that τis defined for each element and depends on the status of the flows.
In this work, the parameter is calculated as below [25,41]:
{( ) ( )2 [ ]2 }− 1/2
2
2 2‖ u ‖ 4(μ + μt )
τ= + + , (39)
Δt he ρ(he )2

where he is taken as the length of the shortest edge in a mesh.

2.2. Standard wall functions

The presence of walls significantly affects the prediction of turbulent flows. An accurate prediction of the near wall region is central
to the computing accuracy of turbulent flows. Wall functions are the way to model the flow characteristics near the wall region, which
are derived from experiments.
The flow near the wall region is assumed to obey the law of the wall [28]:
⎧ u ρuτ y

⎨ uτ = μ
⎪ y+ ≤ 11.225
. (40)

⎪ u 1 ρu y
⎩ = ln τ + B y+ > 11.225
uτ κ μ
√̅̅̅̅
Here, uτ is the friction velocity, defined as ρ . τw is the shear stress on the wall. y is the wall distance. κ is the von Karman constant, and
τw

B is a roughness constant, set as 0.4187 and 5.45, respectively. For the realizable k-ε model, standard wall functions are employed to
modify the turbulent variables near the wall region. The non-dimensional velocity U+ is defined as u/uτ , and the non-dimensional wall
distance y+ is defined as ρuτ y /μ. Turbulent variable k in the wall-adjacent element is resolved. The boundary condition ∂∂nk = 0 for k is
imposed at the wall. Here, n is the vector perpendicular to the wall.
The generated turbulent kinetic energy Gk and turbulent dissipation rate ε at the wall-adjacent elements are calculated based on the
local equilibrium hypothesis.
∂U τw
Gk ≈ τw = τw , (41)
∂y 1/4
κρCμ k1/2 y

and:

6
M. Liu et al. Journal of Computational Physics 501 (2024) 112783

Cμ3/4 k3/2
ε= (42)
κy

3. Fundamentals of the CS-FEM with realizable k-ε model

3.1. Gradient smoothing

S-FEM synthesizes the advantages of FEM and mesh-free methods. The essential idea in S-FEM is to modify the compatible variable
field or construct a variable gradient field based solely on the variable itself. This approach aims to achieve desirable properties by
using a Galerkin model that incorporates the modified/constructed variable field. The gradient smoothing technique is the key
technique used in S-FEM. This section will introduce the theory and fundamentals of the gradient smoothing technique. The smoothed
variable gradient can be calculated via curves for 2D (or surfaces for 3D) integrations along the boundary of the smoothing domains
using the assumed variable values and the normal components on the boundary [8,59]. The procedure is as follows:
Let Ω be a three-dimensional computational domain. And Ωi is a subdomain of Ω enclosed by an outer boundary Γ, ̃ as shown in
Fig. 1, where n is the unit outward normal of Γ. ̃
For simplicity, we shall assume a field variable f(xi ), which is defined at an arbitrary point xi in the subdomain Ωi . The smoothed
gradient of a field variable f(xi ) can be described in the following way:

̃ (xi ) =
∇f ̃ − xi )dΩ, ∀xi ∈ Ωi ,
∇f (xi )W(x (43)
Ωi

where ∇ is the gradient operator, W ̃ is the smoothed function, and should satisfy some basic conditions. Wis
̃ nonzero only in the
vicinity of x and should be positive over the local smoothing domain. The smoothed function should be centered at x and needs to
satisfy the following condition:

̃ − xi ) ≥ 0, W(x
W(x ̃ − xi )dΩ = 1 (44)
Ωi

And the smoothed function can be computed from:



⎨ 1 , x ∈ Ωi

̃ − xi ) = Vi
W(x . (45)

⎩ 0, x ∕
∈Ω i

We require the smoothing function to vanish only out of Ωi to ensure that the smoothed function is differentiable over Ωi . Here, Vi is the
volume of the smoothing domain (SD).
According to the divergence theorem, Eq. (43) could be rewritten as:
∫ ∫
̃ (xi ) = f (x)W(x
∇f ̃ − xi )ndΓ − ̃ − xi )dΩ.
f (x)∇W(x (46)
Γ̃ Ωi

The second term of Eq. (46) is ignored because the smoothed function is the constant that is related to the volume of the smoothing
domain, as shown in Eq. (45). Thus, the gradient of the smoothed function equals zero, and the second term of Eq. (46) can be
neglected. This leads to:
∫ ∫
̃ − xi )ndΓ = 1 f (x)ndΓ.
̃ (xi ) ≈ f (x)W(x
∇f (47)
Γ̃ Vi Γ̃
To simplify Eq. (46), f(xi ) is changed to the boundary integral. The discrete form of f(xi ) can therefore be written as per the Galerkin
procedure as below:

Fig. 1. Gradient smoothing in a generic 3D smoothing domain.

7
M. Liu et al. Journal of Computational Physics 501 (2024) 112783


f (x) = N(xI )f (xI ), (48)
I∈Gi

where Gi presents the supporting nodes of the SD Ωi , and the definition of supporting nodes can be found in the next section. N(xI ) is the
shape function and will be introduced in the next section. f(xI ) is the field variable value. On substituting Eq. (48) into Eq. (47), one
obtains:
∑( 1 ∫ )
∇f (xi ) ≈ N(xI )ndΓ f (xI ), (49)
I∈Gi
Vi Γ̃

where ∇f(xi ) is the smoothed gradient. The right-hand side of Eq. (49) in SD can be rewritten as:
( )
∑ 1∑
∇f (xi ) ≈ N(xJ )nAJ f (xI ). (50)
I∈Gi
Vi J∈Si

Here, J ∈ Si is the Jth surface of Ωi and AJ is the area of Jth surface. The shape function of Gauss point can be calculated by averaging
the shape functions of the points that belong to the same surface Ωi . The smoothed derivative shape function can be described as:
1∑
∇N(xi ) = N(xJ )nAJ . (51)
Vi J∈Si

3.2. On smoothing domain and smoothed shape functions

The fundamental component of S-FEM is the smoothing domains (SDs). The different types of elements can be assigned different
smoothing domains. The common elements used in three-dimensional problems include the 4-node tetrahedral element (T4), 5-node
pyramid element (P5), 6-node wedge element (W6), 8-node hexahedral element (H8), and so on.
We will briefly introduce the construction of the cell-based smoothing domains (CSDs) in such common elements. The description
of two-dimensional problems is available in the references [12,13,35] and not discussed here. The construction of the first smoothing
domain (SD1) is selected to illustrate how to construct smoothing domains in the CS-FEM. For the T4 element, the tetrahedral

Fig. 2. 3D CSDs for (a) the T4 element based on the CS-FEM (CS-FEM-T4), (b) the P5 element based on the CS-FEM (CS-FEM-P5), (c) the W6
element based on the CS-FEM (CS-FEM-W6) and (d) the H8 element based on the CS-FEM (CS-FEM-H8).

8
M. Liu et al. Journal of Computational Physics 501 (2024) 112783

cell-based smoothing domain (CSD) is the element itself, and the supporting nodes for SD1 are nodes 1 to 4. The P5 element can be
divided into four tetrahedral CSDs and one pyramid CSD; the supporting nodes for its SD1 are nodes 1 to 5. The W6 element can be
constructed as two wedge CSDs; the supporting nodes for its SD1 are nodes 1 to 6. For the H8 element, the pyramid CSD is selected
because it can maintain good accuracy [12], and the supporting nodes for its SD1 are nodes 1 to 8. These are illustrated in Fig. 2.
In the present work, the shape function used in the CS-FEM is called the linear point interpolation method (PIM) [60]. The shape
functions for the arbitrary points on the edge of the element are calculated linearly by two related nodes that bind this edge. For the
interior points in the element, the shape functions can be obtained using the radial point interpolation method (RPIM) [61]. More
detailed information can be found in reference [8]. The descriptions of the shape functions for T4, W6, and H8 elements based on the
CS-FEM can be found in previous studies [12,42]. In this section, we will illustrate the shape functions of the P5 element based on the
CS-FEM. The approximation of shape functions of points in the CS-FEM-P5 is presented in Table 1. The shape functions of CSDs in a P5
element are organized as shown in Table 2.

