You are on page 1of 22

Journal of Sound and Vibration 501 (2021) 116044

Contents lists available at ScienceDirect

Journal of Sound and Vibration


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

Vortex-induced vibration mechanism of the NACA 0012 airfoil


based on a method of separating disturbances ✩
Le Han a,b,1, Dasheng Wei a,b,2, Yanrong Wang a,b,3,∗, Xiaojie Zhang a,b,4
a
Beihang University, Beijing 100191, China
b
Collaborative Innovation Center for Advanced Aero-Engine, Beijing 100191, China

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

Article history: Vortex-induced vibration (VIV) is the result of the interaction between the airfoil vibration
Received 31 August 2020 and vortex shedding, and it frequently achieves the peak vibration amplitude in the lock-
Revised 16 February 2021
in region. Therefore, distinguishing the effect of the two disturbances in the lock-in region
Accepted 21 February 2021
is important for understanding VIV. In this study, the NACA 0012 airfoil was investigated
Available online 26 February 2021
at high angles of attack; thus, the vortex could form and shed regularly downstream. The
Keywords: VIV was analyzed numerically and verified using the test data in references. The process of
Vortex-induced vibration (VIV) the vortex shedding was observed to be similar under different incoming flows, vibration
Lock-in amplitudes, and frequency ratios. In the V-shape lock-in region, the phase between the air-
Vortex shedding foil vibration and vortex shedding changes with the frequency ratio. When the frequency
Airfoil vibration ratio is 1, the phase and strength of the vortex shedding remain almost constant with the
Method of separating disturbances vibration amplitude. Based on the features of the vortex shedding, a method of separat-
Energy balance ing the disturbances caused by the airfoil vibration and the vortex shedding is proposed
to investigate the mechanism of VIV from the perspective of energy balance. As the phase
between the airfoil vibration and vortex shedding changes in the lock-in region, the vortex
shedding provides negative aerodynamic damping (AD) in some phases and positive AD
in other phases. The stability of the airfoil depends on the two disturbances mentioned
above. VIV occurs only when the disturbance of the vortex shedding performs positive
work (negative AD) and the disturbance of the airfoil vibration performs negative work
(positive AD). When the vibration amplitude is small, VIV is dominated by the disturbance
of the vortex shedding. When the vibration amplitude is high, the VIV is dominated by
the disturbance of the airfoil vibration. As the vibration amplitude increases, the two form
a balance, and the airfoil exhibits limit cycle oscillations. Based on the assumption of lin-
ear superposition, the theoretical limit vibration amplitudes of the airfoil under the two
conditions in this study were 3° and 13.4°, respectively.
© 2021 Elsevier Ltd. All rights reserved.


Handling Editor: Dr. Lixi Huang.

Corresponding author.
E-mail addresses: hanle@buaa.edu.cn (L. Han), yrwang@buaa.edu.cn (Y. Wang), zxjbuaa@buaa.edu.cn (X. Zhang).
1
Ph.D Candidate, School of Energy and Power Engineering
2
Associate Professor, School of Energy and Power Engineering
3
Professor, School of Energy and Power Engineering
4
Ph.D Candidate, School of Energy and Power Engineering

https://doi.org/10.1016/j.jsv.2021.116044
0022-460X/© 2021 Elsevier Ltd. All rights reserved.
L. Han, D. Wei, Y. Wang et al. Journal of Sound and Vibration 501 (2021) 116044

Nomenclature

A amplitude of the airfoil vibration


c chord length
CD drag coefficient, CD = Fdrag /(0.5ρU∞ 2 c)

CL lift coefficient, CL = Flift /(0.5ρU∞ c )


2

Cp pressure coefficient, C p = (P − P∞ )/(0.5ρU∞ 2 )

C̄ p unsteady pressure coefficient, C̄ p = P̄ /(0.5ρU∞ 2 )

D displacement of the airfoil relative to its equilibrium position


fVS natural frequency of the vortex shedding
fVIB frequency of the airfoil vibration
fr ratio of the frequency of the airfoil vibration to the natural frequency of the vortex shedding, fr =fVIB /fVS
G influence coefficient of the airfoil vibration on the unsteady pressure
P pressure
P̄ unsteady pressure amplitude
Q second invariant of velocity gradient tensor, 1/s2
Re Reynolds number
S airfoil area
St Strouhal number
T period of the vortex shedding (also the period of the airfoil vibration in the lock-in region)
t time
t̄ normalized time, t̄ = t/T
U velocity of flow
U∞ velocity of incoming flow
W aerodynamic work
W̄ normalized aerodynamic work
WDensity aerodynamic work density, WDensity = W/S, J/m2
xi Cartesian coordinates in i direction
α airfoil angle of attack
ϕ phase between disturbance and airfoil vibration
ν kinematic viscosity of air
ρ density
ω angular frequency of the airfoil vibration, rad/s, ω=2π fVIB

Subscripts
RR reconstructed results based on the components of the airfoil vibration and vortex shedding
SR results obtained directly from the numerical simulation
VS component of the vortex shedding
VIB component of the airfoil vibration

1. Introduction

Vortex-induced vibration (VIV) is common in bridges, chimneys, tall buildings, wind turbines, aircraft, and turbomachin-
ery. It is a harmful phenomenon that may result in structural failure. For example, in May 2020, the Humen Bridge in
China vibrated at a significant amplitude after roadblocks were installed, and this was generally considered to be a VIV
phenomenon. Another famous example is the collapse of the Tacoma Bridge in 1940, which was also considered a VIV phe-
nomenon by some scholars. The lock-in phenomenon is a typical feature of VIV. A vortex shedding often forms downstream
of a bluff body and its frequency locks onto a structural natural frequency when their values are close. The range of lock-in
generally expands with increasing vibration amplitude and exhibits a V-shaped region.
VIV has been extensively studied in the structure of cylinders [1–3] and flat plates [4] through simplified models [5–9],
time domain methods [10,11] and experiments [12–15]. In most studies, VIV was investigated using a rigid cylinder with a
spring restraint having one or two degrees of freedom (DOFs). In these scenarios, the force can be modeled using a wake
oscillator such as the van der Pol oscillator [8,9]. These phenomenological models are very useful in understanding VIV.
They can frequently capture the typical features of VIV, such as the lock-in phenomenon, and are consistent with experi-
mental results. However, the mechanisms of VIV and lock-in are still unclear. A better understanding requires a combination
of numerical methods and experiments, which can provide more details about the flow field and be applied to more com-
plex models. For example, Sadeghi et al. [16] experimentally analyzed the VIV of long flexible risers, and Papaioannou et al.
[14] and Rabiee et al. [15] investigated the VIV of tandem cylinders (groups of cylindrical structures) using numerical meth-
ods as they are difficult to describe using simplified models.

