Next Article in Journal
A Digital Twin of the Trondheim Fjord for Environmental Monitoring—A Pilot Case
Previous Article in Journal
Detection and Tracking of Low-Frame-Rate Water Surface Dynamic Multi-Target Based on the YOLOv7-DeepSORT Fusion Algorithm
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Force Element Analysis in Vortex-Induced Vibrations of Side-by-Side Dual Cylinders: A Numerical Study

1
College of Mechanical and Marine Engineering, Beibu Gulf University, Qinzhou 535011, China
2
Guangxi Key Laboratory of Ocean Engineering Equipment and Technology, Qinzhou 535011, China
3
Key Laboratory of Beibu Gulf Offshore Engineering Equipment and Technology (Beibu Gulf University), Qinzhou 535011, China
4
School of Mechanical and Electronic Engineering, Pingxiang University, Pingxiang 337055, China
5
Faculty of Civil Engineering and Mechanics, Kunming University of Science and Technology, Kunming 650500, China
6
Institute of Applied Mechanics, College of Engineering, National Taiwan University, Taipei 10764, Taiwan
*
Authors to whom correspondence should be addressed.
J. Mar. Sci. Eng. 2024, 12(9), 1529; https://doi.org/10.3390/jmse12091529
Submission received: 21 July 2024 / Revised: 30 August 2024 / Accepted: 31 August 2024 / Published: 3 September 2024
(This article belongs to the Section Ocean Engineering)

Abstract

:
A numerical investigation was conducted in this study utilizing Force Element Analysis to explore the vortex-induced vibration (VIV) mechanism of side-by-side dual cylinders under the conditions of Reynolds number Re = 100, mass ratio m* = 10, and spacing ratios L/D ranging from 3 to 6. The hydrodynamic forces by force element formulas were incorporated into the vibration response calculations of elastically supported rigid cylinders using a User-Defined Function (UDF) and the fourth-order Runge–Kutta method. A comprehensive analysis was performed to elucidate the combined effects of the spacing ratio L/D and reduced velocity Ur on the vibration responses, quantifying the hydrodynamic forces involved in the mutual interaction during VIV for side-by-side dual cylinders. The influence mechanisms of inter-cylinder interaction and their effects on the resultant hydrodynamic phenomena were discussed. It was revealed that for side-by-side arranged dual cylinders outside the “lock-in region”, the lift and drag forces are predominantly supplied by the volume vorticity forces in conjunction with surface vortices (including frictional) forces. However, within the “lock-in region”, the surface acceleration lift forces provide greater force contributions, and the volume vorticity lift force contributes significantly to negative values. Notably, alterations to the spacing ratio do not change the proportion of force element components. The amplitudes of the cylinders’ mutual interaction forces are identical in magnitude but opposite in phase. Additionally, the “slapping” phenomenon near the “lock-in region” leads to “bounded” trajectories of cylinders.