3.3. On smoothed finite element discretization

In this work, the equal-order FEM with the stabilized method of SUPG was employed. For two-dimensional flow problems, the Q1-
Q1 FEM with the stabilized method of SUPG was employed. For three-dimensional flow problems, the mixed finite elements were used
including P1-P1 and Q1-Q1 finite elements. In the same case, the same equal-order CS-FEM as the FEM was used. In the next, the equal-
order FEM and CS-FEM with the stabilized method of SUPG will be abbreviated as FEM and CS-FEM, respectively.
The discrete forms of field variables and gradients of field variables are constructed as per Galerkin procedure. As described in Eqs.
(48) and (51), the flow velocity, pressure, turbulent kinetic energy, and turbulent dissipation rate are approximated by:

On ∑
On ∑
On ∑
On ∑1
ui = N(xI )ui (xI ) = NI uIi , ∇ui = ∇N(xI )ui (xI ) = uIi NI (xJ )nAJ , (52)
I=1 I=1 I=1 I=1 J∈Si
Vi


On ∑
On ∑
On ∑On ∑1
p= N(xI )p(xI ) = NI pI , ∇p = ∇N(xI )p(xI ) = pI NI (xJ )nAJ , (53)
I=1 I=1 I=1 I=1
V
J∈Si i


On ∑
On ∑
On ∑On ∑1
k= N(xI )k(xI ) = NI kI , ∇k = ∇N(xI )k(xI ) = kI NI (xJ )nAJ , (54)
I=1 I=1 I=1 I=1
V
J∈Si i


On ∑
On ∑
On ∑On ∑1
ε= N(xI )ε(xI ) = NI εI , ∇ε = ∇N(xI )ε(xI ) = εI NI (xJ )nAJ , (55)
I=1 I=1 I=1 I=1
V
J∈Si i

where On is the total number of nodes in the CS-FEM element, and the index I of variables represents the values at node I. The scheme of
the present method is first order in time and second order in space. The semi-SUPG/SPGP discrete forms of Eqs. (33)-(38) can be
reconstructed as in the solution procedure listed as follows:

Step 1. Calculate the predicted velocity (u∗i ) from:



ρMIJ u∗Ji = ρMIJ unJi − Δt(μ + μt )K IJ unJi − ρΔtSIJ unJi − γΔtGIJi pnJ − ΔtGIJj kJn δij
3 (56)
− ρτΔtQIJ unJi − τΔtH IJi pnJ + ΔtρFIi .

The boundary conditions are imposed on Eq. (56); hence the boundary terms are ignored. That is:

Table 1
Values of shape functions at different points for CS-FEM-P5.
Point Node 1 Node 2 Node 3 Node 4 Node 5 Description

1 1 0 0 0 0 Node
2 0 1 0 0 0 Node
3 0 0 1 0 0 Node
4 0 0 0 1 0 Node
5 0 0 0 0 1 Node
6 1/5 1/5 1/5 1/5 1/5 Element center

9
M. Liu et al. Journal of Computational Physics 501 (2024) 112783

Table 2
Shape functions for CS-FEM-P5.
SD Node 1 Node 2 Node 3 Node 4 Node 5

SD 1 6/25 6/25 6/25 6/25 1/25


SD 2 6/20 6/20 1/20 1/20 6/20
CS-FEM-P5 SD3 1/20 6/20 6/20 1/20 6/20
SD4 1/20 1/20 6/20 6/20 6/20
SD5 6/20 1/20 1/20 6/20 6/20

∫ ∫
∂NI ∂NJ
MIJ = NI NJ dΩ, K IJ = dΩ,
Ω Ω ∂xj ∂xj
∫ ( ) ∫ ∫
∂NJ ∂NJ ∂NJ
SIJ = NI uni dΩ, GIJi = NI dΩ, GIJj = NI dΩ, (57)
Ω ∂xi Ω ∂xi Ω ∂xj
∫ ∫ ∫
∂NI ∂NJ ∂NI ∂NJ
QIJ = unk unj dΩ, H IJi = unk dΩ, FIi = NI NJ fJi dΩ,
Ω ∂xk ∂xj Ω ∂xk ∂xi Ω

where [•] is smoothed term and ∂N/∂xi is calculated via Eq. (51).

Step 2: Calculate the pressure (pn+1 ) from:


( ) ( ) ( )
ρ Δt ρϕ
LIJ pn+1 = − GIJi u ∗
+ γ LIJ p n
− G qn , (58)
J
ρϕ + Δt Ji
ρϕ + Δt J
ρϕ + Δt IJi Ji

where

∂NI ∂NJ
LIJ = dΩ. (59)
Ω ∂xi ∂xi

Step 3: Correct for the predicted velocity (u∗i ) to get velocity (un+1
i ) in the next time step from:
[ ] [ n+1 n ]
n+1 ∗ n+1 n
ρMIJ uJi = ρMIJ uJi + ΔtGIJ − pJ + γpJ − τΔtDIJ pJ − pJ , (60)

where

∂NJ ∂NI
DIJ = unj dΩ. (61)
Ω ∂xj ∂xi

Step 4: Calculate the new auxiliary variable (qn+1


i ) from:

MIJ qn+1
J = GIJ pn+1
J . (62)

Step 5: Calculate the turbulent kinetic energy (kn+1 ) from:


( )
μ
ρMIJ kJn+1 = ρMIJ kJn − Δt μ + t KIJk kJn − ρΔtSIJk kJn − ρτΔtQkIJ kJn + Δt[Gk ]nI − ΔtρEIk , (63)
σk

where

10
M. Liu et al. Journal of Computational Physics 501 (2024) 112783

∫ ∫
∂NI ∂NJ
MIJ = NI NJ dΩ, KIJk = dΩ,
Ω Ω ∂xj ∂xj
∫ ( ) ∫
∂NJ ∂NI ∂NJ
SIJk = NI uni dΩ, QkIJ = unk unj dΩ,
∂xi ∂xk ∂xj
Ω
∫ ⃒ ⃒2
Ω
. (64)
⃒ ⃒
[Gk ]nI = μt ⃒Sijn ⃒ NI dΩ,
Ω

EIk = NI NJ εnJ dΩ
Ω

Step 6: Calculate the turbulent dissipation rate (εn+1 ) from:


( )
μt ε n
ρMIJ εn+1
J = ρ MIJ εn
J − Δt μ + KIJ εJ − ρΔtSIJε εnJ − ρτΔtQεIJ εnJ + ΔtρC1 Sijn EIε − ΔtρC2 PεI , (65)
σε

where
∫ ∫
∂NI ∂NJ
MIJ = NI NJ dΩ, KIJε = dΩ
Ω Ω ∂xj ∂xj
∫ ( ) ∫
∂NJ ∂NI ∂NJ
SIJε = NI uni dΩ, QεIJ = unk unj dΩ, (66)
Ω ∂xi Ω ∂xk ∂xj
∫ ∫ n n
ε ⋅ εJ
EIε = EIk = NI NJ εnJ dΩ, PεI = NI NJ n J √̅̅̅̅̅̅̅̅̅̅ dΩ
Ω Ω kJ + ν ⋅ εnJ

4. Numerical examples

4.1. Mesh turbulence test

Mesh turbulence test can be used to describe the decay of homogeneous turbulence. And the analytical solution of turbulent kinetic
energy and turbulent dissipation rate can be given; therefore, this test was also used to analyze the convergences of the present al­
gorithm. There is an analytical solution [62]:
( ) 1
ε0 1− C2
k = k0 1 + (C2 − 1)x1 0 , (67)
k

( )1−CC2
ε0 2
ε = ε0 1 + (C2 − 1)x1 . (68)
k0

However, there is a difference in the governing equation of turbulent dissipation rate between the realizable k-ε model and that of
Mohammadi and Pironneau [62]. In order to assess the convergences of the present algorithm, the same turbulence model as
Mohammadi and Pironneau [62] was used in conjunction with the presented algorithm. k0 and ε0 are the initial values of k and ε,
respectively. x1 presents the horizontal distance. In this test, k0 and ε0 are set as 0.01 m2/s2 and 0.01 m2/s3. The computational domain
is a rectangle with 1.2 m × 0.3 m. The velocity inlet is set as a uniform flow u = {1, 0, 0}T ; the outlet pressure is assumed to be 0 Pa, and
other boundaries are defined as the no-slip wall. In the solution system, we adopted the Lagrange multiplier method to deal with the
pressure boundary. From step 1 to step 6, the prescribed values on the corresponding degree of freedoms (DOFs) on velocity, turbulent
kinetic energy and turbulent dissipation rate boundaries were applied. The boundary condition will be imposed on the corresponding
boundary node. And solving governing equations to obtain new values of field variables.
Four meshes were selected to analyze the corresponding error and test the convergence behaviors of the present algorithm. The

Table 3
The information of four meshes.
Mesh Number of nodes Number of elements have

