You are on page 1of 22

Available online at www.sciencedirect.

com
ScienceDirect

Comput. Methods Appl. Mech. Engrg. 373 (2021) 113438


www.elsevier.com/locate/cma

A semi-implicit immersed boundary method for simulating viscous


flow-induced sound with moving boundaries
Long Chenga,b , Lin Dua ,∗, Xiaoyu Wanga , Xiaofeng Suna , Paul G. Tuckerb
a School of Energy and Power Engineering, Beihang University, Beijing 100191, China
b Department of Engineering, University of Cambridge, Cambridge CB2 1PZ, UK

Received 27 March 2020; received in revised form 31 July 2020; accepted 10 September 2020
Available online xxxx

Abstract
In this paper, a semi-implicit immersed boundary body force model is derived from the compressible Navier–Stokes
equations, to directly predict the viscous flow-induced sound from moving objects on a fixed Cartesian grid. To overcome the
conflict of grid quality with efficiency in simulating moving-boundary problems with high-order computational aeroacoustics
methods, a prediction–correction technique is utilized. This accurately satisfies no-slip wall boundary conditions at every time
step without any feedback treatment. A particular contribution of the work is the introduction of a numerical model equation
to analyze the body force convergence. This is useful to pre-evaluate the generated Cartesian and body-surface grids. Several
benchmark aeroacoustic problems are simulated to validate the present model. Results show that the unsteady force and far-field
sound directivity agree well with the previous direct numerical simulation results. The work further suggests that the developed
body force model/CAA methods are capable of predicting interaction noise, especially those associated with oscillating multiple
objects.
⃝c 2020 Elsevier B.V. All rights reserved.

Keywords: Semi-implicit; Body force; Prediction–correction; CAA; Moving boundaries

1. Introduction
Flow-induced noise is usually viewed as an undesirable byproduct in engineering applications, especially in
the aircraft industry. Its prediction [1] has undergone great development since 1952, with improved physical
acquaintance of sound sources and the development of numerical methods and high-performance computers.
However, aerodynamic noise generation and propagation mechanisms become more and more complex in modern
aircraft. This mainly reflects in the strong interaction with aeroelasticity [2]. The introduction of aeroelastic effects
not only increases the complexity of predicting aerodynamic performances [3], but also changes sound sources and
their behavior in non-uniform mean flows. This is because the motion effects of wall boundaries are equivalent to
a category of boundary inhomogeneity on the fluid, which is usually the source or sink of vorticity and sound. To
some extent, how to predict flow-induced sound from moving boundaries is one of the most challenging topics in
aeroacoustics communities [4].
∗ Corresponding author.
E-mail address: lindu@buaa.edu.cn (L. Du).

https://doi.org/10.1016/j.cma.2020.113438
0045-7825/⃝ c 2020 Elsevier B.V. All rights reserved.
L. Cheng, L. Du, X. Wang et al. Computer Methods in Applied Mechanics and Engineering 373 (2021) 113438

Hybrid computation and direct computation are two basic classifications for flow-induced sound prediction
methods [5,6]. The representative theory for the first category of methods is Lighthill’s acoustic analogy theory [1]
and the Ffowcs-Williams–Hawkings (FW–H) equation [5], in which the sources should be known in advance. The
other kind of methods resolves different flow scales on the basis of physical problems, such as direct numerical
simulation (DNS) [6]. Recent DNS studies showed that different motion modes of immersed bodies are able to
exert a remarkable influence on both near-field flow structures and far-field sound directivities [7,8]. The findings
indicate that DNS is a powerful tool to perform principle validation [9]. The current direct computation methods
are usually based on well-optimized body-fitted grids. To some degree, they are usually suitable to treat simple
geometries [10] and simple motion modes [7]. This is due to the conflict in grid quality and generation efficiency,
when low dissipation and low dispersion high-order numerical schemes in computational aeroacoustics (CAA) are
adopted.
Unlike the direct imposition of wall boundary conditions in body-fitted-grid-based methods, the immersed
boundary (IB) method, proposed by Peskin [11], converts the fluid–solid interface (FSI) into a category of flow
discontinuity which is strictly described using generalized body force sources in momentum equations. It has
boasted a dominant development in simulating moving-boundary problems [12–14]. However, when it comes to
the compressible flow computation [15], there is very limited relevant literature especially for flow-induced sound
problems. This is possibly due to the lack of suitable body force model when high-order methods are utilized
in high-speed flow or sound wave simulations [16,17]. To the authors’ knowledge, three typical categories of
IB methods have been developed for flow-induced sound or acoustic scattering problems, that is, the ghost-cell
(GC) method [17], the penalization-type methods [18,19], and the influence matrix method (IMM) [16]. The sharp-
interface quality is kept in the GC method via high-order interpolation near wall boundaries, and no explicit body
force models are introduced, which, on the contrary, leads to the inevitable adoption of asymmetric spatial schemes
and filtering/artificial damping stencils near wall boundaries. From the stability analysis in [20,21], the irregularities
of wall boundaries and the necessity of adopting one-sided high-order spatial scheme near wall boundaries make it
prone to excite spurious short waves, which can not only pollute the solution, but also lead to numerical instability
in some circumstances. This solution degradation was observed in [17], but it is worth noting that this phenomenon
is quite common in the dispersive finite difference method (FDM). Trefethen [22] analyzed the group velocity
of FDM, finding that the anisotropy of finite difference grid is partly accountable for the parasitic waves. Hence,
whether for the Cartesian method [20] or GC method [17], how to tackle the parasitic waves and numerical stability
caused by the adoption of asymmetric spatial schemes and filtering/artificial damping stencils near wall boundary
is quite challenging.
By contrast, both the penalization method and IMM inherit the original concept of the IB method for the body
force model to indirectly satisfy the wall boundary conditions. The central high-order spatial schemes are used in
the whole computational domain even near FSI. From the singularity theory derived by Sirovich [23,24], the FSI
is equivalent to a category of flow discontinuity represented by the singular body force terms in discontinuous N–S
equations. The theory is possibly able to better associate the sources of parasite waves near wall boundaries with
discontinuity-induced oscillations [22] in high-order difference approximations. Particularly, the penalization method
is related to two empirical parameters, so accurately choosing these parameters is significant to get a convergent
solution, and it was initially developed to simulate flow-induced sound from stationary objects. Recently, Komatsu
et al. [25] improved the penalization method by correcting the internal energy equation in consideration of the body
force work, thus making the method capable of predicting the flow-induced sound from moving boundaries. By
comparison, the IMM was originally put forward by Su and Lai [26] for incompressible flow simulation, while
Sun et al. [16] and Cheng et al. [27,28] extended it to simulate linear acoustic problems. No empirical parameters
are needed in the IMM. Also, unlike the GC method and the penalization method, it is unnecessary to distinguish
the different sides of the flow discontinuity. Hence, the interpolation and distribution operations for the body force
are isotropic and are fulfilled more efficiently with IMM. In addition, the allowed CFL number is much larger
than the classical feedback IB methods [11,29]. However, the previous studies [16,27] are mainly related to linear
problems with stationary objects. Besides, the time-consuming singular value decomposition (SVD) is used to solve
the boundary force equation with large stiffness. The present work aims at improving the IMM to directly simulate
nonlinear flow-induced sound from moving boundaries with high-order CAA methods. The improved IMM for
compressible Navier–Stokes (N–S) equations adds body force sources into the momentum equations to satisfy the
no-slip wall boundary conditions. The high-efficiency preconditioned conjugate gradient method (PCGM) [30],
2
L. Cheng, L. Du, X. Wang et al. Computer Methods in Applied Mechanics and Engineering 373 (2021) 113438

Fig. 1. Schematic of immersed boundary method framework for moving-boundary problems defined on fixed Cartesian grid and Lagrangian
boundary grid (Γ: Lagrangian curvilinear boundary or the fluid–solid interface (FSI); Ω1 : fluid subdomain; Ω2 : solid/fictitious fluid subdomain;
f: body force density; dx, ∆sk : Background mesh size and the Lagrangian mesh size of the kth segment of boundary Γ; A, B: boundary
forcing points on Γ; blue and green square boxes: the influence regions of points, A and B). . (For interpretation of the references to color
in this figure legend, the reader is referred to the web version of this article.)

rather than SVD, is developed to solve the boundary force equation. To eliminate the possible parasite waves near
the flow discontinuity as mentioned above, different optimized distribution functions [31] are tested. Particularly,
much effort is made to analyze the body force convergence for the first time.
The paper is organized as follows. In Section 2, the IMM for the compressible N-S equations with moving
boundaries and associated high-order finite difference schemes are presented. The convergence and stability of the
IMM is numerically analyzed. In Section 3, several benchmark flow-induced sound cases are investigated to verify
the capability and accuracy of the developed computational model in dealing with flow-induced sound from moving
objects with complex configurations.