2
L. Han, D. Wei, Y. Wang et al. Journal of Sound and Vibration 501 (2021) 116044

The VIV and lock-in phenomena have been extensively investigated for the airfoil structure. Owing to the complexity of
the flow field around the airfoil, numerical and experimental methods are more commonly employed. The process of vortex
shedding from an airfoil has been observed to be similar to that of a cylindrical structure [17,18]. A typical pattern is that
vortices form and alternately shed from the leading and trailing edges. From the perspective of angle of attack, the research
on VIV can be crudely divided into scenarios of small angles of attack (lower than the stall angle of attack) [18–25] and high
angles of attack [26–32]. For small attack angles, the airfoil vibration has a significant effect on the vortex shedding, while
for high attack angles, the vortex can naturally be shed from the airfoil.
The two important aspects of VIV are the lock-in phenomenon and vibration amplitude. Besem et al. [12] numerically
demonstrated that two different lock-in regions existed in their two-DOF system, and the vortex shedding modes in the
two lock-in regions were different. Spiker et al. [25] simulated the tip section of compressor blades to analyze the lock-in
phenomenon in turbomachinery. They observed that the lock-in region differed under the two inter-blade phase angles they
investigated. Cagney and Balabani [33] observed that the far wake was not locked onto the vibration frequency while the
near wake was. Hoskoti et al. [24] used the van der Pol oscillator to simulate the wake and compared it with the numerical
results. They observed that the lock-in region was affected by the frequency ratio, reduced frequency, and mass ratio.
Compared with the lock-in phenomenon caused by structural vibration, the reason for the occurrence of VIV and the vi-
brating amplitude have been more widely studied by scholars from the perspective of structural failure. Poirel et al. [21] re-
vealed that boundary layer separation is important for airfoils to maintain vibration. Conversely, the vibration can effectively
restrain the laminar separation, thereby enhancing the lift and reducing the drag [22]. Bhat and Govardhan [29] observed
that the phase difference between the motion of an airfoil and flow is significantly different in the damping and exciting
scenarios, which is the main reason for VIV. Benner et al. [31] conducted an experiment on the NACA 0021 airfoil to in-
vestigate the VIV of vertical-axis wind turbines. Significant vibration occurs at angles of attack between 60° and 130°. The
highest vibration amplitude and widest lock-in region occur at 90° (the vertical incoming flow). Zhu et al. [23] used numer-
ical methods to study the VIV of a 2D thin airfoil at small angles of attack (-10° to 15°), and they proposed that the angle
of attack has a significant effect on exciting the airfoil vibration.
The vibration amplitude has been observed to behave asymmetrically with changing flow velocity [13,15,34–38], and the
maximum amplitude is not always attained when the natural frequency of the vortex shedding is equal to the frequency
of the structure (fr =1) [35–37], which cannot be explained directly by the theory of forced response. The different DOFs of
the structure may also have a noticeable response in the lock-in region. For example, compared with the DOF transverse to
the flow, the in-line DOF contributes slightly to the vibration amplitude, unless the mass ratio and damping are significantly
small [12]. Besem et al. [30] calculated the aerodynamic damping of VIV. Negative damping is obtained on the right side of
the lock-in region, in which the natural frequency of the vortex shedding is lower than the specified frequency of the airfoil
vibration. This means that the vibration amplitude would be further amplified.
VIV is the result of the interaction between a fluid flow and vibration, and its mechanism is still unknown. The effect
of the vibration on the fluid primarily occurs as the lock-in phenomenon, and the effect of the fluid on the structure is
primarily reflected in the response. An analysis of the two aspects of fluid and vibration can contribute to understanding
VIV. To explore the mechanism of VIV, we numerically investigated the NACA 0012 airfoil in this study. By simulating the
airfoil at high angles of attack, a periodically oscillating vortex shedding forms downstream of the airfoil. First, the effect
of the airfoil vibration on the flow was investigated in the lock-in region. Subsequently, based on the flow feature, the
disturbance in the lock-in region was divided into vibration and vortex shedding components using a disturbance-separating
method. Finally, the effect of the two disturbances on the mechanism of VIV was analyzed based on the energy balance. To
the authors’ knowledge, no effective method currently exists to distinguish the two disturbances in the lock-in region. A
method of separating disturbances is important to explore the effect of the airfoil vibration and the vortex shedding on VIV
and to further understand the physical mechanism of VIV.

2. Numerical methods and verification

2.1. Governing equations and numerical methods

An NACA 0012 airfoil with a chord length of 0.25 m was investigated in this study; it was simulated under the conditions
of high angle of attack and low incoming Mach number. Therefore, the flow field can be described using 2D incompressible
Reynolds-averaged Navier–Stokes equations, which are expressed as follows:

∂ Ūi
=0 (1)
∂ xi
     
∂ Ūi ∂ ŪiŪ j 1 ∂P ∂ ∂ Ūi ∂ Ū j
+ =− + ν +
 
− Ui U j (2)
∂t ∂xj ρ ∂xj ∂xj ∂ x j ∂ xi

where t and xi are the time and Cartesian coordinates, respectively; Ui and Ui are the corresponding velocity and fluctuat-
ing velocity components, respectively; and P, ρ , and ν are the pressure, density, and kinematic viscosity, respectively. The
overbar in Eqs. (1) and (2) represents the time average.

3
L. Han, D. Wei, Y. Wang et al. Journal of Sound and Vibration 501 (2021) 116044

A feature of vortex shedding is that its natural frequency cannot be predicted in advance. Therefore, a time marching
computational fluid dynamics solver, instead of a frequency domain code, was employed to compute the unsteady Reynolds-
averaged Navier–Stokes equations. These equations were discretized using a cell-centered control volume method. A high-
resolution spatial discretization scheme was adopted, with a special nonlinear recipe at each node [39]. A second-order
backward Euler scheme was adopted for the transient term, and the k-ε two-equation turbulence model with wall function
[40,41] was used. The unsteady simulation continuously ran until periodic results were achieved from one cycle to another,
which required 8 to 9 cycles from the initial value. To reproduce the unsteady process of vortex shedding, a temporal sen-
sitivity analysis was conducted as is discussed in Section 2.3. Based on the results, each period of the vortex shedding
contained approximately 40 steps to accurately simulate its unsteady process. After achieving the periodic results, 30-40
complete periods were extracted to ensure accuracy of the Fourier transform. Thus, each pressure signal in this study fre-
quently contained approximately 1500 samples. The simulation results were also compared with the experimental data in
Ref. [30] for verification.

2.2. Computational domain and boundary conditions

A C-type computational domain was adopted by extending the chord lengths 20 times downstream and 10 times up-
stream. Three meshes were assessed for mesh independence and are listed in Table 1. In the end, Mesh 2 was selected such
that the computational domain contained 400 nodes in the circumferential direction, 200 nodes in the radial direction, and
200 nodes in the downstream direction. To satisfy the conditions of the turbulence model, the spacing of the first mesh
layer on the airfoil surface was set to 1 × 10−4 m. The details of the mesh refinement study are provided in Section 2.3,
and the computational domain is shown in Fig. 1.
The NACA 0012 airfoil was simulated in a standard atmosphere. A uniform incoming flow (U∞ ) was imposed at the inlet
boundary, and static pressure was specified at the outlet. The no-slip boundary condition was specified at the airfoil surface.
The simulations were conducted under different angles of attack (α ), vibration frequencies (fVIB ), and vibration amplitudes
(A).

2.3. Mesh independence and temporal sensitivity

Mesh independent analysis was conducted for the NACA 0012 airfoil at different angles of attack and Reynolds number
(Re). Re is defined as follows:
cU∞
Re = (3)
ν

Table 1
Mesh independence simulation for the NACA 0012 airfoil.

Node number Total


Scenario mesh
Circumferential direction (airfoil surface) Radial direction Downstream number
Mesh 1 200 100 100 0.4 × 105
Mesh 2 400 200 200 1.6 × 105
Mesh 3 600 300 300 3.6 × 105

Fig. 1. Computational domain and mesh of the NACA 0012 airfoil.

4
L. Han, D. Wei, Y. Wang et al. Journal of Sound and Vibration 501 (2021) 116044

Fig. 2. Comparison of lift and drag coefficients at different angles of attack.

Fig. 3. Comparison of pressure coefficient at different mesh scales (α =40° and Re=105 ).

where U∞ and c are the velocities of the incoming flow and airfoil chord length, respectively. The parameters of c and ν are
constant, thus Re is proportional to U∞ in this study.
Three mesh scales were simulated for mesh independence and are listed in Table 1. The lift coefficient (CL ) and drag
coefficient (CD ) were calculated up to α =15° (Fig. 2). The lift coefficient and drag coefficient of an airfoil are defined as
Flift
CL = (5)
0.5ρU∞ 2c

Fdrag
CD = (6)
0.5ρU∞ 2c

where the lift per unit length (Flift ) is perpendicular to the direction of U∞ , and the drag per unit length (Fdrag ) is parallel
to the direction of U∞ . The lift coefficient and drag coefficient were calculated under the three mesh scales and compared
with the experimental results in Refs. [42] and [43]: The lift first increased with the angle of attack and began to decrease
at approximately α =12.5°. The drag increased with the angle of attack. A comparison of the numerical results of the three
mesh scales for Re=105 indicated that the lift coefficients of Meshes 2 and 3 were approximately the same, and the lift
coefficient of Mesh 1 was significantly different from those of Meshes 2 and 3 (Fig. 2a). A comparison of the experimental
and numerical results of Mesh 2 at Re=106 indicated that they were in close agreement (Fig. 2b). Subsequently, the three
mesh scales were further simulated at α =40° and Re=105 . The time histories of the pressure coefficient (defined as C p =
(P − P∞ )/(0.5ρU∞2 )) near the trailing edge were compared, and the results are shown in Fig. 3a. The normalized time t̄ in

Fig. 3a was defined as t̄ = t/T , where T is the period of the vortex shedding). The distributions of the pressure coefficient
along the chord length were compared, and the results are shown in Fig. 3b. The results of the three mesh scales were

5
L. Han, D. Wei, Y. Wang et al. Journal of Sound and Vibration 501 (2021) 116044

Fig. 4. Frequency spectra of unsteady pressure coefficient at different time resolutions (α =40° and Re=105 ).