Mesh-A 21×7 120 0.062 m


Mesh-B 41×7 240 0.044 m
Mesh-C 41×15 560 0.029 m
Mesh-D 61×21 1200 0.020 m

11
M. Liu et al. Journal of Computational Physics 501 (2024) 112783

information on different meshes can be obtained in Table 3. The meshes all consist of quadrilateral elements with a uniform distri­
( 2)
h h
bution. The semi-implicit CBS scheme requires the critical time step to be stable [12]. The requirement is Δt ≤ min |u| , 2ν . In this work,
2
the SUPG formulation is selected for the CS-FEM. Thus, one of the requirements (2hν) could be ignored due to the application of SUPG
formulation. Hence, the Courant number |u| ⋅ Δt/h is considered, and the time steps in the next cases are set according to the Courant
number, which is less than 1.
Fig. 3 gives a comparison of the analytical solution and numerical results with the present algorithm. The numerical result ap­
proaches the analytical solution well, as the mesh is refined. The discrepancy between numerical results and analytical solution focuses
on the rear region of the computational domain. The numerical results show the deeper decay of homogeneous turbulence.
To observe clearly the convergence behaviors of the present algorithm, the error of turbulent kinetic energy k and turbulent
dissipation rate ε will be analyzed by the L2 error norm. The L2 error norm is described below:
ek = ‖ k − kans ‖L2 /‖ kans ‖L2 (69)

eε = ‖ ε − εans ‖L2 /‖ εans ‖L2 (70)

Table 4 gives the corresponding errors of turbulent kinetic energy and turbulent dissipation rate. And the visual description is
shown in Fig. 4. The results exhibit the approximate second order in space of the present algorithm. The error decreases gradually with
the increase of the number of mesh. Moreover, the errors of turbulent kinetic energy and turbulent dissipation rate at Mesh-D are
significantly below 0.002 and 0.004 levels, respectively. The present algorithm could provide good accuracy and have better prospects
in dealing with incompressible turbulent flows.
In an attempt to study the influence of the time step on the convergences for the present algorithm, four different time steps, 0.004
s, 0.002 s, 0.001 s and 0.0005 s were selected to test in the same mesh (Mesh-B). The convergences of turbulent kinetic energy and
turbulent dissipation rate in the different time steps, as calculated by the present algorithm are shown in Table 5 and Fig. 5. The results
show the approximate linear trend of the convergence of the present algorithm over time step. The first order in time of the present
method was exhibited.

4.2. 2D turbulent flow in a channel

The first numerical example is a moderate Reynolds number problem of the turbulent flow in a channel, which also serves to
validate the numerical method. This example will mainly discuss the turbulent flow characteristics near the boundary region. The
Reynolds number is set as 12,300 based on the half-width of the channel. The dynamics viscosity of the fluid is set as 0.0001 Pa ⋅ S. The
channel is set as 2 m wide and 40 m long. The power-law velocity profile is used to describe the shape of the turbulent velocity profile,
which is employed to describe inlet horizontal velocity. That is,
⎧( )1/7

⎪ y

⎪ D− , y > 1.0
u ⎨ D/2
= ( )1/7 , (71)
umax ⎪ ⎪ y


⎩ , y ≤ 1.0
D/2

where D is the width of the channel, and y is the coordinate.


The outlet pressure is assumed to be 0 Pa, and other boundaries are defined as the no-slip wall. The k and ε equal 0.05 m2/s2 and

Fig. 3. A comparison of theoretical and calculated turbulent kinetic energy k and turbulent dissipation rate ε by the CS-FEM versus horizontal
distances. (a) Turbulent kinetic energy k and (b) turbulent dissipation rate ε.

12
M. Liu et al. Journal of Computational Physics 501 (2024) 112783

Table 4
The convergences for the CS-FEM in the grid turbulence test.
Mesh have ek eε log10(have) log10(ek) log10(eε)

Mesh-A 0.062 m 0.0287 0.0345 − 1.20901 − 1.54212 − 1.46218


Mesh-B 0.044 m 0.0055 0.0097 − 1.35952 − 2.25964 − 2.01323
Mesh-C 0.029 m 0.0026 0.0068 − 1.54363 − 2.58503 − 2.16749
Mesh-D 0.020 m 0.0016 0.0038 − 1.70898 − 2.79588 − 2.42022

Fig. 4. The corresponding error and order of convergence obtained by the CS-FEM. (a) The L2 error norm of k and (b) the L2 error norm of ε.

Table 5
The influence of different time step on the convergences for the CS-FEM in the grid turbulence test.
Mesh Time step ek eε log10(time step) log10(ek) log10(eε)

0.004 s 0.0399 0.0581 − 2.39794 − 1.39903 − 1.23582


0.002 s 0.0222 0.032 − 2.69897 − 1.65365 − 1.49485
Mesh-B
0.001 s 0.0114 0.0179 − 3 − 1.9431 − 1.74715
0.0005 s 0.0055 0.0097 − 3.30103 − 2.25964 − 2.01323

Fig. 5. The convergences of the CS-FEM in the different time steps. (a) The L2 error norm of k and (b) the L2 error norm of ε.

0.05 m2/s3 at the inlet [46,63]. Fig. 6 illustrates the problem at hand with the boundary conditions. The initial conditions of k and ε are
identical in value to the boundary conditions. For initial velocity conditions, the horizontal velocity is set as 1.23 m/s, and the vertical
velocity is set as zero.
In an attempt to test the convergence behavior of the CS-FEM and FEM with different meshes, three sets of meshes were

13
M. Liu et al. Journal of Computational Physics 501 (2024) 112783

Fig. 6. The geometry and boundary conditions of the channel.

constructed, and detailed information can be found in Table 6. Here, have is the average size of elements. The Mesh-C consists of
quadrilateral elements with 300×60 nodes, as shown in Fig. 7. In Mesh-C, the size of the first element on the wall along the y direction
is set as 0.01 m. The time step is set as 0.002 s to ensure the Courant number is less than 1. The results obtained from the three meshes
were compared under the same initial conditions and boundary conditions. The fully developed velocity profiles calculated by the CS-
FEM and FEM using the three different meshes are shown in Fig. 8. It indicates that the tendencies of velocity in the three meshes are
consistent. However, the results of Mesh-A exhibit some discrepancies compared to the other two meshes and the experimental data.
The medium mesh (Mesh-B) leads to a decrease in error. Notably, Mesh-C demonstrates good agreement with the experimental data.
To clearly observe the mesh convergence behaviors of the CS-FEM and FEM, the accuracy of horizontal velocity will be analyzed
using the L2 error norm. The L2 error norm is described below:
eux = ‖ ux − uxexp ‖L2 /‖ uxexp ‖L2 (72)

Here, uxexp is the experimental result, and ux is the numerical result. ‖ • ‖L2 represents the L2 norm operator.
The mesh convergence behaviors of the CS-FEM and FEM are shown in Table 7 and Fig. 9. Better convergences of velocity are found
for the CS-FEM and FEM with decreasing element size. The results show that the CS-FEM can improve the accuracy and mesh
convergence rate of the FEM.
The mesh convergence test demonstrates that the resolution of Mesh-C is adequate. Therefore, further discussions and analyses in
this case are conducted based on Mesh-C. The non-dimensional parameter U is defined as u/umax, where u presents the velocity value,
and umax is the maximum value of the velocity. umax is set as 1.23 m/s.
The low-order elements in the form of finite element method not satisfying the inf–sup condition or LBB may also work well by
using the numerical stability method [64]. In order to evaluate the effect of numerical stability method, the results at 4.0 s obtained by
SUPG/SPGP methods and without the numerical stability method are compared in Fig. 10. It is evident from Fig. 9(a) that there are
pressure and spatial oscillation, which adversely affects the numerical stability and result in the failure of the solution. Conversely,
Fig. 9(b) demonstrates good numerical stability and effective performance of SUPG/SPGP in handling incompressible turbulent flow.
Therefore, SUPG/SPGP methods will be used to alleviate pressure and spatial oscillation in other incompressible turbulent flow
problems.
The flow fully developed numerical results calculated by the CS-FEM and FEM are presented in Fig. 11. Similar velocity distri­
butions between the CS-FEM and FEM are observed. The result obtained from the CS-FEM has a relatively larger coverage of high-
speed regions compared to those obtained from the FEM. Concerning turbulent kinetic energy, the larger values near the boundary
of pressure outlet regions calculated by the CS-FEM are observed. In this sense, the CS-FEM may obtain a larger turbulent viscosity in
these regions compared to the FEM.
The turbulent boundary layer and the logarithmic velocity profile are important characteristics of turbulent flows. We depicted the
velocity profiles in the wall unit shown in Fig. 12. The law of the wall is also displayed in Fig. 12, and the detailed definition is available
in Section 2.2. In the viscous sublayer region, the present method exhibits a consistent result with the law of the wall. There is an
interim region between the viscous sublayer and the log-law region. The effects of fluid viscosity and turbulence are equally important
in this region. Meanwhile, the velocity profiles obtained by the present method agree well with the law of the wall. The results of the
FEM and CS-FEM agree to y+<30. The velocity profiles calculated by the present method exhibit an apparent logarithmic law.
Nevertheless, the velocity profile obtained by the CS-FEM is closer to the law of the wall than that of the FEM in the fully turbulent
region. Fig. 12 indicates that the proposed scheme based on the CS-FEM can accurately predict the turbulent flow characteristics near
the wall.
Next, the numerical and experimental results on velocity profiles are compared. Fig. 13 displays the fully developed results,
including the results of the SA model and one-equation by Nithiarasu and Liu [46], experimental data [65], and the results of the high
Reynolds number k-ε model by Lam and Bremhorst [66]. For the two present methods, the fully developed results were obtained on the
pressure outlet of the computational domain. It is apparent from Fig. 13 that the result calculated by the CS-FEM is close to the
experiment. Compared to the FEM, the result of the CS-FEM displayed a lower velocity magnitude near the boundary region. Other
than that, the larger velocity was found in the central region with respect to the result of the CS-FEM. In general, the result of the