2. Computational models
The schematic of the model problem is illustrated in Fig. 1. The IB method [11] employs a mixture of Eulerian
and Lagrangian variables to define the flow field and FSI, so two sets of grids are generated: a set of Eulerian
background mesh for domain Ω = Ω1 + Ω2 and a set of Lagrangian boundary mesh for the FSI Γ in Fig. 1.
Correspondingly, physical quantities are called Lagrangian variables if they are functions of the Lagrangian mesh
coordinate, X = (X, Y), on FSI Γ; otherwise, they are Eulerian variables and are functions of the fixed Cartesian
mesh coordinate, x = (x, y). It is noteworthy that the X = (X, Y) is able to be viewed as a time-dependent mapping
from the Lagrangian material coordinate, X = (ζ, t), of the FSI Γ to the fixed Cartesian position on domain Ω, where
the material coordinate, X = (ζ, t), is a widely-adopted definition in Peskin’s work [11,32] if elastic boundaries are
considered. Hence, we know x ∈ Ω, X ∈ Γ and Γ ⊂ Ω. Also, the body force density, f(x), is an Eulerian variable,
while the boundary force density, F(X), introduced on the FSI Γ in Fig. 1 is a Lagrangian variable. Besides, in IB
method [11], the numerical discretization of the governing equations is conducted on the whole domain, Ω. The FSI
Γ is discretized into several boundary forcing points to calculate the boundary force density, F(X), and the body
force density distribution on the Eulerian background mesh points. When considering moving boundaries, only the
Lagrangian boundary mesh moves freely on the Eulerian background mesh accompanied by a time-varying body
force density, f(x). The objects are assumed with no deformation, so only the Lagrangian coordinate, X = (X, Y),
rather than X = (ζ, t) is utilized.

2.1. Governing equations

Flow-induced sound problems can be described by the compressible N–S equations in the non-dimensional
conservative form of
Qt + Ex + G y = S0 , (1)
where the conservative variables are
Q = [ρ, ρu, ρv, E]T , (2)
where E = ρ(u2 +v 2 )/2 + p/(γ −1) is the total energy,
3
L. Cheng, L. Du, X. Wang et al. Computer Methods in Applied Mechanics and Engineering 373 (2021) 113438

and flux vectors are


ρu
⎡ ⎤

ρu 2 + p − τx x
⎢ ⎥
⎢ ⎥
E=⎢ ⎥, (3)
⎢ ⎥

⎣ ρuv − τ xy


(E + p) u − uτx x − vτx y + qx
ρv
⎡ ⎤

ρuv − τx y
⎢ ⎥
⎢ ⎥
G=⎢ ⎥, (4)
⎢ ⎥

⎣ ρv 2
+ p − τ yy


(E + p) v − uτx y − vτ yy + q y
with viscous stress terms written as
∂u ∂u ∂v
[ ( )]
Ma∞
τx x = 2µ −λ + , (5)
Re∞ ∂x ∂x ∂y
∂v ∂u ∂v
[ ( )]
Ma∞
τ yy = 2µ −λ + , (6)
Re∞ ∂y ∂x ∂y
∂u ∂v
( )
Ma∞
τx y = µ + , (7)
Re∞ ∂x ∂y
and heat transfer terms of
Ma∞ ∂T Ma∞ ∂T
qx = − µ , qy = − µ , (8)
(γ − 1)Pr Re∞ ∂ x (γ − 1)Pr Re∞ ∂ y
where µ is the dynamic viscosity, calculated by Sutherland’s Law [33]. Also, λ = 2µ/3 is the bulk viscosity. Finally,
γ p = ρT is the equation of state for perfect gas, and the source terms can be written as
]T
S0 = 0, f x , f y , Q, W ,
[
(9)
where “T ” represents transposition operation. u = (u,v) denotes flow velocity. Also, the Eulerian variable, f =
(fx , f y ), is the body force density same as the above definition in Fig. 1; S0 represents the source term; Q is the
heat source, and W = f • uw is the work done by the body force, where the Eulerian variable, uw (x), is same as
the moving velocity, Uw (X), of the boundary forcing point, X, on FSI Γ in Fig. 1. When a stationary object is
considered, the work W is definitely zero. All quantities are normalized. Specifically, the velocity is normalized by
the inflow sound speed c∞ , density by the inflow static density ρ∞ , time by D/c∞ , where D is the characteristic
length of the object immersed in the Eulerian background mesh in Fig. 1. Also, the pressure is normalized by
ρ∞ c∞ c∞ and viscosity by incoming flow viscosity µ∞ , force by the combined variable ρ∞ c∞ c∞ D. Ma∞ is the
Mach number U∞ /c∞ , and Re∞ = ρ∞ U∞ D/µ∞ is the Reynolds number. Furthermore, Pr = 0.75 is the Prandtl
number and γ = 1.4 is the specific heats ratio.
Particularly, from the internal energy equation in Eq. (10), the work done by the body force for moving objects
is only associated with flow kinetic energy 0.5ρ(u2 +v 2 ). This actually accounts for Galilean invariance in [25] for
moving-boundary problems.
∂ρe
+ ∇ · (ρeu) = S : P + ∇ · q + Q, (10)
∂t
where e = T/ (γ −1) is the internal energy, S:P = −p∇•u + µ(∇•u)2 + Φ, is viscous dissipation function [33], q
= (qx , q y ), Q is the modeled heat source term if specified wall thermodynamic boundary conditions are necessarily
satisfied.

2.2. Influence matrix method for moving-boundary problems

Unlike the explicit IB methods such as virtual spring force [34], penalized force [18], or feedback forcing
method [29], the semi-implicit method—IMM accounts for the influence of adjacent boundary forcing points such
4
L. Cheng, L. Du, X. Wang et al. Computer Methods in Applied Mechanics and Engineering 373 (2021) 113438

as points A and B in Fig. 1 via an influence matrix. The two overlapped squares (‘blue’ and ‘green’ in Fig. 1) denote
the joint influence regions of the boundary forcing points, A and B. In this region, the flow velocity is affected by the
calculated boundary forces on points, A and B, simultaneously. This consideration confirms the accurate satisfaction
of no-slip wall boundary conditions without any feedback iteration [16,26,27]. In general, to achieve the IMM, the
time-marching algorithm for the governing equations in Eq. (1) is divided into three steps:
(1) Prediction step (from current time level n to the intermediate level n + 1/2): perform the time marching for
the discretized Eq. (1) on the whole Cartesian background mesh without applying the velocity constraints – no-slip
wall boundary conditions (WBCs) – on the points near the FSI Γ. (f(x) = 0)
(2) Body force derivation: derive the boundary force equation by utilizing the IMM technique and the deviation
of the expected Lagrangian boundary velocity to the predicted one on the FSI Γ at time level n + 1/2, and then
calculate the body force distribution on the Cartesian background mesh by utilizing the interaction equations [11].
(3) Correction step (from time level n + 1/2 to n + 1): correct the flow velocity at time level n + 1/2 on the
Cartesian background mesh points near the FSI Γ using the calculated body force density, f, and thus obtaining the
accurate flow field at the next time level n + 1.
The detailed mathematical derivation of this algorithm is presented below to satisfy the no-slip WBCs.