Fig. 5. Pressure fluctuation near the airfoil trailing edge at different angles of attack (Re=84500).

almost the same, except near the trailing edge where the results of Mesh 1 differed slightly from those of Meshes 2 and 3.
Thus, we determined that Mesh 2 was reliable and selected it for all the simulations in this study.
The NACA 0012 airfoil was simulated under four different time resolutions. Each unsteady period contained approxi-
mately 20, 40, 80, and 160 steps. The frequency spectra of the unsteady pressure coefficient (defined as C̄ p = P̄ /(0.5ρU∞
2 ))

near the airfoil trailing edge are shown in Fig. 4. The frequency spectra under the four time resolutions were observed to
be almost the same, and only the frequencies of the time resolution containing 20 steps per period were slightly lower
than those of the other three scenarios. Therefore, we determined that 40 steps per unsteady period was a sufficient time
resolution for predicting the unsteady flow.

2.4. Phenomena of natural vortex shedding and frequency lock-in

First, the parameters of the NACA 0012 airfoil were calculated at angles of attack from 10° to 50°, and as shown in Fig. 5,
the periodically oscillating vortex shedding began to appear at 30°. Thus, to investigate the VIV, the airfoil was simulated
at α =40° and 50°. The simulation results at α =40° and different incoming flows were compared with the numerical and
experimental results in Ref. [30] (Fig. 6). The Strouhal number (St) is defined as follows:
fVS csinα
St = (6)
U∞
where fVS and α are the natural frequency of the vortex shedding and the airfoil angle of attack, respectively.
The width of the wind tunnel test section in Ref. [30] was 0.7 m, which had a significant effect on the flow. Therefore,
St was corrected using the empirical formula in Ref. [44]:
St = St(1 − ζ K ) (7)
where K=csinα /Hw, Hw is the width of the wind tunnel test section, and ζ is the corrected coefficient, which was calculated
using the least-squares method to fit the test data at several angles of attack under a certain incoming flow condition. The
results of corrected coefficient ζ and St’ are listed in Table 2.
The Fourier transform results under different Reynolds numbers are shown in Fig. 6a. The frequencies of unsteady flow
were observed to increase with the Reynolds number, the unsteady pressure coefficients changed slightly, and the amplitude

6
L. Han, D. Wei, Y. Wang et al. Journal of Sound and Vibration 501 (2021) 116044

Fig. 6. Frequency of unsteady flow at different Reynolds numbers (α =40°).

Table 2
Corrected coefficients and Strouhal numbers of
the experimental results in Ref. [30].

Reynolds number, Re, × 104 ζ St’

8.5 1.55 0.182


17 1.62 0.152
25 1.56 0.146
34 1.57 0.141
42 1.62 0.140

Fig. 7. Vortex shedding under different magnitudes of the second-order invariant of the velocity gradient tensor, Q (α =40° and Re=84500).

was mainly on the first-order harmonic component. The St results are shown in Fig. 6b, which indicates that St was almost
constant. The St in this study was approximately 0.19, which was slightly higher than that of the corrected experimental
results in Ref. [30]. This error was probably caused by the difference between the computational domain and test section. In
addition, the simulation results in this study were in good agreement with the experimental results in Ref. [30] under low
Reynolds number conditions, which were the investigation conditions in this study. Therefore, the simulation results in this
study were generally reliable.
Subsequently, the instantaneous flow field was described using the second-order invariant of the velocity gradient tensor
(Q), which is expressed as follows:
 Ux     Uy 
 Ux   Ux Ux   Uy 
Q =  Uxy y
Uy
 +  Ux
  xz
z
Uz
 +  Uy
  z
z
Uz

 (8)
x y z y z

here, Ux , Uy , and Uz represent the relative velocity components in Cartesian coordinates. The second invariant Q remains
constant in different coordinate systems, which represents the features of the vortex structure. The flow field was calculated
for Q=1, Q=10, Q=100, Q=10 0 0, and Q=10 0 0 0, which are shown in Fig. 7. The results for Q≥10 0 0 could not fully show
the vortex shedding. The results for Q=1, Q=10, and Q=100 could all show the vortex shedding. A value of Q=100 was
selected. The instantaneous flow field for the fixed airfoil was depicted for Q=100 (Fig. 8). The structure and process of
the vortex shedding were similar for different incoming velocities: the vortices alternately shed off the leading and trailing

7
L. Han, D. Wei, Y. Wang et al. Journal of Sound and Vibration 501 (2021) 116044

Fig. 8. Instantaneous vortex shedding under different Reynolds numbers (Q=100).

Fig. 9. Pressure fluctuation at the monitor in the wake (the position of the monitor is marked in Fig. 1 and the work conditions of the four cases are
marked in Fig. 10).

edges. However, the scale and shedding density of vortices become larger at high Reynolds numbers, which meant that the
strength and frequency of the shedding vortices increased with the incoming flow (also shown in Fig. 6a).
Subsequently, the airfoil was specified to vibrate in pitch about the quarter-chord from the leading edge and analyzed
at α =40° and 50°. The vortices could form and shed off regularly for the two angles of attack, and the two scenarios
exhibited different features of aerodynamic damping in the lock-in region. To investigate the lock-in phenomenon, the NACA
0012 airfoil was specified to vibrate at different frequencies and amplitudes. The oscillating flow field was obtained under
a standard atmospheric environment, α =40° and Re=101400. In the simulation, the amplitude and frequency of the airfoil
vibration were changed, and the results were compared with the experimental results in Ref. [30]. During the simulation
process, the pressure fluctuation at the monitoring point (see Fig. 1) in the wake was used to determine whether lock-in
occurred. In Fig. 9, the time domain signal of the fixed airfoil is on the top left (Fig. 9a), the time domain signal of a case
where the vortex shedding has locked onto the airfoil vibration is on the top right (Fig. 9b), and the time domain signal
of two cases where the vortex shedding are unlocked are at the bottom (Fig. 9c and d). The work conditions of (a)–(d) in
Fig. 9 were marked by (a)–(d) in Fig. 10.

8
L. Han, D. Wei, Y. Wang et al. Journal of Sound and Vibration 501 (2021) 116044

Fig. 10. Lock-in region of the NACA 0012 airfoil (Re=10 0,0 0 0 and α =40°).

Fig. 10 shows the lock-in region with the vibration frequency normalized by the natural frequency of vortex shedding.
This defines the frequency ratio: fr =fVIB /fVS . The black circles represent the numerical simulation results in this study and the
red diamonds describe the experimental data in Ref. [30]. The solid symbols indicate that lock-in occurred, and the hollow
symbols correspond to the unlocked conditions. The black dashed line shows the lock-in region obtained in this study using
the time marching method, the blue dashed line describes the lock-in region obtained using the frequency domain method
in Ref. [30], and the red dashed line is the lock-in region based on the experimental results in Ref. [30]. The time marching
method used in this study presented a V-shaped lock-in region and was in good agreement with the experimental results
under the examined conditions.

3. Effect of the airfoil vibration on the vortex shedding in the lock-in region

3.1. Time-varying history of the vortex shedding

The vortex shedding of the fixed and vibrating airfoils were compared under the conditions described above (standard
atmospheric environment, α =40° and Re=101400). The vibration amplitude was 1.5° (A=1.5°), and the airfoil was prescribed
a frequency equal to a frequency of the vortex shedding (fr =1), which was determined according to the St shown in Fig. 6.
The vortex shedding was depicted for Q=100, and the flow field over a period is shown in Fig. 11. The airfoil was in the
equilibrium position at t̄ =0, 0.5, and 1, at the maximum displacement of the pressure side (the lower side of the airfoil
surface) at t̄ =0.25, and at the maximum displacement of the suction side (the upper side of the airfoil surface) at t̄ =0.75.
The formation and shedding process of the vortex were almost the same for the fixed and vibrating airfoils. The fluid
separated at the leading edge of the suction side, which caused clockwise vortices at the leading edge and anticlockwise

Fig. 11. Instantaneous vortex shedding over a period (Q=100).

9
L. Han, D. Wei, Y. Wang et al. Journal of Sound and Vibration 501 (2021) 116044

Fig. 12. Instantaneous pressure coefficient distribution on the airfoil over a period.