1. Introduction

Vortex-induced vibration (VIV), a significant topic, that plays a pivotal role in engineering applications involving bluff body structures, such as bridge stay cables, marine risers, and heat exchanger tube bundles, profoundly impacts structural stability, fatigue damage, and overall performance [1]. Under specific reduced velocities, contingent upon the system’s mass ratio and damping ratio, the frequency of vortex shedding matches the cylinder’s natural frequency ( f n = ( 1 / 2 π ) k / m , where k is the system stiffness and m is the mass of the cylinder), leading to large-amplitude oscillations, a phenomenon known in the literature as “lock-in” [2].
The presence of two closely spaced cylinders in a fluid flow has garnered considerable attention and has been extensively investigated. As a prototypical configuration of multi-cylinder arrangements, side-by-side dual cylinders exhibit pronounced interference effects on the wake vortex shedding flow [3,4,5]. The VIV characteristics of such configurations are intricately influenced by a complex interplay of vortex-shedding modes, flow field interference effects, and geometric parameters [1,6]. A thorough exploration of these factors is critical for gaining insight into and controlling the VIV behavior of such structures.
Early research predominantly focused on stationary dual cylinders (Alam and Zhou [7], Zhao and Cheng [8], Bai et al. [9]), and single-cylinder vortex-induced vibration (Feng [10], Bearman [11], Williamson [1,6], Williamson and Govardhan [12], Sarpkaya [13], Gabbai and Benaroya [14], Zhao and Cheng [15], Wu et al. [16]). Zdravkovich [4] explored the VIV of two cylinders, revealing that the amplitude of oscillation is largely contingent on the relative positioning of the two cylinders. Williamson [17] observed that within a certain gap range, the wakes of side-by-side dual cylinders synchronize, in-phase or out-of-phase. However, when the distance between two cylinders is less than 2.2D (D is the diameter), only a single wake forms [17,18,19]. Cui et al. [20] investigated the VIV of elastically coupled side-by-side dual cylinders in a cross-flow direction. Chen et al. [21] comprehensively summarized six discernible wake patterns for VIV of two side-by-side cylinders at a Reynolds number Re = 100 and spacing ratios in the range of 2 ≤ L/D ≤ 5 (L is the center-to-center spacing of cylinders). These patterns encompassed irregular, in-phase and out-of-phase flapping, in-phase and anti-synchronous, and biased anti-synchronous modes. Rahmanian [22] studied the VIV of side-by-side dual cylinders with different diameters, delving into the effects of their interaction during the VIV process. Xu et al. [23] systematically examined the impact of the spacing ratio L/D on the two-degrees-of-freedom (2DOF) VIV of side-by-side dual cylinders at Re = 1470 to 10320. They found that even at high spacing ratios of up to 10, the interference between cylinders remained non-negligible for the vibration response; at the same time, six wake patterns, similar to Chen et al. [21], were identified. Chen et al. [24] investigated the VIV of 2DOF side-by-side dual cylinders at Re = 100 and 2 ≤ L/D ≤ 5, exploring the characteristics of hydrodynamic forces, trajectories, phase lags, and wakes in depth. They categorized the response of two cylinders into four distinct branches. Subsequently, Chen et al. [25] examined the effect of unilateral hysteresis of two laterally vibrating cylinders at Reynolds numbers of 60 to 200.
In previous studies, whether dealing with rigidly coupled or individually oscillating elastically mounted cylinders, limitations existed in that they could only provide macroscopic insights into the variations in lift and drag forces across the entire flow field, lacking a comprehensive understanding of the specific impacts of the flow field on object forces. For dual cylinders, for instance, there was no way to ascertain the magnitude of the mutual influence between cylinders. To address this, our group [26] introduced a force theory for incompressible flows around multiple solids, enabling us to examine the force contributions of individual fluid elements to each object. This methodology stems from the Force Element Theory (FET) proposed by Chang [27], which originates from d’Alembert’s principle and represents a discretization-based fluid–structure interaction theory. By dividing the cylinder surfaces into a series of discrete elements (force elements), the hydrodynamic forces exerted on each element by the flow field can be simulated, and when combined with structural dynamics models, the full VIV process can be effectively simulated.
The Force Element Theory has been widely employed in various dynamic-related applications and has gained recognition among scholars (Lee et al. [28], Lin [29], Menon and Mittal [30,31], Yin et al. [32], Luo et al. [33], Xu et al. [34], Moriche et al. [35]). Lin et al. [29] employed the Force Element Theory to analyze the relationship between vorticity in the wake region and total thrust contribution across different regions for a three-dimensional heaving flexible plate. This approach allowed for a deeper understanding of the mechanism behind thrust generation in flexible plates. The Force Element Theory was applied to investigate the forces acting on cylinders and spheres in Stokes accelerating flows at low Reynolds numbers by Xu et al. [34]. They discovered that at extremely low Reynolds numbers, the vorticity within the flow domain contributes almost nothing to the drag force. It is worth noting that both studies utilized the Force Element Analysis to gain insights into the dynamics of fluid–structure interactions at varying scales. In this study, we will leverage the Force Element Analysis to elucidate the VIV mechanisms of side-by-side dual cylinders under different L/D and reduced velocities Ur (here, Ur = U/fnD, U is the free-stream velocity). This will deepen our understanding of their dynamic response characteristics and provide theoretical foundations for structural design, vibration control, and safety assessments in relevant engineering domains.
This study initiates with an introduction to the fluid control equations, which include the Navier–Stokes equations, force element formula, and control equations for VIV systems, complemented by details on the simulation model and fundamental parameters. A grid independence study is then performed. Subsequently, to validate the accuracy of the vortex force formula, pertinent literature verification and analysis are conducted, concentrating on the VIV phenomenon of side-by-side dual cylinders with 2DOF. Frequency and amplitude responses, hydrodynamic coefficients, motion paths, and wake patterns specific to each cylinder are discussed. Ultimately, conclusions with principal findings are presented.