2.2.1. No-slip wall boundary conditions


The derivation of the IMM for compressible N-S equations aims at achieving the accurate satisfaction of no-slip
WBCs, uw (x) = Uw (X), which actually provides an equality constraint for Eq. (1), where Uw (X) represents the
motion velocity of wall boundaries as we mentioned above. The discretized momentum equations are written in the
compact form of
(ρu)n+1 − (ρu)n
= RHSn + f, (11)
∆t
where RHSn represents the summation of advection, diffusion, and pressure gradient terms in the time level n. The
body force density f and the (ρu)n+1 are unknown quantities to be calculated.
• Prediction Step
In the prediction step, f is set to zero, which means wall effects are firstly not considered at intermediate time
level n + 1/2. Then, the intermediate conservative variables at time level n+1/2 are calculated by
(ρu)n+1/2 = (ρu)n + RHSn · ∆t, (12)
(ρu) n+1
= (ρu) + RHS · ∆t + f · ∆t,
n n
(13)
Obviously, the intermediate velocity field in Eq. (12) near the FSI Γ does not satisfy the equality constraint -
uw (x) = Uw (X). The accurate velocity field calculation given in Eq. (13) at the n + 1 time level should include the
exact body force density f, which is still an unknown quantity at the present stage, and will be calculated below.
• Body Force Derivation
The relationship among the body force density f, (ρu)n+1/2 and (ρu)n+1 is obtained by subtracting Eq. (12) from
Eq. (13) as
(ρu)n+1 − (ρu)n+1/2
f= , (14)
∆t
where ρ n+1 and ρ n+1/2 represent the flow density on the domain Ω in the time level n + 1 and the intermediate flow
density in the time level n + 1/2, respectively. In this work, no mass sinks/sources are added into the mass equation
in Eq. (1), so no flow density correction is conducted on the Cartesian grid points near FSI Γ. This makes that ρ n+1
= ρ n+1/2 holds in the whole computational domain Ω. On the other hand, it leads to the imperfect conservation
of mass near FSI Γ, but the previous numerical studies [16,18,25–27] indicate that it will not cause an observable
error for subsonic flow simulation.
By involving the Dirac delta function, the interaction equations [11] about Eulerian variables and Lagrangian
variables are introduced as

U(X) = u(x)δ(x − X)dx, (15)

5
L. Cheng, L. Du, X. Wang et al. Computer Methods in Applied Mechanics and Engineering 373 (2021) 113438

f(x) = F(X)δ(x − X)ds, (16)
Γ
where δ represents the Dirac delta function, Ω = Ω1 + Ω2 is the whole generalized computational domain depicted
in Fig. 1. U(X) = (Ux , U y ) represents the velocity of the Lagrangian boundary forcing point, X, on FSI Γ. ds is the
differential arc of the Lagrangian curvilinear boundary Γ. The interaction equation in Eq. (15) is also best described
as “interpolation” as illustrated in [32]. Both of Eqs. (15) and (16) hold at all time levels, n, n + 1/2 and n + 1.
Then, it is straightforward to obtain Eq. (17) by applying the Eq. (15) at n+1/2 and n+1 time levels, respectively.
U(X)n+1 − U(X)n+1/2 u(x)n+1 − u(x)n+1/2

= · δ (x − X) dx, (17)
∆t Ω ∆t
Uw (X) − U(X)n+1/2

f(x)
= · δ (x − X) dx, (18)
∆t Ω ρ n+1/2 (x)

where U(X)n+1 is also equal to the expected wall moving velocity, Uw (X), and U(X)n+1/2 is the predicted flow
velocity of the boundary forcing point, X, on FSI Γ in Fig. 1 at time level n + 1/2 in the “Prediction step”.
Subsequently, the body force density, f(x), in Eq. (14) is utilized to replace the Eulerian velocity deviation terms in
Eq. (17), and the following equation in Eq. (18) is obtained.
Hence, by substituting Eq. (16) into Eq. (18), the boundary force equation is derived in the generalized form of
Uw (X) − U(X)n+1/2 δ (x − X1 ) · δ (x − X)
∫ [∫ ]
= { dx · F (X1 ) } ds, (19)
∆t X1 ∈Γ Ω ρ n+1/2 (x)
However, Eq. (19) cannot be solved directly due to the singularity of δ function. δ(x-X) is only related to the relative
position of the Lagrangian and Eulerian grid points. When the Eulerian mesh point, x, coincides with the Lagrangian
boundary forcing point, X, it is equal to infinity; otherwise, it is zero. For the purpose of the normalization, the
approximate δ function in Eq. (20) in Ref. [11] is usually adopted.
1 rx ry
δd (r) = φ( )φ( ), (20)
d xdy d x dy
where φ(x) is the one-dimensional template function; dx and dy are the mesh sizes in the two directions; rx and r y
are the distances of the Lagrangian boundary forcing point from the Cartesian grid point in two directions. Besides,
δd (r) function is highly compact. In this work, dx = dy is used near FSI Γ. Furthermore, curvilinear and surface
integrals of Eqs. (15) and (16) are approximated by the Riemann summation [11] in the form of Eqs. (21) and (22).

U(Xk ) = u(x)δd (x − Xk )d xdy, (21)
x
M

f(x) = F (Xk ) δd (x − Xk )∆sk , (22)
k=1
where k = 1, ..., M, M is the discretized number of FSI Γ, ∆sk denotes the arc length of the kth segment for the
curvilinear FSI Γ. Hence, the linear algebraic form of boundary force equation in Eq. (19) is reduced to Eq. (23).
∆UT /∆t = A · F̃T , (23)
where ∆UT /∆t is the velocity deviation matrix of the expected and predicted moving velocity of the boundary
forcing points. Its expression and the unknown boundary force density matrix, F̃T , are denoted by Eqs. (24) and
(25).
Uw (X1 ) − U(X1 )n+1/2
⎡ ⎤
⎢ Uw (X2 ) − U(X2 )n+1/2
..

.
⎢ ⎥
⎢ ⎥
⎢ ⎥
⎢ n+1/2

T
⎢ U w (X k ) − U(X k ) ⎥,

∆U = ⎢ ⎢
.. (24)
.

⎢ ⎥
⎢ ⎥
⎢ ⎥
⎢Uw (X M−1 ) − U(X M−1 )n+1/2 ⎥
⎣ ⎦
n+1/2
Uw (X M ) − U(X M )
6
L. Cheng, L. Du, X. Wang et al. Computer Methods in Applied Mechanics and Engineering 373 (2021) 113438

]T
F̃T = F(X1 ) F(X2 ) · · · ,
[
F(Xk ) · · · F(X M−1 ) F(X M ) (25)
where both of the sizes of the F̃T and ∆UT are M×2 for 2-D problems, and the elements for the first and second
columns in two matrixes represent the force/velocity components in x and y directions, respectively. The size of the
coefficient matrix A is M×M, and its elements are expressed as
[ ]

A p,q = δd (x − Xq ) · δd (x − X p )/ρ n+1/2
· d xdy · ∆sq , (26)
x

where p,q = 1, . . . , M.
In this algorithm, A is called “influence matrix (IM)” and only related to the distribution of boundary forcing
points, X, on the fixed Eulerian background grid and the nearby predicted fluid density. Due to the compact support
quality of δd (r) function, IM A is quasi-triangular and highly-sparse, which makes it efficient to store the IM A
and develop fast solving algorithms.
• Correction Step:
un+1 = (un+1/2 + f · ∆t) / ρ n+1/2 , (27)
In the correction step, the local velocity field, u(x), near the FSI Γ is corrected by utilizing Eq. (27) when the
boundary force density F(X) is solved in Eq. (23). In practical computation, the assumption ρ n+1/2 (x) ≈ ρw (X p )
approximately holds near the Lagrangian boundary forcing point X p and ρ n+1/2 (x) is the fluid density near the
support range of point X p , when the compressibility effects of the boundary layer are not so important that an
assumption of incompressibility is reasonably made in the boundary layer [35,36]. Hence, the density term is able
to be transferred to the left side of Eq. (23) to get ∆(ρU)T /∆t, and then the element of the IM A is only associated
with the distribution function, δd (r), and the relative position of Eulerian and Lagrangian grid points.
In particular, the heat source term, Q, is neglected. It means that no specific temperature boundary conditions are
satisfied, and only body force work W is added into the total energy equation. It is actually reasonable for subsonic
flow in the absence of observable heat transfer or releasing processes [35,36].