Fig. 13. Instantaneous vortex shedding over a period under different frequency ratios (Q=100).

vortices at the trailing edge. They alternately formed and periodically shed off the airfoil suction side, thus forming a Kármán
vortex street downstream.
The pressure coefficient distribution on the fixed and vibrating airfoils was described at several typical moments over
a period (t̄ =0, 0.25, 0.5, 0.75, and 1, which corresponds to Fig. 11), and the results are shown in Fig. 12. The pressure on
the pressure side was higher than that on the suction side, but the fluctuation in pressure was significantly higher on the
suction side. When approaching the trailing edge of the airfoil, the pressure fluctuation gradually increased. A comparison
of the results of the fixed airfoil (Fig. 12a) and vibrating airfoil (Fig. 12b) indicated that the overall feature of the pressure
fluctuation was still similar, but the pressure fluctuation increased and the pressure difference between the pressure and
suction side of the airfoil became larger for the vibrating airfoil. As the vortex shedding primarily occurred on the suction
side and shed off the trailing edge of the airfoil (Fig. 11), we inferred that the pressure fluctuation on the suction side
and near the trailing edge was primarily caused by the vortex shedding. However, as both the airfoil vibration and vortex
shedding caused pressure fluctuations on the airfoil and they were mixed in the flow field, quantitatively distinguishing the
airfoil vibration or vortex shedding components was difficult. Therefore, determining their effect on VIV is difficult.

3.2. Effect of the frequency ratio on the vortex shedding

The frequency of the airfoil vibration was adjusted to investigate the effect of the frequency ratio on the vortex shedding
and pressure fluctuation under the same condition (standard atmospheric environment, α =40°, Re=101400, and A=1.5°).
Fig. 13 shows the flow field for fr =0.93, 1, and 1.03. For all cases, the airfoil was in the equilibrium position at t̄ =0, 0.5,
and 1, and the airfoil was at the maximum displacement of the pressure and suction sides at t̄ =0.25 and 0.75, respectively.
Since the incoming flow was constant, the formation and shedding process were almost the same and the strength of the
vortex shedding changed little at different frequency ratios. However, the phase difference between the airfoil vibration and

10
L. Han, D. Wei, Y. Wang et al. Journal of Sound and Vibration 501 (2021) 116044

Fig. 14. Unsteady pressure coefficient distribution on the airfoil at different frequency ratios.

Fig. 15. Instantaneous vortex shedding over a period under different vibration amplitudes (Q=100).

vortex shedding changed with the frequency ratio. Compared with the results of fr =1, the phase of the vortex shedding at
fr =0.93 was ahead by approximately 0.25T (π /2), and the phase of the vortex shedding at fr =1.03 lagged by approximately
0.25T (π /2).
The distribution of the unsteady pressure coefficient on the airfoil for different frequency ratios is shown in Fig. 14. The
variation in the frequency ratio significantly changed the distribution of the unsteady pressure on the airfoil, particularly in
the area before the 70% chord length. The unsteady pressure amplitude on the airfoil suction side was higher than that on
the airfoil pressure side and was significantly high near the trailing edge. The unsteady pressure distribution was consistent
with the pressure fluctuation results shown in Fig. 12. Combined with the results shown in Fig. 13, the variation in the
frequency ratio changed the phase of the vortex shedding, which in turn changed the unsteady pressure distribution on the
airfoil.

3.3. Effect of the vibration amplitude on the vortex shedding

The effect of the vibration amplitude on the vortex shedding and the pressure fluctuation in the lock-in region was
investigated by specifying different amplitudes of the airfoil vibration. Similarly, the incoming flow remained the same as
that in Section 3.2 (standard atmospheric environment, α =40° and Re=101400), and the frequency ratio was specified as
fr =1. The vibration amplitudes were 0.5°, 1°, 1.5°, and 3°, and the flow field is shown in Fig. 15. Although the vibration
amplitude changed, the formation and shedding process were almost the same, the strength of the vortex shedding changed
little, and the relative phase between the airfoil vibration and vortex shedding was similar at every instant; that is, the
phase of the vortex shedding did not change with the vibration amplitude. This was different from the results obtained

11
L. Han, D. Wei, Y. Wang et al. Journal of Sound and Vibration 501 (2021) 116044

Fig. 16. Unsteady pressure coefficient distribution on the airfoil at different vibration amplitudes.

by changing the frequency ratio. Similar results were obtained in Ref. [45]: The phase of the lift coefficient on the airfoil
remained almost constant with the vibration amplitude when the frequency ratio was 1.
The distribution of the unsteady pressure coefficient on the airfoil at different vibration amplitudes was further com-
pared. As shown in Fig. 16, as the vibration amplitude increased, the unsteady pressure amplitude increased significantly
at the trailing edge and changed slightly before the 70% chord length. Comparing the results under different vibration am-
plitudes (Fig. 16) and frequency ratios (Fig. 14), increasing the vibration amplitude only increased the unsteady pressure
amplitude near the trailing edge, while the variation of the vortex shedding phase affected the distribution of the unsteady
pressure on the entire airfoil.

4. Analysis of the unsteady pressure in the lock-in region

4.1. Method of separating disturbances

The coupled fluid and structure interaction (FSI) method is frequently used to analyze the VIV phenomenon. The FSI
method presents the complete results of VIV containing both the variation in flow and structure. However, the results are
too complex to understand how the structure and flow affect each other and to reveal the mechanism of VIV. For example,
when the structure vibrates in a fluid, the fluid around may supply additional mass, damping, and stiffness, which will
change the natural frequency of the structure. Thus, the status of vibration changes with the frequency, and the flow state
changes accordingly, which complicates the problem. If we do not know the effect of the incoming flow, frequency ratio and
vibration amplitude, understanding the results obtained using the FSI method is difficult.
The coupling between the structure and fluid is generally considered to be affected by the mass ratio, which is defined
as the ratio of the mass per unit span of the structure to the mass flow per unit span around the structure. Compared
with the structure in water, the mass ratio of airfoils in air is high, and the aerodynamic load minimally affects the mass,
damping and stiffness of the structure. Thus, this study investigated VIV using an uncoupled method to analyze the effect
of vibration of the structure on the vortex shedding process and to reveal the mechanism of VIV.
To investigate the effect of airfoil vibration and vortex shedding on VIV, the unsteady pressure contributed by the airfoil
vibration and the vortex shedding in the lock-in region should be separated. The disturbances cannot be separated directly
by Fourier transform in the lock-in region. Therefore, the following four hypotheses are proposed based on the results above.

(1) The disturbances of the airfoil vibration and shedding flow satisfy the linear superposition relationship.
(2) The pressure fluctuations caused by the airfoil vibration and the vortex shedding can both be expressed as Fourier series
with the same fundamental frequency ω (ω=2π fVIB ).
(3) The unsteady pressure amplitude caused by the vortex shedding (P̄VS ) and its phase (ϕVS ) do not change with the vibra-
tion amplitude (A) when the frequency ratio is 1 (fr =1) and the incoming flow remains constant.
(4) The unsteady pressure amplitude caused by the airfoil vibration (P̄VIB ) is proportional to A, and its phase (ϕVIB ) remains
constant with A when the incoming flow remains constant and fr =1.

Here, we elaborate on the above hypotheses. The flow field containing the separated flow is frequently nonlinear, which
was considered in the simulation. Here, the nonlinear flow is assumed to be a linear superposition of a nonlinear part
containing the separated flow and a part containing the unsteady flow caused by the vibration. The justification for this is
the simulation and analysis results in Section 3, in which we describe the vortex shedding with Q and investigate the effect
of the vibration amplitude and frequency ratio on the vortex shedding. We observed that for the simulated conditions, the
effect of the airfoil vibration on the strength of the vortex shedding was weak. Thus, we assume that the flow can be divided
into a linear superposition of two types of disturbances (hypothesis 1).