Table 6
The information of three meshes.
Mesh Number of nodes Number of elements have

Mesh-A 15 × 75 1036 0.2085 m


Mesh-B 30 × 150 4321 0.1217 m
Mesh-C 60 × 300 17,641 0.0645 m

14
M. Liu et al. Journal of Computational Physics 501 (2024) 112783

Fig. 7. The computational mesh and zoomed view close to the wall.

Fig. 8. A comparison of fully developed velocity profiles in different meshes.

Table 7
A comparison of mesh convergences for CS-FEM and FEM.
eux

CS-FEM FEM

have= 0.2085 m 0.0466 0.0361


have= 0.1217 m 0.0258 0.0318
have= 0.0645 m 0.0151 0.0178

Fig. 9. The mesh convergences of the CS-FEM and FEM.

CS-FEM has a good agreement with the experiment compared to other simulations. That is to say, the realizable k-ε model based on the
CS-FEM exhibits reasonably good performance in predicting the flow characteristics near the boundary region.
Next, we moved to test the CS-FEM and FEM behaviors on solving for the severely distorted mesh. The distortion parameter a was
used to control the distorted circumstance and generate the new mesh, which can be described below:
{ ′
x = x + he ⋅ r ⋅ ax
. (73)
y′ = y + he ⋅ r ⋅ ay

15
M. Liu et al. Journal of Computational Physics 501 (2024) 112783

Fig. 10. The effect of SUPG/SPGP on dealing with the incompressible turbulent flow. (a) With no numerical stability method and (b) with SUPG/
SPGP methods.

Fig. 11. The contours of non-dimensional velocity and turbulent kinetic energy calculated by the CS-FEM and FEM. (a) The velocity, and (b) the
turbulent kinetic energy.

Fig. 12. The velocity profiles obtained by the CS-FEM and FEM in the boundary layer region.

The new mesh was generated by changing the distortion parameter to obtain the new node coordinate. Here, x and y are the
original node coordinates of the regular mesh. r is a random number between − 1 and 1. The distortion parameters ax and ay, were set as
0.5 and 0.25 due to the mesh anisotropy, respectively. A larger distortion parameter represented that a more severely distorted mesh

16
M. Liu et al. Journal of Computational Physics 501 (2024) 112783

Fig. 13. A comparison of fully developed velocity profiles in numerical simulations and the experiment.

would be generated. The distorted mesh for this example is displayed in Fig. 14.
The generated mesh was severely distorted, resulting in some elements losing their regular shape. This numerical example is also
calculated using the distorted mesh, and the fully developed velocity profiles are depicted in Fig. 15. The result of the FEM with the
distorted mesh presents a difference from that of the FEM with the regular mesh. Meanwhile, larger differences between the regular
and distorted mesh are found in the central region. On the other hand, the results obtained from the CS-FEM with the distorted mesh
are still in good agreement with the experimental data. This result indicates that the CS-FEM performs well in dealing with the severely
distorted mesh.

4.3. 2D turbulent flow past a circular cylinder

Turbulent flow past a circular cylinder is another well-known example and has a wide range of industrial applications [67]. This
second example is focused on the unsteady transverse oscillating flow downstream of the circular cylinder. The Reynolds number is set
as 10,000 based on the diameter of the cylinder and upstream inflow velocity. The dynamics viscosity of the fluid is set as 0.0001 Pa ⋅ S.
The diameter of the cylinder is prescribed as 1 m (D). The computational domain is 16 m long and 8 m wide. The center of the cylinder
is on the horizontal centerline of the whole computational domain. The inlet is 4 m away from the center of the cylinder. The uniform
velocity u = 1 m/s is set at the inlet, and the vertical velocity component at the inlet is set to zero. The outlet pressure is assumed to be 0
Pa. The k and ε are equal to 0.0025 m2/s2 and 0.0025 m2/s3 at the inlet, respectively, according to the description of Nithiarasu and Liu
[68]. Other boundaries are defined as a no-slip wall. The initial conditions of k and ε are identical in value to the boundary conditions.
conk was set to 100 to ensure the turbulent kinetic energy is much smaller than the value of the field variable. ε was a variable
related to k, if ε was tiny or negative and needed to be replaced. conε was set to 0.01 to reduce the effect of negative or small values of
the turbulent dissipation rate. Similarly, conε was set to 0.01 to reduce the effect of negative or small values of the turbulent viscosity.
The initial condition of the horizontal velocity is set as 1 m/s, and the vertical velocity is set as zero. The time step is set as 0.005 s.
The apparent geometry information and boundary conditions are shown in Fig. 16.
The computational domain is discretized using 15,688 quadrilateral elements. In order to capture the flow characteristics around
the cylinder and predict the transverse oscillating flow, the mesh in the vicinity of the cylinder is refined. The average size of the first
element along the normal direction of the cylinder is 0.015 m. The computational mesh is displayed in Fig. 17.
The contours of pressure and velocity calculated by the CS-FEM in one cycle are presented in Fig. 18. The vortexes oscillate above
and below the horizontal centerline, and the complete periodic vortex shedding phenomenon is clearly observed; this shows that the
CS-FEM can simulate the transverse oscillation and vortex shedding phenomenon around the circular buff body.
Besides the flow characteristics, we will proceed to compare the drag coefficient Cd ( = 2Fd /ρu2 D), lift coefficient Cl ( = 2Fl /ρu2 D),
and Strouhal number St( = fD/u) to further validate the effectiveness of the present method. Here, Fd is the total drag/horizontal force
on the cylinder, Fl is the total lift/vertical force on the cylinder, and f is the oscillation frequency. The histories of drag and lift co­
efficients calculated by the CS-FEM and FEM are shown in Figs. 19 and 20, respectively. The stabilization of the drag coefficient
calculated by the CS-FEM was established after about 200 s. The steady condition for FEM was observed after about 250 s. The faster
convergence and stabilization observed in the CS-FEM may be attributed to the different approach used for handling elements between

Fig. 14. The zoomed view of the distorted mesh.

17
M. Liu et al. Journal of Computational Physics 501 (2024) 112783

Fig. 15. The results calculated by the regular mesh and distorted mesh.

Fig. 16. The geometry and boundary conditions of turbulent flow past a circular cylinder.