2.2.2. Quality of IM and fast-solving algorithms


The choice of distribution function δd is a key ingredient in the indirect forcing IB method [37]. In the original IB
method, template function φ(x) is chosen to satisfy discrete moment conditions. Non-physical oscillations, however,
are generated for the body force calculation in simulating moving-boundary problems [31], which is undesirable for
high-order CAA methods. Yang et al. [31] proposed a smoothing technique on the basis of a Taylor series to make
it satisfy higher discrete moment conditions for the template function φ(x). In the present work, both the original
template function φ(x) in Ref. [11], support width of which is 4, and the smoothed version φs (x) in Ref. [31], support
width of which is 5, are considered for the IMM. The assumption is made that the discretized Lagrangian boundary
forcing points are equally distributed, so the arc length ∆sk of the curvilinear boundary Γ is identical with each other.
Combining Eqs. (22) and (26) in the prediction and correction steps, we can find[∑ that the boundary mesh size ∆sk is
x δd (x − Xq ) · δd (x − X p )d xdy .
]
able to be eliminated to make the IM A symmetric, that is, the element Arp,q =
L ≈ f loor (2n 0 . d x/∆s + 1), when d x/∆s ∼ 1.0 (28)
Setting the support width of the distribution function δd to n0 . dx, illustrated in Fig. 2, the band width L of the
non-zero elements in every row or column of IM A could be evaluated via the approximate relationship in Eq. (28)
for grid size ratio, dx/∆s, not far from 1.0, where “floor” is a function to round the element to the nearest integer
smaller than or equal to the original element. Eq. (28) is obtained for the situation when the FSI Γ is a straight line
as shown in Fig. 2. The band width L is also equal to the boundary forcing point number, 2(k2 − k1) + 1, between
points Xk1 and Xk2 on the FSI Γ. According to the quasi-triangular data structure of the IM A analyzed above, we
know that the non-zero elements only occupy L/M ×100% of the IM A. And the L is a small number in practical
calculations due to the compact quality of the distribution function. Therefore, the jagged diagonal storage (JDS)
scheme [30] is appropriate to store the IM A due to its high compression ratio.
The compact and symmetric quality of IM A makes it possible to develop a high-efficiency solving algorithm for
Eq. (23) rather than SVD recommended by Sun et al. [16]. One of the most efficient algorithms is the PCGM [38],
7
L. Cheng, L. Du, X. Wang et al. Computer Methods in Applied Mechanics and Engineering 373 (2021) 113438

Fig. 2. Principle to evaluate band width L in IM A (Xk1 and Xk2 : the k1th and k2th Lagrangian boundary forcing points on FSI Γ).

Fig. 3. The 2-norm condition number variation with the grid size ratio dx/∆s (CN nor m = log10 (cond(A))).

because the iterative searching approach is only related to matrix multiplication operations, and thus confirming
the best parallel efficiency. The convergence speed of the PCGM, however, is highly dependent on the condition
number of IM A. The relative position between Eulerian and Lagrangian grid points is mainly dependent on the
grid size ratio dx/∆s, so we firstly give the condition number distribution of IM A with varying the grid size ratio
dx/∆s for different typical geometries—flat plate (FP), circular cylinder (CC), and square cylinder (SC)—in Fig. 3.
The background Cartesian mesh size, dx, is fixed to 0.01 for all geometries, and the characteristic length is all 1.0.
The calculated results show that 2-norm condition number of the A increases rapidly, indicating that the IM A
tends to be ill-posed, when dx/∆s is close to and then beyond 1.0, as indicated in Fig. 3. The IM A calculated
with the φs (x) has a little larger condition number than that associated with the φ(x), which is mostly due to the
larger support width of the smoothed function, φs (x). Particularly, the condition number for the CC is remarkably
different from the other two geometries with sharp edges, so, in the preprocessing phase, the quality of the IM A
can be firstly checked to show boundary mesh quality before performing unsteady simulation especially for 3-D
problems.

2.2.3. Prior check model of the IMM convergence


Contrast to the numerical methods based on body-fitted grid, the convergence check of the IMM is a little
different. It is essential to conduct the grid convergence check as well as the body force convergence check. The
body force convergence indicates the satisfaction degree of the no-slip WBCs in IMM. Hence, joint numerical
experiments should be conducted for different combinations of the Eulerian background grid size, dx, and the
Lagrangian boundary grid size, ∆s. However, this joint check will result in huge time consumption if traditional
grid refinement work is done for practical simulation. At the same time, we noticed that the IM A in Eq. (23) is
mainly dependent on the positions of the Lagrangian boundary forcing points on the Eulerian coordinate(domain
8
L. Cheng, L. Du, X. Wang et al. Computer Methods in Applied Mechanics and Engineering 373 (2021) 113438

Ω). The physical meaning of the body force convergence in IMM is that the force acted on the fluid from the wall
boundary Γ needs to be convergent with refining the grid size combination (dx, ∆s) to make the WBCs satisfied
at each time step. In practical simulation, the grid size, dx, is mainly determined by the flow structure resolution
when solving N-S equations, and, therefore, the grid size ratio, dx/∆s, is the major influential factor for the body
force model convergence. We have known the exact relationship between the body force density, f(x), and boundary
force density, F(X), in Eq. (29) from Peskin’s work [32]. The summations in Eq. (29) are equivalent and both of
them represent the resultant force acted on the fluid from the wall boundary Γ.
( M ) ⎛ ⎞
∑ ∑
F(Xi ) ∆s = ⎝ f(xi, j )⎠ d xdy, (29)
i=1 xi,j ∈Ω

To efficiently validate the convergence quality of the IMM, a numerical model equation in Eq. (29) for the
boundary force density is extracted from Eq. (23) to perform the check in a prior phase.
Ax = b/d x, (30)
where b = (1,. . . ,1)T represents the unit velocity deviation. The term dx represents normalization factor
(∑ for different
)
M
IM A. With choosing different combinations of (dx, dx/∆s), the calculated resultant force, Fb = i=1 x i ∆s, is
expected to be convergent to a certain value, and, therefore, we think that the selected grid size combination can
confirm a convergent body force model in IMM.

2.3. Numerical schemes for spatial and temporal integration

The 7-point central dispersion-relation-preserving (DRP) scheme [39] is adopted for the spatial discretization in
the whole computational domain Ω even near the FSI Γ. This is due to the adoption of body force model. The
high-order optimized selective artificial filter in Eq. (31) is used to suppress short waves.
M2
f

Ui = Ui − σ f s j Ui+ j , (31)
j=−M1

where σ f is a constant between 0 and 1, s j represents the filtering coefficient, Ui is the conservative variable in
computational coordinates. The filtering coefficient s j with 9-point stencils in Eq. (31) developed by Bogey and
Bailly [40] is adopted in this work, which means: M1 = M 2 = 4. When encountering the wall boundaries, the
non-central filter with smaller stencils [40] is usually recommended in the frameworks of the body-fitted grid [21],
Cartesian [20], and GC methods [17], as illustrated in Fig. 5(a), while the IMM still utilizes the large-stencil
central filter in the whole domain Ω as shown in Fig. 4(b) due to the adoption of body force model. Many
illustrations [17,41] indicate that one-sided higher-order derivative/filtering schemes are unstable. By contrast, our
present method has great superiority in this respect.
Besides, the 2N-storage 5/6 stages low-dissipation and low-dispersion Runge–Kutta schemes (LDDRK5-6) [42]
are utilized for time integration. In our numerical experiments, it is found that the LDDRK5-6 is more stable in
the long-time behavior than the LDDRK4-6 for flow-induced sound problems. Also, it is known that non-reflecting
boundary conditions (NRBCs) are of vital importance in CAA due to the inevitable truncation of the infinite domain.
Hence, the perfectly matched layer (PML) absorbing boundary conditions for the nonlinear Euler equations [43]
are adopted in the far field.

3. Results and discussion


As we mentioned above, the grid convergence of the IMM can be numerically demonstrated using Eq. (30) in the
preliminary phase without directly performing practical unsteady simulations. In this section, we firstly demonstrate
the body force convergence of the IMM for two kinds of geometries (CC and SC) without doing unsteady simulation,
and give a new criterion of choosing mesh size in the combined IMM/CAA methods. Then, to quantitatively validate
our present IMM/CAA for simulating flow-induced sound from moving boundaries, two kinds of benchmark flow-
induced sound problems (single cylinder and flat plate cascade) are investigated in this section. The first kind of
simple cases are to study the present method’s practical convergence, especially its capability of dealing with objects
9
L. Cheng, L. Du, X. Wang et al. Computer Methods in Applied Mechanics and Engineering 373 (2021) 113438

Fig. 4. Filter stencils near wall boundaries for different high-order computational aeroacoustics method. # represents filter stencil points,
represents filtered points.