12
L. Han, D. Wei, Y. Wang et al. Journal of Sound and Vibration 501 (2021) 116044

In the lock-in region, the airfoil vibrates periodically, and the process of the vortex shedding is similar to that of the
fixed airfoil, which causes approximate harmonic pressure fluctuations. Therefore, we inferred that the pressure fluctuations
caused by the airfoil vibration and the vortex shedding are periodic and have the same period. As is well known, a periodic
disturbance can be identified by its fundamental frequency and corresponding Fourier series (hypothesis 2).
As discussed in Section 3, the unsteady pressure in the lock-in region is caused by the disturbance of the airfoil vibration
and vortex shedding. When the frequency of the airfoil vibration is equal to the natural frequency of the vortex shedding,
that is, fr =1, the process, strength, and phase of the vortex shedding are almost constant with the vibration amplitude. This
phenomenon was also confirmed in the lift coefficient and the unsteady flow field in Ref. [45]. Therefore, hypothesis 3 is
proposed.
When analyzing the problems of aeroelastic stability, aerodynamic damping is generally considered to be equivalent
to viscous damping. Thus, the aerodynamic force introduced by the airfoil vibration is expressed as a linear function of
the vibration velocity. When the vibration frequency is constant, the aerodynamic force is proportional to the vibration
displacement (hypothesis 4).
The unsteady pressure on the airfoil is described as
P = Paverage + P˜ (9)
where P is the pressure, Paverage is the mean pressure, and P˜ is the disturbance on the airfoil. Based on hypotheses 1 and
2, considering only the first-order harmonic of the unsteady pressure, which contributes to the aerodynamic work of the
airfoil, the pressure of node i can be expressed as
P˜i = P̄SR sin(ωt + ϕSR ) = P̄VIB sin(ωt + ϕVIB ) + P̄VS sin(ωt + ϕVS ) (10)
where ω is the angular frequency of the airfoil vibration, and P̄SR and ϕSR represent the unsteady pressure amplitude and
phase of node i, respectively, which could be obtained directly from the simulation results. The unsteady pressure on
the airfoil can be further expressed as the sum of the airfoil vibration and vortex shedding components. To ensure that
Eq. (10) holds for any value of t, P̄SR , ϕSR , P̄VIB , ϕVIB , P̄VS , and ϕVS must satisfy the following relationships:
P̄SR cos(ϕSR ) − P̄VIB cos(ϕVIB ) − P̄VS cos(ϕVS ) = 0 (11)

P̄SR sin(ϕSR ) − P̄VIB sin(ϕVIB ) − P̄VS sin(ϕVS ) = 0 (12)


Based on hypothesis 4, the pressure caused by the airfoil vibration on node i can be expressed as
P̄VIB = GA, G ≥ 0 (13)
where G is the influence coefficient. If results are at two vibration amplitudes (A1 and A2 ), based on Eqs. (11)–(13), we
obtain
P̄SR,1 cos(ϕSR,1 ) − P̄SR,2 cos(ϕSR,2 ) + G(A2 − A1 )cos(ϕVIB ) = 0 (14)

P̄SR,1 sin(ϕSR,1 ) − P̄SR,2 sin(ϕSR,2 ) + G(A2 − A1 )sin(ϕVIB ) = 0 (15)


By adding the square of Eq. (14) to the square of Eq. (15), we obtain

2
P̄SR,1
+ P̄SR
2
,2
− 2P̄SR,1 P̄SR,2 cos(ϕSR,1 − ϕSR,2 )
G= (16)
( A 1 − A 2 )2
Define C1 and C2 as
P̄SR,1 cos(ϕSR,1 ) − P̄SR,2 cos(ϕSR,2 ) P̄SR,1 sin(ϕSR,1 ) − P̄SR,2 sin(ϕSR,2 )
C1 = , C2 =
G ( A1 − A2 ) G ( A1 − A2 )
It follows that if C1 ≥0,
ϕVIB = arcsinC2 (17)
while if C1 <0,
ϕVIB = π − arcsinC2 (18)
Thus, we can now determine P̄VIB and ϕVIB . The unsteady pressure caused by the disturbance of the airfoil vibration is
obtained. Subsequently, substituting P̄VIB and ϕVIB into Eqs. (11) and (12), we obtain
 2  2
P̄VS = P̄SR cos(ϕSR ) − GA1 cos(ϕVIB ) + P̄SR sin(ϕSR ) − GA1 sin(ϕVIB ) (19)

Define D1 and D2 as
P̄SR cos(ϕSR ) − GA1 cos(ϕVIB ) P̄SR sin(ϕSR ) − GA1 sin(ϕVIB )
D1 = , D2 =
P̄VS P̄VS

13
L. Han, D. Wei, Y. Wang et al. Journal of Sound and Vibration 501 (2021) 116044

Fig. 17. Unsteady pressure distribution of the airfoil vibration and vortex shedding on the airfoil at α =40° and Re=101400 (fr =1).

Fig. 18. Unsteady pressure distribution of the airfoil vibration and the vortex shedding on the airfoil at α =50° and Re=92900 (fr =1).

It follows that if D1 ≥0,

ϕVS = arcsinD2 (20)


while if D1 <0,

ϕVS = π − arcsinD2 (21)


The unsteady pressure caused by the disturbance of the vortex shedding is determined by P̄VS and ϕVS . Based on the
calculated values of G, ϕVIB , P̄VS , and ϕVS , the unsteady pressure P̄RR and ϕRR on the airfoil can be reconstructed for other
vibration amplitudes using the follow equation.

P˜RR = P̄RR sin(ωt + ϕRR ) = P̄VIB sin(ωt + ϕVIB ) + P̄VS sin(ωt + ϕVS )

= GAsin(ωt + ϕVIB ) + P̄VS sin(ωt + ϕVS ) (22)

4.2. Unsteady pressure caused by the airfoil vibration and vortex shedding

The disturbance in the lock-in region was analyzed based on the above method under two conditions (one was α =40°,
Re=101400, and fr =1, and the other was α =50°, Re=92900, and fr =1). The results at A=1.5° and 3° were used to determine
the unsteady pressure of the airfoil vibration and vortex shedding disturbances, P̄VIB , ϕVIB , P̄VS , and ϕVS .
The separated unsteady pressures under the two conditions are shown in Figs. 17 and 18. The unsteady pressure ampli-
tude caused by the vortex shedding was significantly higher than that of the airfoil vibration under the vibration amplitudes
examined. Both of these disturbances had higher unsteady pressure amplitudes near the trailing edge. The unsteady pres-
sure amplitude of the airfoil vibration increased with the vibration amplitude. The unsteady pressure amplitude of the vortex
shedding increased significantly with amplitude and formed a peak near the trailing edge. This peak was significantly wider
at α =50° and Re=92900 than that at α =40° and Re=101400.

14
L. Han, D. Wei, Y. Wang et al. Journal of Sound and Vibration 501 (2021) 116044

Fig. 19. Unsteady pressure distribution of the reconstructed amplitude and phase and the simulation results on the airfoil at A=4.5° (fr =1).

The unsteady pressure under other vibration amplitudes was reconstructed using G, ϕVIB , P̄VS and ϕVS , which were cal-
culated based on the results of A=1.5° and 3°. The unsteady pressure for a vibration amplitude of A=4.5° was reconstructed
and compared with the simulation results. As shown in Fig. 19, the distributions of the reconstructed unsteady pressure
amplitude and phase were in good agreement with the overall trend of the simulation results. However, there were some
slight difference. At α =50° and Re=92900, the main difference was that the reconstructed unsteady pressure was slightly
lower than the simulation results on the suction side. At α =40° and Re=101400, the reconstructed unsteady pressure was
slightly lower than the simulation results only at the trailing edge of the suction side.

5. Analysis of the vortex-induced vibration in the lock-in region

5.1. Aerodynamic work at different vibration amplitudes

The aerodynamic work (W) per vibration period of the airfoil can be obtained using

1  j  
nnode ntstep
W = Si Pi + Pi j−1 Dij − Dij−1 (23)
2
i=1 j=1

where nnode and ntstep represent the total number of nodes on the airfoil and total time steps in a vibration period, respec-
tively, Si describes the area of node i on the airfoil, and Di is the displacement of node i relative to its equilibrium position.
Instead of aerodynamic damping, the aerodynamic work W was used to evaluate the aeroelastic stability of the airfoil. The
main reasons were to avoid distortion at small vibration amplitudes or singularity at zero vibration amplitude, and it was
also convenient to analyze the effect of the disturbances on VIV. A positive aerodynamic work means that the aerodynamic
damping is negative and the airfoil is excited. Otherwise, the airfoil is damped.
The total aerodynamic work of VIV was calculated based on the reconstructed unsteady pressure and was compared
against simulation results as shown in Fig. 20. The aerodynamic work is normalized by the absolute aerodynamic work of
PVIB for A=4.5°. PVIB is the unsteady pressure caused by the airfoil vibration, WVIB is the aerodynamic work done by PVIB and
W̄ = W/|WVIB,A=4.5 |. The total aerodynamic work obtained using the reconstructed unsteady pressure was in good agreement
with the simulation results and changed at a rate proportional to the square of the vibration amplitude. The aerodynamic
work decreased as the vibration amplitude increased at α =50° and Re=92900, but it initially increased and then decreased
with the vibration amplitude at α =40° and Re=101400. By analyzing the separated aerodynamic work, we observed that
the difference in the pattern between the two incoming flows was caused by the phase of the vortex shedding.
The aerodynamic work of each disturbance was analyzed based on the separated unsteady pressure above. Because the
frequency ratio was maintained at 1, the phase of the vortex shedding (ϕVS ) remained constant at different vibration ampli-
tudes. The absolute aerodynamic work is shown in Fig. 21. It has been normalized to the maximum value for convenience.
The aerodynamic work of the airfoil vibration disturbance was negative under both incoming flows, which corresponded
to a damping of the airfoil vibration. The aerodynamic work of the airfoil vibration increased at a rate proportional to the
square of the vibration amplitude. The aerodynamic work of the vortex shedding disturbance was positive at α =40° and
Re=101400, which excited the airfoil vibration, and was negative at α =50° and Re=92900, which damped the airfoil vibra-
tion. The aerodynamic work of the vortex shedding disturbance increased linearly with the vibration amplitude.
Fig. 21 shows that the aerodynamic work of the two disturbances intersected with the increase in the vibration am-
plitude. Before the intersection point, the disturbance of the vortex shedding dominated the aerodynamic work, and thus
determined the vibration status of the airfoil. After the intersection, the disturbance of the airfoil vibration determined
whether the airfoil was excited or suppressed. Thus, we understood why the total aerodynamic work initially increased and
then decreased at α =40° and Re=101400. The point of W̄ = 0 in Fig. 20 corresponds to the intersection point in Fig. 21.