2. Control Equations and Numerical Methods

2.1. Flow Field Setting

The computational domain and boundary conditions are established as depicted in Figure 1. The outer boundary of the computational domain is rectangular, and the streamwise and cross-flow directions are Lx = 150D and Ly = 100D. The distance from the cylinder centers to the inlet boundary is 50D and to outlet boundary is 100D. The two cylinders are symmetrically positioned with respect to each other, with the center-to-center distance between them and the top and bottom boundaries being 50D. This arrangement ensures that the influence of the nearby walls is minimized, allowing a well-developed wake behind the cylinders and small computational error [36]. For the simulation setup, the inlet is specified as a velocity inlet where Dirichlet boundary conditions are applied, setting the velocity components as u = U (streamwise component) and v = 0 (cross-flow component). The outlet is set as Neumann boundary conditions, representing a zero-pressure condition. The top and bottom boundaries are set as symmetry boundaries. No-slip conditions are enforced on the surfaces of both cylinders. The fluid medium within the domain is water.

2.2. Control Equations

In this study, the flow is assumed to be two-dimensional (2D), incompressible, and laminar. Thus, the flow is governed by the non-dimensionalized Navier–Stokes equations:
v t + ( v ) v = P + 1 R e 2 v
and the continuity equation as:
v = 0
where v represents the non-dimensional velocity vector; P = P*/ρU2 is the non-dimensional pressure, where P* is the pressure and ρ is the density; Re = ρUD/μ is the Reynolds number, where μ is the dynamic viscosity of water; and t = τU/D is non-dimensional time, where τ is flow time.
Consider a viscous flow field of an incompressible and uniform incoming flow passing through two bodies, designated as C1 and C2. The surfaces of these two bodies are represented by S1 and S2, respectively, and SR denotes the boundary of the far-field domain.
In the Pressure Method (PM), it is typically necessary to first determine the pressure distribution exerted by the fluid and the frictional forces resulting from viscous effects on the surface of the object. These forces distributed over the surface of the object are then integrated to obtain the total force, which can be subsequently decomposed into a component parallel to the direction of motion, representing drag or thrust, and a component perpendicular to the direction of motion, representing lift. As an example, the force coefficient on cylinder C1 surface S1 in i-direction can be expressed as follows:
C t o t a l 1 = S 1 P ( n i ) d A + 1 R e S 1 n × ω   i d A
Here, C t o t a l 1 represent the total hydrodynamic force coefficient in i-direction, and C t o t a l 1 = F t o t a l 1 / 0.5 ρ U 2 D , F t o t a l 1 is the total hydrodynamic force.
On the other hand, we may take another viewpoint from the perspective of the vorticity force. This is particularly useful for separated flow, mainly featured by its distribution of vorticity in the flow field. Based on Chang et al. [26], it is necessary to introduce an auxiliary potential flow ϕ in the moving direction of the object in the flow field, where ϕ rapidly diminishes as it moves away from the object concerning the force exerted on cylinder C1 in the i-direction. The parameter ϕ is defined as follows:
2 ϕ = 0 o n   V R n ϕ = n i o n   S 1 n ϕ = 0 o n   S 2 ϕ = 0 a t   i n f i n i t y
The force element equation that quantifies the hydrodynamic forces acting on C1 in the i-direction can be deduced from Equation (1) by taking inner products with ϕ :
C t o t a l 1 = S 1 ϕ v t   n d A + 1 2 S 1 v 2   ϕ n d A V R v   × ω ϕ d V + 1 R e S 1 n × ω   ϕ + i d A + S 2 ϕ   v t n d A + 1 R e S 2 n   × ω ϕ d A
here, the first term on the right-hand side of the equation signifies the magnitude of the acceleration forces acting on the surface S1 of C1. The second term denotes the contribution of the velocity on the surface S1 of C1. Please note that in this study, the cylinders are symmetric, and this term is zero. The third term denotes the contribution of the volume vorticity in the flow field to the forces on C1. The fourth term accounts for the combined effect of surface vorticity (including frictional) forces on the surface S1. The fifth term reflects the influence of the acceleration at the surface S2 of C2 on the forces experienced by C1. Lastly, the sixth term captures the impact of the surface vorticity at the surface S2 of C2 on the forces acting on C1.
In this case, i denotes the drag direction, C t o t a l 1 becomes C d 1 , and the five terms on the right-hand side of the Equation (5) become Cda1, Cdv1, Cds1 Cda21, and Cdv21, so Equation (5) can be written as:
C d 1 = C d a 1 + C d ν 1 + C d s 1 + C d a 21 + C d ν 21
where Cda1, Cdv1, and Cds1 denote force coefficients of surface acceleration, the volume vorticity, and the surface vorticity (including frictional) acting on cylinder C1 in the drag direction, respectively. The subscript ‘1’ signifies forces pertinent to cylinder C1. Cda21 and Cdv21 represent force coefficients of surface acceleration and surface vorticity acting on cylinder C1 in the drag direction, which represent the influence of cylinder C2 on cylinder C1.
Similarly, for forces acting in the lift direction, Equation (5) can be written as:
C l 1 = C l a 1 + C l ν 1 + C l s 1 + C l a 21 + C l ν 21