Fig. 5. Numerical experiments for the grid convergence of the IMM to solve the boundary force equation for different geometries and
different distribution functions ((a1),(a2): resultant forces Fb exerted on the circular cylinder (CC) and square cylinder (SC); (b1), (b2): the
proportion of nonzero elements in the IM A for the CC and SC).

with sharp edges; the other is to validate the capability of the IMM to tackle multi-bodies interaction noise, which
is known to be quite difficult to simulate for high-order CAA methods with traditional body-fitted grid. Here, we
10
L. Cheng, L. Du, X. Wang et al. Computer Methods in Applied Mechanics and Engineering 373 (2021) 113438

firstly give some quantity definitions in Eqs. (32), (33), (34) and (35).
C F L = (1 + Ma ∞ ) · ∆t / d x, (32)

CL = f y dx /(0.5ρ∞ U∞2
D), (33)
∫ Ω

Cd = f x dx / (0.5ρ∞ U∞
2
D), (34)

C p = ( p − p∞ ) / (0.5ρ∞ U∞
2
), (35)
where p in Eq. (35) is the time-averaged pressure. The Strouhal number is defined as St = fU ∞ /D, where f and
U∞ are the natural vortex shedding frequency and inflow velocity, respectively. fx and f y are body force density on
the Cartesian grid. For practical computation, the relationship in Eq. (29) is utilized, so the lift and drag coefficients
are obtained at every time step by integrating the boundary force density, F (X), along the FSI Γ. For moving-
(ρV) /∂t] dx, of the subdomain Ω2 should be subtracted from the

boundary problems [31], the inertial force, Ω2 [∂
Ω f (x) dx, firstly. For simplicity, the fluid in the subdomain Ω2 is treated as a ‘frozen’ object, which

total force,
means Ω2 [∂ (ρV) /∂t] dx = ma, where m and a denote the total fluid mass of Ω2 and the associated acceleration

velocity of the object, respectively.

3.1. Grid convergence of the IMM

For two geometries, CC and SC, the calculated resultant forces are plotted in Fig. 5(a) for different cases. They
yield an obvious convergent tendency for the resultant force Fb with decreasing the mesh size dx both for the CC
with a smooth curvilinear boundary Γ and SC with sharp edges. Meanwhile, we find that the grid size ratio dx/∆s
has a prominent influence on the convergence for different geometries, apart from affecting the condition numbers
of the IM A in Fig. 3. Only when the grid size ratio dx/∆s is in a suitable range could a convergent solution be
achieved.
Specifically, in Fig. 5(a), the convergence behaviors for the CC and SC are quite different. It has a wider
convergent range, approximately (0.5, 1.5), of the grid size ratio dx/∆s for the CC, compared to the convergent
range, (0.5, 1.045), for the SC, which is possibly due to the existence of the sharp edges for the SC. If the grid
size ratio dx/∆s is too small (<0.5), a nonphysical solution is obtained for the two configurations; otherwise (dx/∆s
>3.0), the IM A becomes ill-posed and the solution is unstable. In Fig. 5(b), it shows the proportion of nonzero
elements in IM A with different distribution functions. And it illustrates that the IM A is highly sparse if the grid
size is less than 0.025 for all cases, and, in this case, the non-zero elements only make up less than 5% of the
IM A, indicating quite low demands for computer memory. In addition, the final convergent values for the force
Fb are different for two distribution functions, and the influence on the no-slip WBCs is hard to evaluate only by
solving Eq. (30), so the practical unsteady simulation in the following is more suitable to be utilized to investigate
the influence. Combining the above investigations of the condition number for the IM A in Fig. 3 with the present
results in Fig. 5, it is appropriate to choose the grid size ratio dx/∆s near or a little smaller than 1.0 which is able to
efficiently achieve a convergent solution as well as avoid numerical stiffness. The previous numerical experiments
[16,26] also provided similar illustrations.

3.2. Flow-induced sound from a single object

Laminar flow around a single object (CC and SC) is a simple nonlinear case, so it is widely used to validate
numerical methods. Sufficient DNS results [7,9,10,18,25] both for the near-field flow structures and far-field acoustic
information are available. In this section, the stationary cylinder streaming (Ma∞ = 0.2, Re∞ = 150) simulated
by Inoue and Hatakeyama [10] and the oscillating cylinder streaming (Ma∞ = 0.2, Re∞ = 150) fulfilled via the
penalization method [25] are studied. The computational domain is [−100,100] × [−100,100]. The grid size ratio
dx/∆s near objects is set to 0.95, which has been demonstrated to achieve a convergent body force model for the
CC and SC in Fig. 5. A stretched grid is utilized in the far field to reduce grid points. Here, a stretched grid ratio,
8h, is used, which gives approximately 27 points per wave length, also called point-per-wave (PPW) number, in
the far field.
11
L. Cheng, L. Du, X. Wang et al. Computer Methods in Applied Mechanics and Engineering 373 (2021) 113438

Fig. 6. (Color online) Pressure coefficient and far-field sound pressure level pr ms on the circle r p = 75 (fixed circular cylinder, Re∞ = 150,
dx = 0.017).

For the fixed CC case, the time histories of the unsteady pressure, p, at points (0, 75) and the instantaneous pres-
sure field are depicted in Fig. 6(a). Convergent solution of the far-field sound pressure is observed. Quantitatively,
the pressure coefficients, C p and root-mean-square (RMS) of sound pressure at numerical probes distributed on the
circle with the radius r p = 75, for different CFL numbers and different distribution functions are plotted in Fig. 6(b1)
and (b2). Good agreement is obtained with the previous body-fitted grid DNS results [10]. Besides, the adopted
CFL number in this paper is obviously larger than the 0.3∼0.5 in the GC method [17] and the 0.3 in the body-fitted
grid [10] for the stationary case, which is one of the superiority for the IMM over other interpolation-based IB
methods [17]. Moreover, from Fig. 6(b1), it also suggests that both the original function, φ(x), and the smoothed
one, φs (x), achieve a correct prediction for the unsteady flow problems, so the small discrepancy in Fig. 5(b1) does
not make any inherent difference in the unsteady force prediction on the object and its radiated sound field. Four
groups of mesh size, dx = 0.040, 0.020, 0.017, and 0.010, are chosen to perform the grid independence test in
Table 1. Agreement is obtained with the previous DNS results [10,25] both for the unsteady force and radiated SPL
with refining the grid size. Besides, the local convergence is also investigated based on the quantities, C L(r ms) and
pr ms . The linear least square fitting is utilized to process the logarithm deviations of the calculated results with the
DNS results [10] on body-fitted grid with refining the mesh, and thus obtaining the convergence orders, 1.53 and
0.89, for C L(r ms) and pr ms , respectively.
12
L. Cheng, L. Du, X. Wang et al. Computer Methods in Applied Mechanics and Engineering 373 (2021) 113438

Table 1
Solution for the fixed Circular Cylinder (CC) case (Re∞ = 150, pr ms is taken at r p = 75, θ = 90◦ ; BFG: body-fitted grid; In.: incompressible;
VP: volume penalization).
Method dx C L(r ms) Cd St pr ms (×105 ) Local convergence order
IMM 0.040 0.3544 1.348 0.177 7.601
1.53(C L(r ms) )
IMM 0.020 0.3681 1.349 0.180 8.047
/ 0.89(pr ms )
IMM 0.017 0.3661 1.348 0.180 7.990
IMM 0.010 0.3661 1.344 0.182 7.806
BFG [10] – 0.3677 1.320 0.183 7.900
VP [25] – 0.3499 1.300 0.183 7.710 –
BFG, In. [44] – 0.3748 1.334 0.182 –
‘Universal’ [45] – 0.3555 – 0.183 –