15
L. Han, D. Wei, Y. Wang et al. Journal of Sound and Vibration 501 (2021) 116044

Fig. 20. Normalized total aerodynamic work at different vibration amplitudes (the aerodynamic work is normalized by the absolute aerodynamic work of
PVIB under A=4.5° and fr =1).

Fig. 21. Normalized absolute aerodynamic work of the two disturbances at different vibration amplitudes (the aerodynamic work is normalized by the
absolute aerodynamic work of PVIB under A=4.5° and fr =1).

The aerodynamic work density (WDensity = W/S) on the airfoil was studied at A=4.5° for both conditions (the total aero-
dynamic work is negative, W<0) and at A=0.5° for α =40° and Re=101400 (the total aerodynamic work was positive, W>0).
As shown in Fig. 22, the aerodynamic work density changed significantly on the suction side and only slightly on the pres-
sure side. When the vibration amplitude was 4.5°, the reconstructed aerodynamic work density was in good agreement
with the simulation results for both conditions. The area in the middle chord length of the suction side and the area near
the trailing edge of the pressure side provided positive work. The area near the trailing edge of the suction side provided
negative work, which was significantly larger than the positive work. Therefore, the total aerodynamic work was negative at
A=4.5° for both conditions.
When the vibration amplitude was 0.5°, the positive work done by the area in the middle chord length of the suction
side was greater than the negative work done by the area near the trailing edge. Thus, the total aerodynamic work was
positive at A=0.5° at α =40° and Re=101400. However, although both of the reconstructed and simulation results indicated
a positive aerodynamic work at A=0.5° (Fig. 20), an apparent difference was observed near the trailing edge (Fig. 22b), which
was caused by the phase difference of the vortex shedding between the reconstructed and simulation results.

16
L. Han, D. Wei, Y. Wang et al. Journal of Sound and Vibration 501 (2021) 116044

Fig. 22. Distribution of the aerodynamic work density on the airfoil (fr =1).

Fig. 23. Profile of the vortex shedding at different vibration amplitudes (α =40°, Re=101400, and fr =1).

Fig. 24. Effect of the vortex shedding phase on the distribution of the aerodynamic work density.

5.2. Effect of the vortex shedding phase on the aerodynamic work

In the previous section, we argued that the phase of the vortex shedding affected the distribution of the aerodynamic
work density (Fig. 22b) and caused a difference in the pattern between two flows (Fig. 20). In Fig. 23, the flow fields for
A=0.5°, 1.5°, and 3° were described at t̄ =0 (α =40°, Re=101400, and fr =1). Compared with A=1.5° and 3°, which were used
to obtain the unsteady pressure components (G, ϕVIB , P̄VS , and ϕVS ), the phase of the vortex shedding at A=0.5° was slightly
ahead by approximately π /20. After correcting the vortex shedding phase at A=0.5°, ϕ ’VS =ϕ VS +π /20, the flow field of the
three amplitudes were in the same phase. The distribution of the aerodynamic work density on the airfoil was in good

17
L. Han, D. Wei, Y. Wang et al. Journal of Sound and Vibration 501 (2021) 116044

Fig. 25. Phase of the unsteady pressure caused by the vortex shedding on the airfoil under different incoming flows.

Table 3
Variation in the phase
of the vortex shedding
relative for Re=92900.

Re, × 104 ࢞ϕ VS

9.12 -0.1π
9.29 0
9.46 0.1π
9.63 0.2π
9.80 0.25π
9.97 0.35π
10.14 0.4π
10.47 0.5π

agreement with the simulation results (Fig. 24). This proved that the difference between the reconstructed and simulation
results in Fig. 22b was primarily caused by the phase of the vortex shedding.
The significant question is why the phase for A=0.5° was slightly different from the phase for A=1.5° and 3°. In practice,
the natural frequency of the vortex shedding cannot be accurately predicted; thus, fr is not strictly 1. When the vibration
amplitude is large, the interaction between the airfoil vibration and flow is strong, and the lock-in region is wide. Therefore,
the phase of the vortex shedding is not very sensitive to fr . When the vibration amplitude is small, the phase of the vortex
shedding is more sensitive to fr . Therefore, a small error in fr may cause a variation in the vortex-shedding phase (Fig. 23).
The sensitivity of the vortex-shedding phase to fr can be confirmed by comparing the vortex shedding phases under different
vibration amplitudes at the same frequency ratio other than 1, which will not be shown here.
In addition to the aerodynamic work density distribution, the phase of the vortex shedding also affects the pattern of
the aerodynamic work with the vibration amplitude (Fig. 20). The aerodynamic work of the vortex-shedding disturbance
was positive and negative for α =40° and Re=101400 and α =50° and Re=92900, respectively. This difference was caused by
the variation in the vortex-shedding phase. As shown in Fig. 25, the phase had a significant difference of approximately π /3
between the two conditions. The variation in the vortex-shedding phase can change the aerodynamic work significantly.
The phase of the vortex shedding is known to change with the frequency ratio in the lock-in region. Thus, the frequency
ratio was varied to further investigate the effect of the vortex shedding on the aerodynamic work. Two feasible methods
can be used to adjust the numerical frequency ratio. One is to directly specify the frequency of the airfoil vibration. The
other is to adjust the natural frequency of the vortex shedding by changing the incoming flow, which is consistent with
the actual scenario. The second method was selected in this study and the incoming flow was changed under an angle of
attack of 50° and vibration amplitude of 1.5°. The lock-in phenomenon occurred between the Reynolds numbers of 91200
and 104700. As shown in Fig. 26, the phase of the vortex shedding increased with the incoming flow in the lock-in region.
As the unsteady pressure near the trailing edge contributed more aerodynamic work, the moment that the vortex fell off
the airfoil trailing edge was selected as the iconic feature to confirm the phase differences of the vortex shedding under
different incoming flows. The accuracy of the phase differences was mainly determined by the number of time steps in the
vibration period and the error was no more than 0.05π in this study. The variation in the phase relative for Re=92900 is
shown in Fig. 26 and listed in Table 3.
Subsequently, the effect of the vortex shedding phase on the aerodynamic work was investigated. The aerodynamic work
was calculated based on the simulation results under different incoming flows and the reconstructed unsteady pressure at
different phases of the vortex shedding ϕVS . The reconstructed unsteady pressure P̄RR and ϕRR were calculated based on Eq.
(22), where G, ϕVIB , and P̄VS used the values obtained at Re=92900 and ϕVS varied from -π to π based on the ϕVS at
Re=92900. The results are shown in Fig. 27. The simulation results showed that vortex shedding frequency locked onto the
vibration frequency between Re=91200 and 104700, but the unsteady pressure can be reconstructed at any phase based on

18
L. Han, D. Wei, Y. Wang et al. Journal of Sound and Vibration 501 (2021) 116044

Fig. 26. Instantaneous vortex shedding profile when it fell off the airfoil trailing edge under different Reynolds numbers.

Fig. 27. Effect of the vortex shedding phase on the aerodynamic work (the aerodynamic work is normalized by the aerodynamic work at Re=104700 and
A=1.5°).