2.3. Vibration Response System Model

In this study, the VIV model of cylinders is simplified to a mass-spring-damper system, permitting the cylinder to vibrate freely in both streamwise and cross-flow directions. The vibration governing equations of the cylinder are as follows:
x ¨ i = F D i t m c m x ˙ i k m x i
y ¨ i = F L i t m c m y ˙ i k m y i
where m is the mass of the cylinder (m = m/mf, where mf is the mass of the displaced fluid), c = 2mω0ζ is damping coefficient, where ω0 is the natural circular frequency, ζ is damping ratio, x is the displacement of the cylinder in the streamwise direction, x ˙ is the velocity of the cylinder in the streamwise direction, and x ¨ is the acceleration of the cylinder in the streamwise direction. Similarly, y , y ˙ , and y ¨ are the displacement, velocity, and acceleration in the cross-flow direction, respectively. As a function of time, F D i t and F L i t are the drag force in the streamwise direction and the lift force in the cross-flow direction, respectively.
The computational process for the VIV response using the Force Element Analysis involves several critical steps. Initially, the Force Element Analysis is coupled with the system vibration equations through User-Defined Functions (UDFs), which are then integrated into the ANSYS-FLUENT computational platform by the Finite Volume Method (FVM). The hydrodynamic force elements are computed. Subsequently, these hydrodynamic forces, which stem from the solutions of Equation (5), representing the forces in streamwise and cross-flow directions, are transferred into Equations (8) and (9) for vibration response calculations. A fourth-order Runge–Kutta method [15] is employed to discretize and solve Equations (8) and (9), effectively capturing the dynamic behavior of the system under consideration. The utilization of the CG_MOTION macro is to relay the calculated vibration response data back to the flow field for dynamic mesh updates. The motion of the cylindrical structure’s boundary is facilitated through the dynamic meshing technique. The movement of the grid is governed by diffusion equations, ensuring that the grid adapts to the boundary motion while preserving its quality. For detailed insights into the boundary motion diffusion method, one may refer to Han et al. [37].
This comprehensive approach ensures an accurate representation of the complex interaction between the fluid flow and the structural motion of cylinders, providing a reliable framework for studying VIV phenomena at the specified parameter ranges. The UDF and the dynamic mesh technique are essential approaches in achieving high-fidelity VIV simulations of cylinders under varying flow conditions.

2.4. Numerical Verification

2.4.1. Flow Meshing

As shown in Figure 2, the grid refinement around cylinders follows a “C”-type pattern to capture the complex flow phenomena accurately. Local grid refinement zones are placed around each cylinder to better observe the vortex-induced vibration effects and the vortex shedding behavior. Each cylinder’s surface is discretized with 240 nodes [38]. The grid expansion ratio is set below 1.05, employing an exponential growth method. The smallest grid size, which is adjacent to the cylinder surface, is characterized by a nondimensionalized height h* = hfirst/D = 0.005, where hfirst is near-wall first grid height and y+ < 1. This grid configuration and boundary condition setup are designed to ensure that the simulations accurately resolve the flow physics while maintaining computational efficiency at the same time.