Fig. 17. The mesh and zoomed view of turbulent flow past a circular cylinder.

the CS-FEM and FEM. The discrepancy between the CS-FEM and FEM was found to be 5.7% after stabilization.
The frequency of vortex shedding and periodic flow is presented in Fig. 20. The results calculated by the CS-FEM maintain
satisfactory stability and are more steady than that of FEM. Therefore, the data after 275 s are selected for discussion. The correct
vortex shedding rules could be clearly observed. The mean amplitude of the lift coefficient calculated by the CS-FEM is close to that of
the standard k-ε model by Nithiarasu and Liu [68].
The time evolution of drag and lift coefficients was compared between the present method and reference solutions. We assumed
that oscillation periods start simultaneously, and the results are depicted in Fig. 21. The drag coefficients calculated by the FEM and CS-
FEM correspond to the upper and lower sides of the experiment, respectively. Nevertheless, the result is acceptable because it is closer
to the experimental data than that of Nithiarasu and Liu [68]. A faster oscillation frequency was observed in Nithiarasu and Liu [68], as
shown in Fig. 21(b). However, the Strouhal numbers calculated by the present method are closer to those of the experiment.
Additionally, the present method exhibits consistent results with Nithiarasu and Liu [68] in terms of the lift coefficient amplitude,
with only slight differences observed between the CS-FEM and FEM. Further comparisons of drag and lift coefficients between

18
M. Liu et al. Journal of Computational Physics 501 (2024) 112783

Fig. 18. Contours of velocity and pressure calculated by the CS-FEM in one cycle.

Fig. 19. History of drag coefficient calculated by the CS-FEM and FEM.

numerical simulations and the experiment were conducted, as shown in Table 8. The results show that the present method agrees well
with the experiment. Overall, these results highlight the impressive performance of the present method in predicting the
vortex-shedding phenomenon around the bluff body. Next, we move on to test the behavior of the CS-FEM for 3D turbulent flows.

4.4. 3D turbulent flow past an obstacle

The turbulent flow past an obstacle is another common flow phenomenon, which includes flow separation characteristics.
Therefore, this example was selected to evaluate the performance of the realizable k-ε model based on the CS-FEM to deal with flow

19
M. Liu et al. Journal of Computational Physics 501 (2024) 112783

Fig. 20. History of lift coefficient calculated by the CS-FEM and FEM.

separation. The Reynolds number is 3115, calculated based on the obstacle’s height and the inlet velocity. The dynamic viscosity of the
fluid is set as 0.000018 Pa ⋅ S. The computational domain’s length, height, and width are 1.2 m, 0.3 m, and 0.03 m, respectively. The
obstacle’s height has a characteristic size of 0.04 m, and the length of the obstacle is prescribed as 0.01 m. The inlet is 0.32 m away
from the obstacle. The horizontal uniform velocity is assumed to be 1.17 m/s at the inlet, and the other velocity components are set as
zero. The values of k and ε at the inlet are specified as 0.0024 m2/s2 and 0.07 m2/s3, respectively, according to the description of the
tutorial guide of STAR-CCM+ [70]. The outlet pressure is set as 0 Pa, and the plane z = 0 is defined as a symmetry condition. The
remaining boundaries are defined as no-slip walls. The initial conditions of k and ε are set as 0.0024 m2/s2 and 0.07 m2/s3, respec­
tively, while the initial velocity conditions are identical in value to the velocity inlet conditions. The settings of conk and conε are the
same as the previous example. The time step for the simulation is 0.0002 s. Detailed information on computational configuration and
boundary conditions can be found in Fig. 22.
The large velocity gradient and flow separation are mainly concentrated near the obstacle. Thus, the mesh around the obstacle is
refined. The surface mesh of the computational domain is displayed in Fig. 23. In order to reduce the computing time, the W6 elements
mixed with H8 elements are applied to discretize the computational domain, which consists of 75,832 H8 elements, 14,592 W6 el­
ements, and 78,903 nodes. The average size of the first cell along the normal direction of the wall is 0.002 m.
STAR-CCM+ user guide [70] provides the numerical simulation result calculated by the realizable k-ε model and experimental data
for this case. Fig. 24 shows the velocity and pressure distributions calculated by the CS-FEM and FEM. In the upstream region, the
results of CS-FEM and FEM have a similar distribution, consistent with STAR-CCM+. The result calculated by the CS-FEM in the
downstream region has a smaller velocity distribution in the recirculation region. The vortices in the recirculation region calculated by
the CS-FEM and FEM show similar structures. The differences between the CS-FEM and FEM are found mainly in the downstream
region for the pressure distribution.
Fig. 25 shows the Q-criteria isosurface calculated by the CS-FEM and FEM on the vortex structures. Three vortex structures can be
found in Fig. 25, and two small vortices are found in the upstream and downstream regions. In the downstream region, the vortex
calculated by the CS-FEM is larger than that of the FEM.
Further comparison of the horizontal velocity along a line on the symmetry plane at x = 0.38 m is shown in Fig. 26. In the
recirculation region near the bottom wall, we observe fairly similar results calculated by the CS-FEM and FEM. The magnitude appears
closer to the experiment than that from the STAR-CCM+. The present method and STAR-CCM+ show slight differences in about y =
0.075 m. In the upper region, the velocity profile of the CS-FEM is closer to the experiment and STAR-CCM+ than that of FEM. Overall,
these results indicate that the present method can predict the flow characteristics and obtain acceptable results in dealing with tur­
bulent flow past an obstacle. Furthermore, the CS-FEM performs better than the FEM, as its results are closer to the experiment.

4.5. 3D turbulent flow in the dimpled channel

Turbulent flow in a channel is common in many internal flows. Energy loss, which results in pressure drop, is a perennial issue
prompting a strong motivation to reduce drag. A growing body of literature recognizes the importance of reducing energy loss, and
considerable attention has been paid to employing innovative channel designs such as the dimpled channel [71–73] and others.
Therefore, the turbulent flow in the dimpled channel was selected as the fourth example to investigate the wide range of applications of
the present method. The Reynolds number is prescribed as 5000 based on the height of the channel (H = 1.0 m). The dynamic viscosity
of the fluid is set as 0.0002 Pa ⋅ S. The length and width of the computational domain are set as 16H and 8H, respectively. The diameter
of the dimple (D) is equal to the channel’s height, and the dimple depth ratio to channel height is δ/H = 0.4. As mentioned in Section
4.1, the power-law velocity profile was also used to describe inlet horizontal velocity.

20
M. Liu et al. Journal of Computational Physics 501 (2024) 112783

Fig. 21. The time evolution of drag and lift coefficients.

Table 8
A comparison of the Strouhal number, mean of drag, and amplitudes of lift coefficient values.
St Mean (Cd) Amplitude (Cl)

CS-FEM 0.20 1.16 0.26


FEM 0.20 1.23 0.27
Standard k-ε model by Nithiarasu and Liu [68] 0.24 0.85 0.27
Experimental Data [69] 0.21 1.2 /

21
M. Liu et al. Journal of Computational Physics 501 (2024) 112783

Fig. 22. The computational configuration of turbulent flow past an obstacle.

Fig. 23. The surface mesh and zoomed view of turbulent flow past an obstacle.

⎧( )1/7 /

⎪ z

⎪ 2H − , z > H 2
u ⎨ H/2
= ( )1/7 / . (74)
umax ⎪
⎪ z


⎩ ,z ≤ H 2
H/2

Here, z is the coordinate, and umax is set as 1.0 m/s.


The vertical and transverse velocities are set as zero. The k and ε equal 0.0005 m2/s2 and 0.005 m2/s3 at the inlet. The outlet
pressure is assumed to be 0 Pa. The front and back planes are set as symmetry planes, and the no-slip wall boundary condition is
applied to all the other boundaries. The initial conditions of k and ε are set as 0. 0005 m2/s2 and 0. 005 m2/s3, respectively. The settings
of conk and conε are the same as the previous example. The initial condition of the horizontal velocity is set as umax, and the lateral and
vertical velocities are set as zero. The detailed descriptions of geometry information and boundary conditions are shown in Fig. 27.
The surface mesh of the computational domain and zoomed view of the dimple are shown in Fig. 28. The mesh around the dimple
was refined to improve the computing accuracy and capture the flow characteristics near the dimple region. The computational
domain was discretized using mixed mesh, including T4, P5, and H8 elements. The total elements are 223,048, which consist of
117,444 T4 elements, 9112 P5 elements, 96,492 H8 elements, and 134,085 nodes. The average size of the first element along the
normal direction of the wall is 0.015 m.
The results for CS-FEM and FEM were obtained at 20 s. Fig. 29 presents the numerical results calculated by the CS-FEM and FEM. As
shown in Fig. 28(a), similar velocity distributions are observed in the region located in front of the dimple for both the CS-FEM and
FEM. However, notable differences can be found in the upper boundary region, particularly in the rear region of the dimple, where the
CS-FEM exhibits larger velocity distributions. Meanwhile, the lower velocity distributions can be found near the bottom boundary
compared to the upper boundary region. The low-velocity region and vortex structure are found in the dimpled region with the vortex
structure, as shown in Fig. 29(b). These calculated results are consistent with those of Amsha et al. [72] and Hyun et al. [73], which
were obtained using the non-linear model with k-ε model using the finite volume method (FVM) and experiment, respectively. The
present method can accurately capture the vortex characteristics.
The comparisons of the horizontal velocity along a line on the planes at x=− 0.15 and x = 0.85 are shown in Figs. 30 and 31,
respectively. The intuitive positions of these two lines are available in Fig. 27. The results of the CS-FEM and FEM have a similar trend
around the dimpled region. Meanwhile, the result of the CS-FEM appears closer to the experiment than that of FEM. Around the upper
boundary region, the CS-FEM has the larger velocity magnitude. A comparison of velocity at x = 0.85 in Fig. 31 reveals more sig­
nificant differences near the boundary between the CS-FEM and FEM. For the CS-FEM, the center of the SD is selected to calculate the
wall distance, whereas for the FEM, the wall distance is calculated based on the coordinates of the Gauss point. These different ways of
calculating the wall distances lead to different results, with the results computed by the CS-FEM exhibiting results closer to the
experiment than that of the FEM. Overall, the CS-FEM performs better in solving for the dimpled channel flow compared to the FEM.
The mixed mesh based on CS-FEM can be well applied to discretize the geometry and obtain acceptable results.