Eq. (22), which cannot judge whether lock-in occurred. Thus, the reconstructed aerodynamic work was depicted by the blue
solid line where lock-in occurred (-0.1π ≤ ϕVS ≤0.5π ) and plotted by the blue dotted line in the unlocked region. The trend
of the reconstructed aerodynamic work was in agreement with the simulation results (black squares). The total aerodynamic
work became positive when the phase increased by more than 0.22π , corresponding to a Reynolds number of approximately
970 0 0. Positive aerodynamic work will cause the vibration of the airfoil to be amplified.
As we introduced in Section 2.4, the vortex shedding changed with the incoming flow. The constant P̄VS at Re=92900
caused the error between the reconstructed results and the simulation results. A strategy to reduce this error is to correct
the disturbance of the vortex shedding at the corresponding incoming flow or to calculate the above work at a smaller
vibration amplitude, at which the vortex shedding phase is more sensitive to the incoming flow and the lock-in region is
narrower. The first strategy was used here that the fixed airfoil under different incoming flows was simulated to correct the
unsteady pressure amplitude caused by the vortex shedding P̄VS . The unsteady pressure amplitudes P̄VS on the airfoil surface

19
L. Han, D. Wei, Y. Wang et al. Journal of Sound and Vibration 501 (2021) 116044

Fig. 28. Unsteady pressure amplitude of the vortex shedding P̄VS under different incoming flows.

Fig. 29. Prediction of the theoretical limit vibration amplitude (the aerodynamic work is normalized by the absolute aerodynamic work of PVIB at A=17.5°).

under several incoming flows are shown in Fig. 28. P̄VS increased with Re, which meant doing more positive aerodynamic
work than the uncorrected situation. As shown in Fig. 27, the aerodynamic work calculated after correcting P̄VS (red line)
was closer to the simulation results than the uncorrected results.

5.3. Mechanism of the vortex-induced vibration and limited cycle oscillations

As shown in Fig. 21, the aerodynamic work of the airfoil vibration disturbance and vortex shedding disturbance increased
at the square rate of and linearly with the vibration amplitude, respectively. As shown in Fig. 27, the aerodynamic work in
the lock-in region changed with the phase of the vortex shedding. Thus, the vibration states of the airfoil are as follows:
1 When the aerodynamic work of the vortex shedding disturbance is negative (WVS <0), the vibration is damped. Because
the vortex shedding has a greater effect on the aerodynamic work than the airfoil vibration disturbance at a small vibra-
tion amplitude (before the intersection point in Fig. 29), the vibration is suppressed. The system is stable regardless of
the aerodynamic work of the airfoil vibration disturbance (WVIB ).
2 When both the aerodynamic work of the vortex shedding and the airfoil vibration disturbances are positive (WVS >0 and
WVIB >0), the vibration amplitude continues to increase, which is similar to the phenomenon of flutter and belongs to
the category of self-excited instability.
3 When the aerodynamic work of the vortex shedding disturbance is positive and the aerodynamic work of the airfoil
vibration is negative (WVS >0 and WVIB <0), the vibration of the airfoil is initially amplified. When the vibration amplitude
reaches the intersection point in Fig. 29 (|WVS |=|WVIB |), the aerodynamic work becomes balanced and the vibration
amplitude no longer increases. Thus, limit cycle oscillations are presented.
Therefore, VIV is considered to be the third status, which occurs only when the aerodynamic work of the vortex shedding
disturbance is positive and the aerodynamic work of the airfoil vibration is negative (WVS >0 and WVIB <0).
Finally, the theoretical limit vibration amplitude was evaluated in this study. As shown in Fig. 27, we know that the limit
vibration amplitude was obtained when the phase of the vortex shedding performed the maximum positive aerodynamic

20
L. Han, D. Wei, Y. Wang et al. Journal of Sound and Vibration 501 (2021) 116044

work in the lock-in region, not when fr was equal to 1. Therefore, the theoretical limit vibration amplitude can be obtained
by calculating the energy balance of the airfoil vibration and vortex shedding when the vortex shedding performed the
most positive aerodynamic work. The results were described in Fig. 29. Compared with Fig. 21, the aerodynamic work of the
vortex shedding disturbance WVSmax in Fig. 29 (red lines) was calculated at the vortex shedding phase ϕ
VS where the unsteady
pressure of the vortex shedding PVS did the maximum aerodynamic work. At α =40° and Re=101400, the theoretical limit
vibration amplitude was A’1 =2.99°. At α =50° and Re=92900, the theoretical limit vibration amplitude was A’2 =13.4°. If other
forms of damping were considered or the lock-in phenomenon did not occur at the phase of the maximum aerodynamic
work, the limit vibration amplitude would be smaller than the predicted results. In addition, note that compared with fluid–
structure coupling analysis methods, the vibration amplitude obtained by calculating the energy balance was based on the
decoupled method above. Therefore, it cannot directly provide the precise limit amplitude of VIV. However, this method is
faster and more efficient in preliminarily determining the risk of VIV in engineering.

6. Conclusions

The fixed and vibrating NACA 0012 airfoils were simulated in this study to investigate the VIV. In the lock-in region, the
disturbances are separated into the components of the airfoil vibration and vortex shedding to reveal the mechanism of VIV.
The numerical codes were verified by the experimental results in the literatures, and the method of separating disturbances
in the lock-in region was examined using the simulation results. The vortex shedding was described by the second-order
invariant of the velocity gradient tensor Q, which is helpful to distinguish its feature from the airfoil vibration. The main
conclusions are as follows:
(1) Comparing the flow field under different conditions, the process of the vortex shedding is similar. The natural frequency
of the vortex shedding increases with the incoming flow. Comparing the flow field in the lock-in region, the phase
between the vortex shedding and the airfoil vibration changes with the frequency ratio, which may cause a significant
variation in the unsteady pressure. When the frequency ratio is equal to 1 and the incoming flow remains constant, the
phase and strength of the vortex shedding are almost the same under different vibration amplitudes, and the unsteady
pressure amplitude near the trailing edge of the airfoil increases with the vibration amplitude.
(2) A method for separating the disturbance in the lock-in region is presented in this paper. The unsteady pressure is divided
into the components of the airfoil vibration and vortex shedding disturbances. The unsteady pressure caused by the air-
foil vibration is proportional to the vibration amplitude and has a higher amplitude near the trailing edge. The unsteady
pressure caused by the vortex shedding is much higher at the suction side of the airfoil and increases significantly near
the trailing edge.
(3) The aerodynamic work of the airfoil vibration and vortex shedding disturbances are proportional to the square of the
vibration amplitude and vary linearly with the vibration amplitude, respectively. Therefore, the effect of the airfoil vibra-
tion on aerodynamics work gradually exceeds that of the vortex shedding, which means that the dominant parameter
of the aerodynamic work changes from the disturbance of the vortex shedding to the disturbance of the airfoil vibration
as the vibration amplitude increases. The aerodynamic work of the vortex shedding disturbance changes with the phase
of the vortex shedding. In the lock-in region, the vortex shedding performs negative work in some phases and positive
work in some other phases.
(4) The stability of the airfoil primarily depends on the aerodynamic work of the two disturbances. The occurrence of VIV
must satisfy the condition that the disturbance of the vortex shedding performs positive work and the disturbance of
the airfoil vibration performs negative work. If the aerodynamic work of the vortex shedding disturbance is negative,
the vibration would be suppressed. If the aerodynamic works of the two disturbances are both positive, the vibration
amplitude will continue to increase, which is similar to the phenomenon of flutter. Regardless of other forms of damping
and under the assumption of linear superposition, the theoretical limit vibration amplitude of VIV can be preliminarily
obtained by calculating the energy balance. For the airfoil in this study, the theoretical limit vibration amplitudes under
the two conditions examined were 3° and 13.4°, respectively.

Declaration of Competing Interest

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

CRediT authorship contribution statement

Le Han: Conceptualization, Methodology, Software, Formal analysis, Investigation, Writing – review & editing. Dasheng
Wei: Data curtion, Visualization, Writing – original draft. Yanrong Wang: Resources, Supervision, Funding acquisition.
Xiaojie Zhang: Validation.

Acknowledgment

This work was supported by the National Nature Science Foundation of China (Grant No. 51475022) .