2.4.2. Independence Verification

As shown in Table 1, a grid independence study was conducted in the VIV of side-by-side cylinders with 2DOF. The computational parameters were set as Re = 100, m* = 2, L/D = 3, Ur = 6, and ζ = 0. Here, A x * = A x / D denotes the dimensionless amplitude in the x-direction, where A x is the streamwise amplitude and A y is the cross-flow amplitude, and A y * = A y / D represents the dimensionless amplitude in the y-direction.
The results reveal that the outcomes obtained with a first-layer grid spacing of h* = hfirst/D = 0.005 were in excellent agreement with those from a finer grid spacing of h* = 0.0025, with a maximum discrepancy of merely 1.6%. This suggests that the simulation accuracy was satisfactory with a grid spacing of h* = 0.005. Furthermore, the independence of the simulation with respect to the time step was confirmed, where results with a dimensionless time step of ∆tU/D = 0.0025 matched very closely with those from a finer time step of ∆tU/D = 0.00125, with a maximum difference of only 0.6%. Consequently, a dimensionless time step of ∆tU/D = 0.0025 was selected for the simulations.
Comparisons with results reported by Chen et al. [24] have also been made, showing a favorable agreement, which further validates the reliability of the present numerical approach. These analyses ensure that the computational parameters chosen provide accurate and robust solutions for the VIV problem under investigation.

2.4.3. Model and Method Validation

In this section, the force element formulation was utilized to validate the numerical methodology for the VIV of side-by-side cylinders with 2DOF. The simulation parameters were consistent with those used by Chen et al. [24], including Re = 100, m* = 2, ζ = 0, and L/D = 3. A comparison of the computed results with those from the literature, as illustrated in Figure 3, reveals that the trends in the mean lift coefficient C ¯ l and the mean drag coefficient C ¯ d of the two cylinders as functions of the reduced velocity Ur were in good agreement. This indicates that the numerical approach adopted could be capable of providing accurate predictions of VIV responses for side-by-side cylinders. Consequently, Force Element Analysis and the vibration model employed herein were validated as suitable tools for studying such phenomena.
Concurrently, we present the simulation results for VIV of a single cylinder at Re = 100 and Ur = 5, as depicted in Figure 4. It is observed that the lift and drag coefficients obtained by solving the Navier–Stokes equations using the Force Element Analysis (FEA) align almost perfectly with those derived from the Pressure Method (PM), which further proves the accuracy of our computational approaches in capturing the hydrodynamic forces associated with VIV phenomena.

3. Results and Discussion

This section focuses on the numerical simulation of VIV for side-by-side cylinders oscillating independently in the streamwise and cross-flow directions at Re = 100, m* = 10, and ζ = 0 to gain higher amplitude responses. The computations encompass a range of spacing ratios L/D is 3 to 6 and reduced velocities Ur spanning from 2 to 12.

3.1. Frequency and Amplitude Response

3.1.1. Frequency Response

The structural response of the cylinder is dominated by its natural frequency and vortex shedding frequency. When the system’s natural frequency fn is close to the vortex shedding frequency fs, a phenomenon known as “lock-in” occurs, characterized by a range of Ur. Within this region, the dimensionless frequency f* = fs/fn ≈ 1. This involves a synchronization mechanism or locking mechanism that leads to large-amplitude oscillations of each cylinder, as described by De Lange [39] and Mittal and Kumar [40]. This synchronization is a critical aspect of VIV phenomena, where the fluid–structure interaction results in enhanced vibrations when the frequencies align. Understanding the “lock-in” region is crucial for predicting and mitigating excessive vibrations in structures subject to fluid flows.
Under the parameters of the VIV response of cylinders for low Reynolds number, following the classification by Chen et al. [24], the vibration response is divided into three phases before, during, and after the “lock-in” region: the initial branch (IB), the lower branch (LB), and the desynchronization branch (DB). The maximum amplitude occurs on LB. The classification of vibration response into these branches is critical for understanding the complex dynamics of VIV. The initial branch typically corresponds to a region where the vibration frequency is close to the natural frequency of the system, while the lower branch is marked by significant amplitude amplification due to the “lock-in” phenomenon. The desynchronization branch, after the “lock-in” region, is characterized by a departure from resonance, resulting in decreased amplitudes and often complex, multi-frequency behavior.
Figure 5 depicts the dimensionless vibration frequency trend of side-by-side cylinders as a function of Ur. The dashed line represents vortex shedding frequency (Strouhal number St = 0.185) for a single cylinder (SC) in flow, as calculated by Jiang et al. [41]. According to the definition of reduced velocity Ur = u/fnD, Strouhal number St = fsD/u, and frequency ratio f* = fs/fn, we have Ur = u/fnD = f*u/fsD = f*/St. Prior to the “lock-in” region (on the IB), the dimensionless frequency curve closely follows the Strouhal line. After the “lock-in” region (on the DB), as L/D increases, the dimensionless frequency curve diverges from the Strouhal line. Given the symmetrical arrangement of the cylinders, the frequencies of cylinders C1 and C2 are identical; hence, only the variation in the VIV frequency of C1 with Ur is shown in Figure 5. When L/D ≤ 4, the “lock-in” region spans approximately Ur = 5 to 6. As the spacing ratio increases, for L/D ≥ 5, the range of the “lock-in” region expands to approximately Ur = 5 to 7.