Particularly, we present the time-averaged body force density distribution in Fig. 7 to investigate the influence
of grid size dx and grid size ratio dx/∆s. The results indicate that the value of the body force density on the FSI
is inversely proportional to the grid size dx, while it is unaffected by the grid size ratio dx/∆s when the ratio is
chosen from a convergent range in Fig. 5. The inversely proportional trend is mainly because the introduction of
1-D boundary force F(X) (2-D corresponding to 3-D problems) in Eq. (16) makes the density f(x) have the quality
of 1-D Dirac delta function [32], which is inversely proportional to the grid size dx. This quality for the IMM is
prominently different from that of the penalization method [18], which utilizes a 2-D boundary force model for
2-D problems (3-D model for 3-D problems). This difference makes the IMM only need a lower computational
effort because only the nearby points of the FSI Γ are tackled, not the whole extended fluid subdomain Ω2 , but,
at the same time, solving the boundary force equation in Eq. (23) is really time-consuming and dependent on the
geometry and the grid size ratio dx/∆s.
Next, the transversely vibrating case is further simulated to validate the developed IMM/CAA for moving-
boundary problems. The grid and computational domain are all kept same as the above stationary case, and the
minimum grid size dx, is 0.017 (Lagrangian boundary forcing point number, M = 180). Correspondingly, CFL =
0.8 is used. The motion equation is given by Eq. (36).
Y (t) = Ym sin (2π f e t + ϕ0 ) , (36)
where Ym = 0.2 is the vibration amplitude, nondimensionalized by the cylinder diameter D. fe is the vibration
frequency, nondimensionalized by c∞ /D, and fe = 0.15Ma∞ is adopted for the oscillating CC case, while fe =
0.14Ma∞ is used for the oscillating SC case, and ϕ0 is the initial phase angle, equal to 0◦ in this part. Particularly,
to keep same as in Ref. [25], the Reynold number, Re∞ = 150, is adopted for the oscillating CC case, while
Re∞ = 100 is used for the oscillating SC case. Both the original template functions φ(x) and φs (x) are considered.
By contrast, a frequently adopted distribution function φ0 (x) = 1 − |x| (support [−1,1]) in [26] is chosen for
comparison. The three template functions are plotted in Fig. 8. It is worth noting that no numerical oscillations are
observed for the stationary CC case above when φ0 (x) is adopted.
The calculated lift coefficients C L with different template functions, are plotted in Figs. 9(a) and 10(a) for the CC
case and SC case, respectively. Intensive numerical oscillation occurs for Figs. 9(a) and 10(a) with 2-support-width
template function φ0 (x), and quite slight oscillation is observed for the force with 4-support-width template function
φ(x), while, by contrast, no numerical oscillation is observed with the smoothed template function φs (x). It exhibits
the significance of the approximate Dirac delta function with high-order vanishing discretized moments [31] in
high-order CAA methods from the oscillating CC and SC cases. But it is worth noting that the inherent discrepancy
is actually negligible for the unsteady flow simulation with adopting the two template functions, φ(x) and φs (x).
Furthermore, agreement is obtained in Figs. 9(b) and 10(b) for the far-field sound directivities compared to results
simulated by the volume penalization method [25]. Specifically, the instantaneous Mach number and pressure
contours are illustrated in Fig. 11 for the CC and SC. It suggests that the interior flow of the objects does not
affect the really concerned unsteady flow in fluid subdomain Ω1 . It indirectly proves the original concept of body
force model that the FSI Γ can be extended as a generalized discontinuity [23]. In addition, the IMM/CAA solver has
a good resolution for the strong-gradient flow near the sharp edges of the SC from Fig. 11(a2) and (b2), where the
Mach number and pressure have an abrupt jump across the interface, although it is widely thought that the adoption
13
L. Cheng, L. Du, X. Wang et al. Computer Methods in Applied Mechanics and Engineering 373 (2021) 113438

Fig. 7. The influence of grid size dx and grid size ratio dx/∆s on the distribution of time-averaged body force density on the fluid/solid
interface (r p = 0.5) ((a) different grid sizes dx (CFL = 0.8), 1- original distribution, 2- normalized distribution using 100dx; (b) different grid
size ratios dx/∆s (dx = 0.01, CFL = 0.8); (c) time-averaged body force in two directions. Fixed circular cylinder (CC) case, Re∞ = 150).

Fig. 8. Template function.

14
L. Cheng, L. Du, X. Wang et al. Computer Methods in Applied Mechanics and Engineering 373 (2021) 113438

Fig. 9. (Color online) Lift coefficient variations for the transversely oscillating cylinder in a flow at Re∞ = 150 and Ma∞ = 0.2 and the
corresponding sound pressure directivity at r p = 80 (Oscillating circular cylinder (CC) case, fe = 0.15Ma∞ ).

Fig. 10. (Color online) Lift coefficient variations for the transversely oscillating cylinder in a flow at Re∞ = 100 and Ma∞ = 0.2 and the
corresponding sound pressure directivity at r p = 80 (Oscillating square cylinder (SC) case, fe = 0.14Ma∞ ).

of the approximate Dirac delta function makes it degrade to only first-order accuracy for the spatial schemes near
FSI Γ [24,46].
Moreover, the grid refinement work is done, and the grid size ratio dx/∆s, 0.95, is chosen from a convergent
range in Fig. 5(a). The smoothed template function φs (x) is always utilized for all cases in Table 2. We still compare
the near-field lift and drag coefficients, C L(r ms) and Cd , as well as the far-field sound pressure level pr ms (×105 )
(r p = 80, θ = 90◦ ) in three different grid sizes. It indicates that the shedding vortex frequency, St · Ma∞ , is equal
to the vibration frequency fe for these cases. We could see that the force coefficients are close to the results [25]
calculated by other methods with refining the grid near FSI Γ. The adopted size grid, dx = 0.010, above, causes
(0.035–0.72) dB and (0.0048–0.072) dB discrepancy of the lift/drag coefficients and pr ms , respectively, from the
results calculated by volume penalization method [25] and the body-fitted grid method [25] for the oscillating
cylinder cases in Table 2. The present discrepancy is mainly caused by the insufficient grid near the FSI Γ, while
the body-fitted grid is superior to resolve the flow near wall boundaries [10]. To better resolve the boundary layer,
especially high Reynolds number flow, adaptive mesh refinement (AMR) is essential for practical application.
Correspondingly, the local convergence orders are also obtained for the vibration circular cylinder (CC) and
square cylinder (SC) cases. Similar orders associated with unsteady lift are obtained for CC (0.74) and SC (0.98)
cases, while almost 1.5 convergence order associated with the far-field sound pressure is achieved for SC cases,
compared to the small order- 0.47 in the CC cases. The large discrepancy indicates that the convergence speed of
15
L. Cheng, L. Du, X. Wang et al. Computer Methods in Applied Mechanics and Engineering 373 (2021) 113438

Fig. 11. Instantaneous Mach number and pressure contours for the oscillating circular cylinder (CC) (Re∞ = 150, fe = 0.15Ma ∞ ) and
square cylinder (SC) (Re∞ = 100, fe = 0.14Ma ∞ ) ((a): Mach number contours; (b): pressure contours).

Table 2
Comparison for the flapping vibrating cylinder (pr ms is taken at r p = 80, θ = 90◦ ; BFG: body-fitted grid; VP: volume penalization).
Case Method dx CL(rms) Cd St Pr ms (×105 ) Local convergence order
IMM 0.04 0.0777 1.338 0.15 3.730
0.74(C L(r ms) )
Vibrating CC IMM 0.020 0.0721 1.313 0.15 3.661
/0.47(Pr ms )
Re∞ = 150 IMM 0.017 0.0716 1.310 0.15 3.624
f e = 0.15Ma∞ IMM 0.010 0.0700 1.305 0.15 3.597
BFG [25] – 0.0654 1.289 0.15 3.423 –
VP [25] – 0.0654 1.281 0.15 3.599
IMM 0.04 0.3577 1.883 0.14 11.165
0.98(C L(r ms) )
Vibrating SC IMM 0.020 0.2922 1.704 0.14 9.272
/1.42(Pr ms )
Re∞ = 100 IMM 0.017 0.2982 1.677 0.14 9.323
f e = 0.14Ma∞ IMM 0.010 0.2910 1.627 0.14 9.200
BFG [25] – 0.2678 1.569 0.14 8.869 –
VP [25] – 0.2898 1.498 0.14 9.276

the far-field sound pressure seems quite case-by-case, but it is noteworthy that unlike the near-field uniform grid,
the non-uniform grid in our simulation is also a significant factor for some uncertainties for the convergence testing.
Hence, the order testing for far-field local variables is actually not strict enough.

3.3. Flow-induced sound from multi-bodies interaction

In this section, sound generated from an oscillating cascade of flat plates with zero thickness as illustrated in
Fig. 12 is simulated. This multi-body interaction model has been widely adopted in analyzing aeroelastic instability
such as flutter and high-cycle fatigue in turbomachinery [47–50]. When it comes to the radiated sound field, little
16
L. Cheng, L. Du, X. Wang et al. Computer Methods in Applied Mechanics and Engineering 373 (2021) 113438

Fig. 12. Sketch of the flat plate cascade.