22
M. Liu et al. Journal of Computational Physics 501 (2024) 112783

Fig. 24. Contours of velocity and pressure in the symmetry plane.

4.6. 3D turbulent flow in a model upper human airway

The complex three-dimensional turbulent flow in the upper human airway was selected as the final numerical example in this paper
to show the capability of the CS-FEM in dealing with complex turbulent flow. The turbulent flow characteristics in the upper human
airway are crucial for analyzing the mechanisms behind numerous related problems, such as “sleep apnoea” and relative vocal cord
problems. More recently, particle movement in the upper human airway has played a pivotal role in discussions related to diseases
affecting the upper airway [74,75], especially after the outbreak of the COVID-19 pandemic [76,77].
For the simplified model and boundary conditions used in this paper, one can refer to Nithiarasu and Liu [68] and Gemci et al. [78],
see Fig. 32. The inlet velocity is set as 1.64 m/s, and the k and ε are equal to 0.0025 m2/s2 and 0.0025 m2/s3 at the inlet, respectively.
The initial conditions of k and ε are set as 0. 0025 m2/s2 and 0. 0025 m2/s3, respectively. The settings of conk and conε are the same as
the previous example. The initial conditions of all three components of velocity are set to zero. The density of air is assumed to be
1.225 kg/m3, and the dynamic viscosity is taken as 1.8 × 10− 5 N⋅s/m2.

23
M. Liu et al. Journal of Computational Physics 501 (2024) 112783

Fig. 25. Q-criteria isosurface (Q = 0.2) for turbulent flow past an obstacle.

Fig. 26. Comparison of horizontal velocity ux along a line on the symmetry plane at x = 0.38 m.

The surface mesh of the upper human airway and a zoomed view of the boundary layer mesh are displayed in Fig. 33. Mixed meshes
composed of T4, P5, W6, and H8 elements were also applied to discretize the computational domain. The computational domain
consists of 170,996 T4 elements, 5450 P5 elements, 47,632 W6 elements, and 43,600 H8 elements.
The spatial streamlines, calculated using the CS-FEM with mixed mesh, permit the observation of the flow characteristics, as shown
in Fig. 34. Meanwhile, the flow characteristics were also calculated using the FVM software STAR-CCM+. The strong convective flow
phenomenon occurs near the narrow region, which is close to the epiglottis of the upper human airway. The results indicate that the
flow characteristics calculated using the CS-FEM are consistent with those reported by Nithiarasu and Liu [68]. The main recirculation
regions are clustered near the inlet and narrow regions, with vortex formation occurring from the narrow to wide inlet region. The flow
accelerates as it passes through the narrow region, and another vortex is produced in the region transitioning from narrow to wide.
Additionally, a vortex region can be observed in the rear region. The characteristics of recirculation regions are close to that of
STAR-CCM+.
A comparison of the non-dimensional velocity in a characteristic line behind the narrow region was made, as depicted in Fig. 35.
The intuitive position of the characteristic line is available in Fig. 32. The results calculated by the CS-FEM were basically the same as

24
M. Liu et al. Journal of Computational Physics 501 (2024) 112783

Fig. 27. The computational configuration of turbulent flow in the dimpled channel.

Fig. 28. The surface mesh and zoomed view of turbulent flow in the dimpled channel.

Fig. 29. Flow characteristics calculated by the CS-FEM and FEM in the plane at y = 0. (a) Contour of velocity and (b) the streamlines.

STAR-CCM+. Though the high-speed region of STAR-CCM+ is slightly larger than that of the CS-FEM, the differences are still
acceptable. In this sense, the CS-FEM with mixed mesh can predict the flow characteristic and capture the flow structures in the upper
human airway.

25
M. Liu et al. Journal of Computational Physics 501 (2024) 112783

Fig. 30. Comparison of horizontal velocity ux along a line on the plane (y = 0) at x=− 0.15.

Fig. 31. Comparison of horizontal velocity ux along a line on the plane (y = 0) at x = 0.85.

Fig. 32. The simplified upper human airway model and computational boundary conditions.

5. Conclusions

This investigation aims to assess the feasibility and performance of the realizable k-ε model based on CS-FEM in solving 2D and 3D
turbulent flow problems. The feasibility of CS-FEM is convincingly supported by several numerical examples, including mesh tur­
bulence, the prediction of boundary velocity profiles, transverse oscillation around a circular cylinder, flow separation around an
obstacle, flow in the dimpled channel, and flow in the upper human airway. A significant result of this study is the better performance
of the Q1-Q1 CS-FEM in handling turbulent flows near boundary regions and regions with strong convection compared to the Q1-Q1
FEM. Meanwhile, the results show that the Q1-Q1 and P1-P1 mixed with Q1-Q1 CS-FEM could accurately predict the flow

26
M. Liu et al. Journal of Computational Physics 501 (2024) 112783

Fig. 33. The surface mesh and zoomed view of the boundary layer.

Fig. 34. The streamlines calculated by the CS-FEM and STAR-CCM+.

Fig. 35. A comparison of velocity in a characteristic line.

characteristics of the above-mentioned turbulent flow types. Utilizing mixed mesh based on CS-FEM also provided an effective method
to deal with complex 3D turbulent flows. These results have profound implications for understanding how to apply the S-FEM to solve
turbulent flow problems. This work contributes to the expanding field of S-FEM in predicting turbulent flows and develops a numerical
simulation solver based on S-FEM for 3D turbulent flows. Further work will explore the performance of CS-FEM in dealing with
compressible turbulent flows, offering valuable insights in this area.

Credit author statement

Mingyang Liu and Guangjun Gao contributed the central idea, analyzed most of the data, and wrote the initial draft of the paper.
The research direction was provided by Jiang Chen. The remaining authors contributed to refining the ideas, carrying out additional

27
M. Liu et al. Journal of Computational Physics 501 (2024) 112783

analyses and finalizing this paper.

CRediT authorship contribution statement

Mingyang Liu: Data curation, Methodology, Validation, Writing – original draft, Writing – review & editing. Chen Jiang: Funding
acquisition, Methodology, Supervision, Writing – review & editing. Boo Cheong Khoo: Formal analysis, Methodology, Supervision,
Writing – review & editing. Huifen Zhu: Data curation, Formal analysis. Guangjun Gao: Formal analysis, Funding acquisition, Project
administration, Supervision, Writing – review & editing.

Declaration of competing interest

No conflict of interest exits in the submission of this manuscript, and the manuscript is approved by all authors for publication.

Data availability

Data will be made available on request.

Acknowledgement

Authors appreciate the support from the National Natural Science Foundation of China-China State Railway Group Co., Ltd.
Railway Basic Research Joint Fund (Grant No. U2268217), National Natural Science Foundation of China (Grant No. 12002395), The
starting fund for scientific research of Central South University, National Numerical Wind Tunnel Project (NNW2018-ZT1A02) and
China Scholarship Council (Grant No. 202006370122).