3.1.2. Amplitude Response

This subsection will analyze the variations in amplitude, which correspond to the periods before (on the IB), during (on the LB), and after (on the DB) the “lock-in” region. Figure 6 illustrates the trends in the dimensionless cross-flow amplitude ( A y * ) and the dimensionless streamwise amplitude ( A x * ) of the cylinders with different L/D varying from 3 to 6, under 2DOF. For a clearer exposition of the cylinders’ vibration patterns, the amplitude variations for a single cylinder are also presented for comparison. Since the amplitudes of vibration for both cylinders are nearly identical when experiencing VIV at low Reynolds numbers, which is same as Chen et al. [24,25]; this study exclusively reports the amplitude characteristics of cylinder C1 for each spacing ratio.
From Figure 6, it is evident that the maximum amplitude responses for side-by-side cylinders with L/D from 3 to 6 occur at approximately Ur = 5. Specifically, the greatest dimensionless cross-flow amplitude ( A y * ) is approximately 0.56, while the largest dimensionless stream-wise amplitude ( A x * ) is at its peak when L/D = 3, reaching a value of 0.12. As L/D increases, the resonance range for larger spacing ratios visibly broadens, a fact also observable in Figure 5. At this point, the dimensionless amplitudes A y * and A x * gradually approach a single cylinder as Ur varies. Notably, amplitudes A y * for different spacing ratios are significantly less than those of a single cylinder at Ur = 8.
These observations reveal the complex interactions between cylinders under different spacing ratios L/D and reduced velocities Ur. As L/D increases and Ur deviates from the “lock-in” region, the amplitudes of cylinders gradually converge towards those of a single cylinder.