21
L. Han, D. Wei, Y. Wang et al. Journal of Sound and Vibration 501 (2021) 116044

References

[1] A. Khalak, C. Williamson, Motions, Motions, forces and mode transitions in vortex-induced vibrations at low mass-damping, J. Fluids Struct. 13 (1999)
813–851.
[2] C. Williamson, R. Govardhan, A brief review of recent results in vortex-induced vibrations, J. Wind Eng. Ind. Aerodyn. 96 (6) (2008) 713–735.
[3] R.D. Gabbai, H. Benaroya, An overview of modeling and experiments of vortex-induced vibration of circular cylinders, J. Sound Vib. 282 (3) (2005)
575–616.
[4] J. Chen, Y.C. Fang, Lock-on of vortex shedding due to rotational oscillations of a flat plate in a uniform stream, J. Fluids Struct. 12 (6) (1998) 779–798.
[5] R. Landl, A mathematical model for vortex-excited vibrations of bluff bodies, J. Sound Vib. 42 (2) (1975) 219–234.
[6] P. Bearman, Circular cylinder wakes and vortex-induced vibrations, J. Fluids Struct. 27 (5) (2011) 648–658.
[7] R. Ogink, A. Metrikine, A wake oscillator with frequency dependent coupling for the modeling of vortex-induced vibration, J. Sound Vib. 329 (26)
(2010) 5452–5473.
[8] M.L. Facchinetti, E.D. Langre, F. Biolley, Coupling of structure and wake oscillators in vortex-induced vibrations, J. Fluids Struct. 19 (2) (2004) 123–140,
doi:10.1016/j.jfluidstructs.20 03.12.0 04.
[9] P.A. Opinel, N. Srinil, Application of wake oscillators to two-dimensional vortex-induced vibrations of circular cylinders in oscillatory flows, J. Fluids
Struct. 96 (2020) 103040.
[10] M. Thorsen, S. Savik, C. Larsen, A simplified method for time domain simulation of cross-flow vortex-induced vibrations, J. Fluids Struct. 49 (2014)
135–148.
[11] M. Thorsen, S. Savik, C. Larsen, Time domain simulation of vortex-induced vibrations in stationary and oscillating flows, J. Fluids Struct. 61 (2016)
1–19.
[12] F. Besem, J. Thomas, R. Kielb, et al., An aeroelastic model for vortex-induced vibrating cylinders subject to frequency lock-in, J. Fluids Struct. 61 (2016)
42–59.
[13] W.L. Chen, F. Xu, Investigation of a hybrid approach combining experimental tests and numerical simulations to study vortex-induced vibration in a
circular cylinder, J. Sound Vib. 331 (5) (2012) 1164–1182.
[14] G.V. Papaioannou, D.K.P. Yue, M.S. Triantafyllou, et al., On the effect of spacing on the vortex-induced vibrations of two tandem cylinders, J. Fluids
Struct. 24 (6) (2008) 833–854.
[15] A.H. Rabiee, M. Esmaeili, Simultaneous vortex- and wake-induced vibration suppression of tandem-arranged circular cylinders using active feedback
control system, J. Sound Vib. 469 (2020) 115131.
[16] M.Y. Sadeghi, F. Chasparis, M.S. Triantafyllou, et al., Chaotic response is a generic feature of vortex-induced vibrations of flexible risers, J. Sound Vib.
330 (2011) 2565–2579.
[17] J. Gostelow, W. Carscallen, M. Platzer, On vortex formation in the wake flows of transonic turbine blades and oscillating airfoils, ASME, 2005 Paper
No. GT2005-69128, doi:10.1115/GT2005-69128.
[18] J. Young, J.C.S. Lai, Oscillation frequency and amplitude effects on the wake of a plunging airfoil, AIAA J. 42 (10) (2004) 2042–2052.
[19] J. Young, J.C.S. Lai, Vortex lock-in phenomenon in the wake of a plunging airfoil, AIAA J. 45 (2) (2007) 485–490.
[20] M.A. Ashraf, J. Young, J.C.S. Lai, Oscillation frequency and amplitude effects on plunging airfoil propulsion and flow periodicity, AIAA J. 50 (11) (2012)
2308–2324.
[21] D. Poirel, V. Métivier, G. Dumas, Computational aeroelastic simulations of self-sustained pitch oscillations of a NACA0012 at transitional Reynolds
numbers, J. Fluids Struct. 27 (8) (2011) 1262–1277.
[22] J.M. Lei, J.M. Zhang, J.P. Niu, Effect of active oscillation of local surface on the performance of low Reynolds number airfoil, Aerosp. Sci. Technol. 99
(2020) 105774.
[23] B. Zhu, J. Lei, S. Cao, Numerical simulation of vortex shedding and lock-in characteristics for a thin cambered blade, ASME J. Fluids Eng. 129 (10)
(2007) 1297–1305.
[24] L. Hoskoti, D. Aravinth, A. Misra, et al., Frequency lock-in during nonlinear vibration of an airfoil coupled with van der Pol oscillator, J. Fluids Struct.
92 (2020) 102776.
[25] M. Spiker, R. Kielb, J. Thomas, K.C. Hall, Application of enforced motion to study 2-d cascade lock-in effect, AIAA, 2009 Paper No. 2009-892, doi:10.
2514/6.2009-892.
[26] D. Poirel, L. Goyaniuk, A. Benaissa, Frequency lock-in in pitch-heave stall flutter, J. Fluids Struct. 79 (2018) 14–25.
[27] D. Tang, E. Dowell, Experimental aerodynamic response for an oscillating airfoil in buffeting flow, AIAA J. 52 (6) (2014) 1170–1179.
[28] Z.Y. Li, L.H. Feng, J.J. Wang, Individual influence of pitching and plunging motions on flow structures over an airfoil during dynamic stall, Chin. J.
Aeronaut. 33 (3) (2019) 840–851.
[29] S. Bhat, R. Govardhan, Stall flutter of NACA0012 airfoil at low Reynolds numbers, J. Fluids Struct. 41 (2013) 166–174.
[30] F. Besem, J. Kamrass, J. Thomas, et al., Vortex-induced vibration and frequency lock-in of an airfoil at high angles of attack, ASME J. Fluids Eng. 138
(2015) 011204.
[31] B. Benner, D. Carlson, B. Seyed-Aghazadeh, et al., Vortex-induced vibration of symmetric airfoils used in vertical-axis wind turbines, J. Fluids Struct.
91 (2019) 102577.
[32] A. Gross, M. Agate, J. Little, et al., Numerical simulation of plunging wing section at high angles of attack, AIAA J. 56 (7) (2018) 2514–2527.
[33] N. Cagney, S. Balabani, Wake modes of a cylinder undergoing free streamwise vortex-induced vibrations, J. Fluids Struct. 38 (2013) 127–145.
[34] L.B. Zhang, A. Abdelkefi, H.L. Dai, et al., Design and experimental analysis of broad band energy harvesting from vortex-induced vibrations, J. Sound
Vib. 408 (2017) 210–219.
[35] Y.X. Ma, W.H. Xu, T. Pang, et al., Dynamic characteristics of a slender flexible cylinder excited by concomitant vortex-induced vibration and time-vary-
ing axial tension, J. Sound Vib. 485 (2020) 115524.
[36] Y. Qu, A.V. Metrikine, A wake oscillator model with nonlinear coupling for the vortex-induced vibration of a rigid cylinder constrained to vibrate in
the cross-flow direction, J. Sound Vib. 469 (2020) 115161.
[37] W. Wang, B.W. Song, Z.Y. Mao, et al., Numerical investigation on vortex-induced vibration of bluff bodies with different rear edges, Ocean Eng. 197
(2020) 106871.
[38] O. Shoshani, Deterministic and stochastic analyses of the lock-in phenomenon in vortex-induced vibrations, J. Sound Vib. 434 (2018) 17–27.
[39] T.J. Barth, D.C. Jesperson, The design and application of upwind schemes on unstructured meshes, AIAA, 1989 Paper No. 1989-0366, doi:10.2514/6.
1989-366.
[40] B.E. Launder, B.I. Sharma, Application of the energy-dissipation model of turbulence to the calculation of flow near a spinning disc, Lett. Heat Mass
Transf. 1 (2) (1974) 131–137.
[41] B.E. Launder, D.B. Spalding, The numerical computation of turbulent flows, Comput. Methods Appl. Mech. Eng. 3 (2) (1974) 269–289.
[42] W.J. McCroskey, K.W. McAlister, L.W. Carr, et al., An experimental study of dynamic stall on advanced airfoil sections, Summary of the Experiment, 1,
NASA, Washington, DC, 1982 NASA-TM-84245.
[43] S. Mittal, P. Saxena, Hysteresis in flow past a NACA 0012 airfoil, Comput. Methods Appl. Mech. Eng. 191 (2002) 2179–2189.
[44] J. Chen, Y.C. Fang, Strouhal numbers of inclined flat plates, J. Wind Eng. Ind. Aerodyn. 61 (1996) 99–112.
[45] T. Khan, J. Zhang, W. Li, et al., Local vibrations and lift performance of low Reynolds number airfoil, Propul. Power Res. 6 (2) (2017) 79–90.

22

You might also like