Table 3
Parameter setup of the computational cascade model.
Parameter Value
Cascade model Flat plates with zero thickness
Motion way Prescribed flapping motion
Chord length, C/C 1
Gap to chord, g/C 0.77
Mean incidence angle 0.0
Inter-blade-phase-angle, IBPA 90.0◦
Vibration amplitude, Ym /C 0.4%, 10%
Vibration frequency, fe = f ′ C/c0 0.22
Inlet mach number, Ma∞ 0.65
Reynolds number, Re∞ 500

results have been published due to the previously mentioned grid difficulty in high-order CAA methods [28]. Hence,
the emphasis is put on the multi-bodies interaction sound here. Specifically, the parameter setup is illustrated
in Table 3. It is worth noting that the inter-blade-phase angle (IBPA) of the flapping vibration is 90◦ , so ϕ0 =
90◦ N0 in Eq. (32), where N0 = 1, 2, 3, 4 denotes flat plate blade number in Fig. 12. Both a small and a large
amplitudes, Ym /C = 0.4% and 10%, are studied. They represent linear and nonlinear flow-induced sound situations,
respectively. It is available for the inviscid exact solution for the linear case [51]. The computational domain is
[−6,6]×[−1.538,1.538]. Periodic boundary conditions are utilized in the upper and lower boundaries y = ±1.538.
Besides, CFL is 0.8, and the grid size ratio, dx/∆s = 0.95.
For a smaller vibration amplitude, Ym /C = 0.4%, except for the deviation near leading edge (LE), our calculated
results have a good agreement with the analytical solution – FINEL [47,51] – as shown in Fig. 13. The unsteady
loading is defined as the unsteady pressure coefficients difference, ∆C p = C p,upper − C p,lower , between the upper
and lower surface of the flat plate. The fast Fourier transformation (FFT) is utilized to obtain the amplitude of the
unsteady pressure difference, that is ‘Am(∆C p )’ in Fig. 13, where C p represents the unsteady pressure coefficient
in Eq. (35). To make the results comparable with different vibration amplitudes, Ym , the Am(∆C p )/Ym at the first
harmonic of vibration frequency, f = f e , is depicted in Fig. 13. For the deviation near the LE in Fig. 13, it is
worth mentioning that the FINEL solution [51] is based on the convective wave equation that is linear and inviscid,
so theoretically, the unsteady lift loading near LE is expected to approach to infinity due to the LE singularity.
However, for the viscous flow solution in our work, it actually represents the real viscous situation only allowing a
finite value of unsteady loading near LE. Hence, the deviation near LE in Fig. 13 from the FINEL solution [47,51] is
reasonable. For the large-vibration amplitude case (Ym /C = 10%), similarly, the Am(∆C p )/Ym at the first harmonic
of vibration frequency, f = f e , is depicted in Fig. 13. The curve is not coincident with the smaller vibration amplitude
case. It indicates that the unsteady pressure loading amplitude, Am(∆C p ), is not linearly dependent on the vibration
amplitude, Ym .
17
L. Cheng, L. Du, X. Wang et al. Computer Methods in Applied Mechanics and Engineering 373 (2021) 113438

Fig. 13. FFT amplitude distributions of the unsteady pressure coefficient difference, ∆C p , of the upper and lower surfaces for the Plate 1.

Fig. 14. Lift coefficients, C L , and their corresponding FFT spectra (the FFT magnitude for the smaller vibration amplitude, Ym /C = 0.4%,
has been multiplied by a factor 25).

The corresponding lift coefficients are given in Fig. 14. It is observed that for the large vibration amplitude case,
observable distortion appears both for the lift and drag curves, indicating the appearance of nonlinearity. It can
be further confirmed from the second harmonic component in the FFT spectra. We also record the instantaneous
unsteady pressure both for the upstream point (−3, 0) and the downstream point (3, 0) in Fig. 15(a), and FFT spectra
are also given in Fig. 15(b). Compared to the results of small vibration amplitude, Ym /C = 0.4%, distorted curves
of the unsteady pressure appear for the larger amplitude, Ym /C = 10%. The higher harmonics on the FFT spectra
in Fig. 15(b) suggest that nonlinear acoustic waves are generated and propagated to the upstream and downstream.
Furthermore, the instantaneous vorticity and pressure contours with different vibration amplitudes are depicted in
Figs. 16 and 17. Compared to the symmetric wake in Fig. 16(a) for the small vibration amplitude, large-magnitude
periodic vortex shedding is observed in Fig. 16(b) for the large vibration amplitude case. When it comes to the
unsteady pressure waves as shown in Fig. 15, strong pressure wave is generated near the LEs and TEs in Fig. 17(b)
for the large vibration amplitude especially for the upstream. This yields the strong nonlinearity shown in Fig. 15.
By contrast, due to the small vibration amplitude, the sound is generated by the oscillating boundary layer in a linear
behavior, so the generated sound wave is very weak in Fig. 17(a). Particularly, the unsteady pressure distribution near
FSI is smooth and no spurious short waves or numerical instability are observed near the interface, which indicates
that our developed IMM is of great capability to describe the both linear and nonlinear pressure wave generation
and propagation processes. Generally, the accurate capture both for the blade unsteady loading and radiated sound
18
L. Cheng, L. Du, X. Wang et al. Computer Methods in Applied Mechanics and Engineering 373 (2021) 113438

Fig. 15. Unsteady pressure histories for upstream and downstream points, (−3, 0) and (3, 0), and their FFT spectra for different vibration
amplitudes (p0 is the mean pressure).

Fig. 16. Instantaneous vorticity fields at t = 190 for the prescribed flapping flat plate cascade with different vibration amplitudes.

structures is quite useful in analyzing the pressure signals associated with the aeroelastic instability of blades [3]
or finding effective noise control strategies in turbomachinery [28,49,50]. By contrast, it is challenging to finish
this simulation by utilizing moving-mesh technique or dynamic overset grid technique in high-order CAA methods
due to the grid contradiction. Hence, the present developed body force model—IMM, to some extent, exhibits the
superiority of simulating this kind of flow-induced sound problems using high-order CAA methods.

4. Conclusions
In our study, a semi-implicit body force model, called IMM, is developed for high-order CAA methods to directly
predict flow-induced sound generated by moving boundaries. The no-slip wall boundary condition is implemented
in a prediction–correction manner in this model, so it is accurately satisfied in each time step without any feedback
operation and without exciting grid-to-grid oscillation near time-varying fluid–solid interface due to the adoption
of smoothed Dirac delta function. Particularly, the introduction of the body force model makes it feasible and
efficient to discretize the spatial derivatives using symmetric schemes and to use symmetric-stencil filters even near
the interface, which better confirms the numerical stability and low dissipation and dispersion characteristics of
high-order CAA schemes. The grid convergence of the IMM is proved by introducing a model equation for the first
time. It is found that it is better to keep the grid size ratio dx/∆s a little smaller than 1.0, thus confirming both the
boundary-force-solving efficiency and convergent solution.
19
L. Cheng, L. Du, X. Wang et al. Computer Methods in Applied Mechanics and Engineering 373 (2021) 113438

Fig. 17. Instantaneous pressure fields at t = 190 for the prescribed flapping flat plate cascade with different vibration amplitudes, Ym /C =
0.4% and 10%.

Then, good agreement is obtained with the previous DNS results based on the body-fitted grid and penalization
method for the simple oscillating CC and SC cases. It indicates the capability and accuracy of the present IMM/CAA,
especially for sound generated by moving objects with sharp edges, and the local convergence order ranges from
0.47 to 1.47 dependent on local variables and cases. Furthermore, a multi-bodies interaction case—oscillating flat
plate cascade is simulated in a laminar flow to show our present IMM/CAA in predicting flow-induced linear and
nonlinear sound waves.
Last but not least, the present algorithm has a great potential to be extended to 3D. This is why we paid much
attention to the analysis of the IMM features, and then developed the corresponding low-memory storage scheme
and fast solving algorithm. But honestly speaking, there is much work to do especially for our parallel solver
development, so the 3D extension study is still ongoing.

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.

Acknowledgments
The present study was supported by the Natural Science Foundation of China (NSFC) (Grant Nos. 51790514
and 51706006) and the China Scholarship Council (CSC) (Grant No. 201806020045). The author – Long Cheng –
would also appreciate Dr. Xiaojue Zhu from the School of Engineering and Applied Sciences at Harvard University
for his valuable advice about the convergence discussion of the developed semi-implicit IB body force model using
the model equation.