References

[1] T. Belytschko, Y.Y. Lu, L. Gu, Element-free Galerkin methods, Int. J. Numer. Methods Eng 37 (1994) 229–256.
[2] H.A. Chowdhury, A. Wittek, K. Miller, G.R. Joldes, An element free Galerkin method based on the modified moving least squares approximation, J. Sci. Comput.
71 (2017) 1197–1211.
[3] E.J. Sellountos, J. Tiago, A. Sequeira, Meshless velocity - vorticity local boundary integral equation (LBIE) method for two dimensional incompressible Navier-
Stokes equations, Int. J. Numer. Method. Heat Fluid Flow 29 (2019) 4034–4073, https://doi.org/10.1108/HFF-06-2018-0310.
[4] N. Sheikhi, M. Najafi, V. Enjilela, Extending the meshless local Petrov–Galerkin method to solve stabilized turbulent fluid flow problems, Int. J. Comput.
Methods 16 (2019) 1850086, https://doi.org/10.1142/S021987621850086X.
[5] G.Y. Zhang, Y. Li, X.X. Gao, D. Hui, Z. Zong, Smoothed point interpolation method for elastoplastic analysis, Int. J. Comput. Methods 12 (2015) 1540013.
[6] G.Y. Zhang, G.R. Liu, Meshfree cell-based smoothed point interpolation method using isoparametric pim shape functions and condensed rpim shape functions,
Int. J. Comput. Methods 08 (2011) 705–730.
[7] G.R. Liu, T.T. Nguyen, K.Y. Dai, K.Y. Lam, Theoretical aspects of the smoothed finite element method (SFEM), Int. J. Numer. Methods Eng. 71 (2010) 902–930.
[8] G.R. Liu, N.T. Trung, Smoothed Finite Element Methods, CRC Press, Boca Raton, 2010. http://books.google.com/books?id=mtub8EjxcsgC&pgis=1.
[9] G.R. Liu, K.Y. Dai, T.T. Nguyen, A smoothed finite element method for mechanics problems, Comput. Mech. 39 (2007) 859–877.
[10] J. Chen, C. Wu, S. Yoon, Y. You, A stabilized conforming nodal integration for Galerkin mesh-free methods, Int. J. Numer. Methods Eng. 50 (2001) 435–466.
[11] C. Jiang, X. Han, G.R. Liu, Z.-Q. Zhang, G. Yang, G.-J. Gao, Smoothed finite element methods (S-FEMs) with polynomial pressure projection (P3) for
incompressible solids, Eng. Anal. Bound Elem. 84 (2017) 253–269, https://doi.org/10.1016/j.enganabound.2017.07.022.
[12] C. Jiang, Z. Zhang, X. Han, G. Liu, T. Lin, A cell-based smoothed finite element method with semi-implicit CBS procedures for incompressible laminar viscous
flows, Int. J. Numer. Methods Fluids 86 (2018) 20–45, https://doi.org/10.1002/fld.4406.
[13] M. Liu, G. Gao, H. Zhu, C. Jiang, A cell-based smoothed finite element method stabilized by implicit SUPG/SPGP/fractional step method for incompressible
flow, Eng. Anal. Bound. Element 124 (2021) 194–210, https://doi.org/10.1016/j.enganabound.2020.12.018.
[14] T. Wang, G. Zhou, C. Jiang, F. Shi, X. Tian, G. Gao, A coupled cell-based smoothed finite element method and discrete phase model for incompressible laminar
flow with dilute solid particles, Eng. Anal. Bound Elem. 143 (2022) 190–206.
[15] T. He, Modeling fluid–structure interaction with the edge-based smoothed finite element method, J. Comput. Phys. 460 (2022) 111171.
[16] G. Zhou, T. Wang, C. Jiang, F. Shi, Y. Wang, L. Zhang, Modeling of particle-laden flows with n-sided polygonal smoothed finite element method and discrete
phase model, Appl. Math. Model. 120 (2023) 355–381.
[17] M. Liu, J. Wang, H. Zhu, S. Krajnovic, G. Gao, A numerical study on water spray from wheel of high-speed train, J. Wind Eng. Ind. Aerodyn. 197 (2020) 104086.
[18] H. Tian, Review of research on high-speed railway aerodynamics in China, Transp. Saf. Environ. 1 (2019) 1–21.
[19] C.M.J. Tay, B.C. Khoo, Y.T. Chew, Use of DES in mildly separated internal flow: dimples in a turbulent channel, J. Turbul. 18 (2017) 1180–1203.
[20] Y. Chen, Y.T. Chew, B.C. Khoo, Enhancement of heat transfer in turbulent channel flow over dimpled surface, Int. J. Heat Mass Transf. 55 (2012) 8100–8121.
[21] M. Atlar, Ö. Gören, others, Effect of turbulence modelling on the computation of the near-wake flow of a circular cylinder, Ocean Eng. 37 (2010) 387–399.
[22] S.A. Orszag, Analytical theories of turbulence, J. Fluid Mech. 41 (1970) 363–386.
[23] J.W. Deardorff, A numerical study of three-dimensional turbulent channel flow at large Reynolds numbers, J. Fluid Mech. 41 (1970) 453–480.
[24] P.R. Spalart, Comments on the feasibility of LES for wings, and on a hybrid RANS/LES approach, in: Proceedings of First AFOSR International Conference on
DNS/LES, Greyden Press, 1997.
[25] H.K. Versteeg, W. Malalasekera, An Introduction to Computational Fluid dynamics: the Finite Volume Method, Second, Perason Education Limited, Harlow,
2007.
[26] M. Shirzadi, P.A. Mirzaei, M. Naghashzadegan, Improvement of k-epsilon turbulence model for CFD simulation of atmospheric boundary layer around a high-
rise building using stochastic optimization and Monte Carlo Sampling technique, J. Wind Eng. Ind. Aerodyn. 171 (2017) 366–379.
[27] L. Yang, J. Zhou, S. Fan, Q. Zheng, H. Zhang, Method and numerical simulation for evaluating the effects of water film on the performance of low-speed axial
compressor, Aerosp. Sci. Technol. 84 (2019) 306–317.
[28] ANSYS FLUENT, ANSYS FLUENT Theory Guide, (2010).
[29] L. Han, N. Riviere, M. Chatelain, E. Mignot, Recirculation zone downstream lateral expansions of open channel flow, Phys. Fluids 32 (2020) 115119.
[30] J. Wang, D. Liu, G. Gao, Y. Zhang, J. Zhang, Numerical investigation of the effects of sand collision on the aerodynamic behaviour of a high-speed train
subjected to yaw angles, J. Appl. Fluid Mech. 12 (2019) 379–389.

28
M. Liu et al. Journal of Computational Physics 501 (2024) 112783

[31] M. Liu, J. Wang, H. Zhu, S. Krajnovic, G. Gao, A numerical study of snow accumulation on the bogies of high-speed trains based on coupling improved delayed
detached eddy simulation and discrete phase model, Proc. Inst. Mech. Eng., Part F: J. Rail Rapid Transit 233 (2019) 715–730.
[32] N. Cuong Nguyen, S. Terrana, J. Peraire, Large-eddy simulation of transonic buffet using matrix-free discontinuous Galerkin method, AIAA J. 60 (2022)
3060–3077.
[33] P. Fernandez, N.C. Nguyen, J. Peraire, The hybridized discontinuous Galerkin method for implicit large-eddy simulation of transitional turbulent flows,
J. Comput. Phys. 336 (2017) 308–329.
[34] O.C. Zienkiewicz, R. Codina, A general algorithm for compressible and incompressible flow—Part I. the split, characteristic-based scheme, Int. J. Numer.
Methods Fluids 20 (1995) 869–885, https://doi.org/10.1002/fld.1650200812.
[35] M. Liu, G.J. Gao, H. Zhu, C. Jiang, G. Liu, A cell-based smoothed finite element method for arbitrary polygonal element to solve incompressible laminar flow,
Int. J. Comput. Methods (2021) 1–28, https://doi.org/10.1142/S0219876221500171.
[36] T. He, An efficient selective cell-based smoothed finite element approach to fluid-structure interaction, Phys. Fluids 32 (2020) 067201.
[37] N. Ahmed, S. Rubino, Numerical comparisons of finite element stabilized methods for a 2D vortex dynamics simulation at high Reynolds number, Comput.
Methods Appl. Mech. Eng. 349 (2019) 191–212.
[38] S.M. Modirkhazeni, V.G. Bhigamudre, J.P. Trelles, Evaluation of a nonlinear variational multiscale method for fluid transport problems, Comput. Fluids 209
(2020) 104531.
[39] D.M. Hawken, H.R. Tamaddon-Jahromi, P. Townsend, M.F. Webster, A Taylor–Galerkin-based algorithm for viscous incompressible flow, Int. J. Numer.
Methods Fluids 10 (2010) 327–351.
[40] J. Donea, A Taylor–Galerkin method for convective transport problems, Int. J. Numer. Methods Eng 20 (1984) 101–119, https://doi.org/10.1002/
nme.1620200108.
[41] T.J. Hughes, M. Mallet, M. Akira, A new finite element formulation for computational fluid dynamics: II. Beyond SUPG, Comput. Methods Appl. Mech. Eng. 54
(1986) 341–355.
[42] M. Liu, G. Gao, H. Zhu, C. Jiang, G. Liu, A cell-based smoothed finite element method (CS-FEM) for three-dimensional incompressible laminar flows using mixed
wedge-hexahedral element, Eng. Anal. Bound. Elem. 133 (2021) 269–285.
[43] H.G. Choi, H. Choi, J.Y. Yoo, A fractional four-step finite element formulation of the unsteady incompressible Navier-Stokes equations using SUPG and linear
equal-order element methods, Comput. Methods Appl. Mech. Eng. 143 (1997) 333–348, https://doi.org/10.1016/S0045-7825(96)01156-5.
[44] C. Wervaecke, H. Beaugendre, B. Nkonga, A fully coupled RANS Spalart-Allmaras SUPG formulation for turbulent compressible flows on stretched-unstructured
grids, Comput. Method. Appl. Mech. Eng. 233–236 (2012) 109–122.
[45] D. Codoni, C. Johansen, A. Korobenko, A Streamline-Upwind Petrov–Galerkin formulation for the analysis of hypersonic flows in thermal non-equilibrium,
Comput. Methods Appl. Mech. Eng. 398 (2022) 115185.
[46] P. Nithiarasu, An efficient artificial compressibility (AC) scheme based on the characteristic based split (CBS) method for incompressible flows, Int. J. Numer.
Methods Eng. 56 (2003) 1815–1845, https://doi.org/10.1002/nme.712.
[47] X. He, G.D. Doolen, T. Clark, Comparison of the lattice boltzmann method and the artificial compressibility method for Navier–Stokes equations, J. Comput.
Phys. 179 (2002) 439–451.
[48] D.Z. Turner, K.B. Nakshatrala, K.D. Hjelmstad, On the stability of bubble functions and a stabilized mixed finite element formulation for the Stokes problem, Int.
J. Numer. Methods Fluids 60 (2010) 1291–1314.
[49] R. Codina, J. Blasco, Stabilized finite element method for the transient Navier–Stokes equations based on a pressure gradient projection, Comput. Methods Appl.
Mech. Eng. 182 (2000) 277–300, https://doi.org/10.1016/S0045-7825(99)00194-2.
[50] R. Codina, J. Blasco, G.C. Buscaglia, A. Huerta, Implementation of a stabilized finite element formulation for the incompressible Navier–Stokes equations based
on a pressure gradient projection, Int. J. Numer. Methods Fluids 37 (2010) 419–444.
[51] G. Wang, Y. Hong, S. Huo, C. Jiang, An immersed edge-based smoothed finite element method with the stabilized pressure gradient projection for
fluid–structure interaction, Comput. Struct. 270 (2022) 106833.
[52] M. Liu, G. Gao, B.C. Khoo, Z. He, C. Jiang, A cell-based smoothed finite element model for non-Newtonian blood flow, Appl. Math. Comput. 435 (2022) 127480.
[53] T. He, Stabilization of a smoothed finite element semi-implicit coupling scheme for viscoelastic fluid–structure interaction, J. Nonnewton Fluid Mech. 292
(2021) 104545.
[54] F. Ilinca, D. Pelletier, Positivity preservation and adaptive solution for the k-ε model of turbulence, AIAA J. 36 (1998) 44–50.
[55] J. Kim, P. Moin, Application of a fractional-step method to incompressible Navier-Stokes equations, J. Comput. Phys 59 (1985) 308–323.
[56] K. Liu, R.H. Pletcher, A fractional step method for solving the compressible Navier–Stokes equations, J. Comput. Phys. 226 (2007) 1930–1951.
[57] G.K. Despotis, S. Tsangaris, Fractional step method for solution of incompressible Navier-Stokes equations on unstructured triangular meshes, Int. J. Numer.
Methods Fluids 20 (2010) 1273–1288.
[58] P. Nithiarasu, O. Zienkiewicz, Analysis of an explicit and matrix free fractional step method for incompressible flows, Comput. Methods Appl. Mech. Eng. 195
(2006) 5537–5551.
[59] W. Zeng, G.R. Liu, Smoothed finite element methods (S-FEM): an overview and recent developments, Arch. Comput. Method. Eng. 25 (2018) 397–435, https://
doi.org/10.1007/s11831-016-9202-3.
[60] K.Y. Dai, G.R. Liu, T.T. Nguyen, An n-sided polygonal smoothed finite element method (nSFEM) for solid mechanics, Finite Element. Anal. Design 43 (2007)
847–860, https://doi.org/10.1016/j.finel.2007.05.009.
[61] G.R. Liu, Meshfree Methods: Moving Beyond the Finite Element Method, 2009th ed., CRC Press, Boca Raton,USA, 2009.
[62] B. Mohammadi, O. Pironneau, Analysis of the K-Epsilon Turbulence Model, John Wiley & Sons, Ltd, 1993.
[63] H. Stefan, Lars Johansson, Erik Davidson, Olsson, Numerical simulation of vortex shedding past triangular cylinders at high Reynolds number using a k-ε
turbulence model, Int. J. Numer. Methods Fluids 16 (1993) 859–878.
[64] J. Li, Y. He, Z. Chen, A new stabilized finite element method for the transient Navier–Stokes equations, Comput. Methods Appl. Mech. Eng. 197 (2007) 22–35.
[65] J. Laufer, Investigation of turbulent flow in a two-dimensional channel, 1951.
[66] C. Lam, K. Bremhorst, A modified form of the k-ε model for predicting wall turbulence, J. Fluids Eng 103 (1981) 456–460, https://doi.org/10.1115/1.3240815.
[67] S. Peng, Y.-L. Xiong, X.-Y. Xu, P. Yu, Numerical study of unsteady viscoelastic flow past two side-by-side circular cylinders, Phys. Fluids 32 (2020) 083106.
[68] P. Nithiarasu, C.B. Liu, An artificial compressibility based characteristic based split (CBS) scheme for steady and unsteady turbulent incompressible flows,
Comput. Methods Appl. Mech. Eng. 195 (2006) 2961–2982, https://doi.org/10.1016/j.cma.2004.09.017.
[69] H. Schlichting, K. Gersten, Boundary-Layer Theory, Springer Berlin Heidelberg, 2017.
[70] Siemens, STAR-CCM+ User’s Guide, Siemens Inc, 2014.
[71] C. Tay, B. Khoo, Y. Chew, Mechanics of drag reduction by shallow dimples in channel flow, Phys. Fluids 27 (2015) 035109.
[72] K.A. Amsha, T. Craft, H. Iacovides, Computational modelling of the flow and heat transfer in dimpled channels, Aeronautic. J. 121 (2017) 1066–1086.
[73] H.G. Kwon, S.D. Hwang, H.H. Cho, Measurement of local heat/mass transfer coefficients on a dimple using naphthalene sublimation, Int J Heat Mass Transf 54
(2011) 1071–1080.
[74] E. Frederix, A.K. Kuczaj, M. Nordlund, M. Bělka, F. Lizal, J. Jedelskỳ, J. Elcner, M. Jícha, B. Geurts, Simulation of size-dependent aerosol deposition in a realistic
model of the upper human airways, J. Aerosol Sci. 115 (2018) 29–45.
[75] N.L. Phuong, N.D. Khoa, K. Ito, Comparative numerical simulation of inhaled particle dispersion in upper human airway to analyse intersubject differences,
Indoor Built Environ. 29 (2020) 793–809.

29
M. Liu et al. Journal of Computational Physics 501 (2024) 112783

[76] J.K. Mutuku, W.-C. Hou, W.-H. Chen, An overview of experiments and numerical simulations on airflow and aerosols deposition in human airways and the role
of bioaerosol motion in COVID-19 transmission, Aerosol Air Qual. Res. 20 (2020) 1172–1196.
[77] H. Mortazavi, H.M. Beni, F. Aghaei, S.H. Sajadian, SARS-CoV-2 droplet deposition path and its effects on the human upper airway in the oral inhalation,
Comput. Methods Programs Biomed. 200 (2021) 105843.
[78] T. Gemci, T. Corcoran, K. Yakut, B. Shortall, N. Chigier, Spray Dynamics and Deposition of Inhaled Medications in the Throat, ILASS-Europe, Zurich, 2001,
2001.

30

You might also like