References
[1] M.J. Lighthill, On sound generated aerodynamically. I. general theory, Proc. R. Soc. Lond. Ser. A Math. Phys. Eng. Sci. 211 (1107)
(1952) 564–587.
[2] M. Wang, J.B. Freund, S.K. Lele, Computational prediction of flow-generated sound, Annu. Rev. Fluid Mech. 38 (1) (2005) 483–512.
[3] D.S. Whitehead, Force and Moment Coefficients for Vibrating Aerofoils in Cascade, H.M. Stationery Office, London, UK, 1962, R &
M, 3754.
20
L. Cheng, L. Du, X. Wang et al. Computer Methods in Applied Mechanics and Engineering 373 (2021) 113438

[4] K.S. Brentner, F. Farassat, Modeling aerodynamically generated sound of helicopter rotors, Prog. Aerosp. Sci. 39 (2–3) (2003) 83–120.
[5] J.E.F. Williams, D.L. Hawkings, Sound generation by turbulence and surfaces in arbitrary motion, Philos. Trans. R. Soc. A Math.
Phys. Eng. Sci. 264 (1151) (1969) 321–342.
[6] C.K.W. Tam, H. Ju, Aerofoil tones at moderate Reynolds number, J. Fluid Mech. 690 (2012) 536–570.
[7] O. Inoue, Effect of initial condition on the sound generation by flow past a rotary-oscillating circular cylinder, Phys. Fluids 18 (11)
(2006) 77.
[8] A. Sharma, S.K. Lele, Sound generation due to unsteady motion of a cylinder, Phys. Fluids 23 (4) (2011) 125.
[9] O. Inoue, M. Mori, N. Hatakeyama, Control of aeolian tones radiated from a circular cylinder in a uniform flow, Phys. Fluids 15 (6)
(2003) 1424–1441.
[10] O. Inoue, N. Hatakeyama, Sound generation by a two-dimensional circular cylinder in a uniform flow, J. Fluid Mech. 471 (471) (2002)
285–314.
[11] C.S. Peskin, Flow patterns around heart valves: A numerical method, J. Comput. Phys. 10 (2) (1972) 252–271.
[12] W.X. Huang, H.J. Sung, An immersed boundary method for fluid–flexible structure interaction, Comput. Methods Appl. Math. 198
(33–36) (2009) 2650–2661.
[13] Y. Mori, C.S. Peskin, Implicit second-order immersed boundary methods with boundary mass, Comput. Methods Appl. Math. 197
(25–28) (2008) 2049–2067.
[14] J.H. Pan, M.J. Ni, N.M. Zhang, A consistent and conservative immersed boundary method for MHD flows and moving boundary
problems, J. Comput. Phys. 373 (2018) 425–445.
[15] Y. Ma, J. Cui, N.R. Vadlamani, P.G. Tucker, Hierarchical geometry modelling using the immersed boundary method, Comput. Methods
Appl. Math. 355 (2019) 323–348.
[16] X. Sun, Y. Jiang, A. Liang, X. Jing, An immersed boundary computational model for acoustic scattering problems with complex
geometries, J. Acoust. Soc. Am. 132 (5) (2012) 3190–3199.
[17] J.H. Seo, R. Mittal, A high-order immersed boundary method for acoustic wave scattering and low-Mach number flow-induced sound
in complex geometries, J. Comput. Phys. 230 (4) (2011) 1000–1019.
[18] Q. Liu, O.V. Vasilyev, A Brinkman penalization method for compressible flows in complex geometries, J. Comput. Phys. 227 (2)
(2007) 946–966.
[19] L. Wang, F.B. Tian, J. Lai, An immersed boundary method for fluid–structure–acoustics interactions involving large deformations and
complex geometries, 2019, arXiv preprint arXiv:1905.00182.
[20] K.A. Kurbatskii, C.K.W. Tam, Cartesian boundary treatment of curved walls for high-order computational aeroacoustics schemes, AIAA
J. 35 (1) (1997) 133–140.
[21] C.K.W. Tam, J.C. Webb, Z. Dong, A study of the short wave components in computational acoustics, J. Comput. Acoust. 01 (01)
(1993) 1–30.
[22] L.N. Trefethen, Group velocity in finite difference schemes, SIAM Rev. 24 (2) (1982) 113–136.
[23] L. Sirovich, Initial and boundary value problems in dissipative gas dynamics, Phys. Fluids (1958-1988) 10 (1) (1967) 24–34.
[24] A. Liang, X. Jing, X. Sun, Constructing spectral schemes of the immersed interface method via a global description of discontinuous
functions, J. Comput. Phys. 227 (18) (2008) 8341–8366.
[25] R. Komatsu, W. Iwakami, Y. Hattori, Direct numerical simulation of aeroacoustic sound by volume penalization method, Comput.
Fluids 130 (2016) 24–36.
[26] S.W. Su, M.C. Lai, C.A. Lin, An immersed boundary technique for simulating complex flows with rigid boundary, Comput. Fluids 36
(2) (2007) 313–324.
[27] L. Cheng, L. Du, X. Wang, L. Wu, X. Jing, X. Sun, Influence of non-uniform mean flow on acoustic scattering from complex
geometries, Comput. Fluids 163 (2017) 20–31.
[28] L. Cheng, L. Du, X. Sun, Computation of flow-induced sound from gust interacting with a vibrating cascade using body force model,
in: 25th AIAA/CEAS Aeroacoustics Conference, Delft, Netherlands, 2019, 2510.
[29] D. Goldstein, R. Handler, L. Sirovich, Modeling a no-slip flow boundary with an external force field, J. Comput. Phys. 105 (2) (1993)
354–366.
[30] Y. Saad, Iterative Method for Sparse Linear Systems, Vol. 82, SIAM, 2003.
[31] X. Yang, X. Zhang, Z. Li, G.W. He, A smoothing technique for discrete delta functions with application to immersed boundary method
in moving boundary simulations, J. Comput. Phys. 228 (20) (2009) 7821–7836.
[32] C.S. Peskin, The immersed boundary method, Acta Numer. 11 (2003) 479–517.
[33] H. Schlichting, Boundary-Layer Theory, McGraw-Hill, Springer, 1979.
[34] L. Zhu, C.S. Peskin, Simulation of a flapping flexible filament in a flowing soap film by the immersed boundary method, J. Comput.
Phys. 179 (2) (2002) 452–468.
[35] D.C. Wilcox, Turbulence Modeling for CFD, Vol. 2, DCW Industries, Canada, CA, 2006.
[36] R.M.C. So, T.B. Gatski, T.P. Sommer, Morkovin hypothesis and the modeling of wall-bounded compressible turbulent flows, AIAA J.
36 (9) (2015) 1583–1592.
[37] R. Mittal, G. Iaccarino, Immersed boundary method, Annu. Rev. Fluid Mech. 14 (37) (2004) 239–261.
[38] M.R. Hestenes, E. Stiefel, Methods of conjugate gradients for solving linear systems, J. Res. Natl. Bur. Stand. 49 (1952) 409–436.
[39] C.K.W. Tam, J.C. Webb, Dispersion-relation-preserving finite difference schemes for computational acoustics, J. Comput. Phys. 107
(2) (1993) 262–281.
[40] C. Bogey, C. Bailly, A family of low dispersive and low dissipative explicit schemes for flow and noise computations, J. Comput.
Phys. 194 (1) (2004) 194–214.
21
L. Cheng, L. Du, X. Wang et al. Computer Methods in Applied Mechanics and Engineering 373 (2021) 113438

[41] M.R. Visbal, D.V. Gaitonde, On the use of higher-order finite-difference schemes on curvilinear and deforming meshes, J. Comput.
Phys. 181 (1) (2002) 155–185.
[42] D. Stanescu, W.G. Habashi, 2N-storage low dissipation and dispersion Runge–Kutta schemes for computational acoustics, J. Comput.
Phys. 143 (2) (1998) 674–681.
[43] F.Q. Hu, X.D. Li, D.K. Lin, Absorbing boundary conditions for nonlinear Euler and Navier–Stokes equations based on the perfectly
matched layer technique, J. Comput. Phys. 227 (9) (2008) 4398–4424.
[44] C. Liu, X. Zheng, C.H. Sung, Preconditioned multigrid methods for unsteady incompressible flows, J. Comput. Phys. 139 (1) (1998)
35–57.
[45] C. Norberg, Fluctuating lift on a circular cylinder: review and new measurements, J. Fluid Struct. 17 (1) (2003) 57–96.
[46] R.J. Leveque, Z. Li, The immersed interface method for elliptic equations with discontinuous coefficients and singular sources, SIAM
J. Numer. Anal. 31 (4) (1994) 1019–1044.
[47] L. He, Integration of 2-D fluid/structure coupled system for calculations of turbomachinery aerodynamic/aeroelastic instabilities, Int. J.
Comput. Fluid D 3 (3–4) (1994) 217–231.
[48] K.C. Hall, F.C. Edward, Calculation of unsteady flows in turbomachinery using the linearizedeuler equations, AIAA J. 27 (6) (1989)
777–787.
[49] X. Sun, S. Kaji, Control of blade flutter using casing with acoustic treatment, J. Fluid Struct. 16 (5) (2002) 627–648.
[50] Y. Sun, X. Wang, L. Du, X. Sun, Effect of acoustic treatment on fan flutter stability, J. Fluid Struct. 93 (2020) 102877.
[51] D.S. Whitehead, The calculation of steady and unsteady transonic flow in cascades, in: Report CUED, Cambridge University Engineering
Department, UK, 1982, A-Turbo/TR118.

22

You might also like