3.2. Lift and Drag Coefficients

As depicted in Figure 7a,b, at reduced velocities Ur = 2 and 3, the lift and drag coefficients of cylinder C1 are out of phase, and the phase differences between the individual force element components and the total lift and drag forces are less than π/2, indicating positive contributions to the total forces. During this stage, cylinder C1’s surface acceleration force, and the forces exerted by cylinder C2 on cylinder C1 are negligible. Cylinder C2 exhibits force element components that are oppositely symmetric to those of cylinder C1 in lift direction and identical in drag direction. Due to the surface interaction forces, the lift curve shifts, causing its equilibrium position to no longer be proximate to the x-axis. This offset is observable in subfigures (I) and (II) of Figure 7 at lower reduced velocities. Subfigure (III) of Figure 7 elucidates the magnitude of the surface interaction forces, with solid lines representing forces on cylinder C1 and dashed lines for cylinder C2. Surface vortices and friction forces seem to make a more substantial contribution here. Notably, in the drag direction, the surface interaction forces are of equal magnitude and phase, causing the curves to overlap. In lift direction, the surface vorticity lift components Clv12 and Clv21 of the two cylinders have equal amplitudes but opposite phases. As Ur increases, the amplitudes of all lift and drag coefficients rise.
As shown in Figure 7c, when Ur is approximately 4, nearing the end of the IB and the proximity of the “lock-in” region, within the IB, the phase differences between the force element components and the total lift and drag forces remain less than π/2, signifying positive contributions. Figure 7d,e illustrates the scenario at Ur = 5 and 6 when the system enters the “lock-in” region. The amplitudes of lift and drag force element components are markedly heightened, peaking at Ur = 5. At this point, in the lift direction, the volume vorticity lift coefficient Clv1 of cylinder C1 has phase differences greater than π/2 relative to the total lift coefficient Cl1, making negative contributions and inhibiting lift variations. Similarly, the surface acceleration force coefficient Cda1 in the drag direction for cylinder C1 exhibits a “large cyclic” behavior, with an overall phase difference greater than π/2 compared to the total drag coefficient Cd1, suppressing total drag variations. The acceleration force Cla21 in the drag direction produced by cylinder C2 on cylinder C1 noticeably increases. Given the stable periodicity of each force component, the total lift and drag coefficients of the cylinders manifest as steady non-sinusoidal periodicities. Cylinder C2 mirrors the phenomena observed in cylinder C1.
When Ur = 7 to 12, the vibration exits the “lock-in” region and enters the desynchronization zone. As observed in Figure 7f, when Ur = 7, akin to the situation at Ur = 4, the curves of the lift and drag force element components display a “multi-frequency” phenomenon. At this juncture, the phase difference between the surface acceleration force coefficient Cla1 in lift direction and the total lift coefficient Cl1 is greater than π/2, inhibiting total lift variations. Other force element components have phase differences less than π/2 relative to them, promoting the total lift variations. Moreover, within the desynchronization zone, the curves of lift and drag force element components gradually stabilize as Ur increases. As shown in Figure 7g, when Ur = 8, the “multi-frequency” phenomenon disappears.
Observations from Figure 8 reveal that when Ur is small, within the initial branch and as Ur increases, the amplitudes of lift and drag force element components show a growing trend. Particularly noticeable is the increase in the surface acceleration lift coefficient Cla21 exerted by cylinder C2 on cylinder C1. The phase differences between the force element component curves and the total lift and drag forces are minimal, with all force components contributing positively to the total lift and drag forces. Among these, the volume vorticity force and surface friction force of the cylinder play a dominant role in the variations in the total lift and drag forces, whereas the contributions of the force element components due to inter-cylinder interactions are relatively smaller, having a limited impact on the total lift and drag forces of the cylinders. When Ur = 5, entering the “lock-in” region, the amplitudes of lift and drag forces reach their maxima, with the amplitudes of all force element component curves notably higher than in the initial branch, a phenomenon consistent with the case for L/D = 3.
At Ur = 5 and 6, the phase difference between the vortex lift coefficient Clv1 of cylinder C1 and the total lift coefficient Cl1 exceeds π/2, making a negative contribution to the lift. The phase difference between the surface acceleration force coefficient Cda1 and the total drag coefficient Cd1 is also greater than π/2, suppressing the variations in total drag. The acceleration force coefficient produced by cylinder C2 on cylinder C1 in the drag direction shows a significant increase. Within this interval, the surface friction force and surface acceleration force coefficient curves exerted by cylinder C2 on cylinder C1 are out of phase by more than π/2 with the total lift coefficient curve of C1, inhibiting the variations in lift. At Ur = 7, the force element components of lift and drag coefficients exhibit a “multi-frequency” phenomenon, similar to that observed for a spacing ratio L/D = 3 (please note that no multi-frequency phenomena occurred when L/D = 3 and Ur = 4). At this point, the phase difference between the surface acceleration force coefficient Cla1 and the total lift coefficient Cl1 exceeds π/2, suppressing the variations in lift. The phase difference between the surface acceleration force coefficient Cda1 exerted on cylinder C1 and the total drag coefficient Cd1 is also greater than π/2, restraining the variations in drag. After Ur > 7, the curves of the force element components of lift and drag coefficients tend to stabilize.
As illustrated in Figure 9, the behavior of the force element components of the lift and drag coefficients for the cylinders with L/D = 5, as a function of reduced velocity, is similar to that observed for L/D = 3 and 4. However, at each reduced velocity, the amplitudes of the surface friction and surface acceleration force coefficients in the lift and drag directions are notably diminished compared to those at L/D = 3 and 4, indicating a significant reduction in the surface interaction forces. Furthermore, the total lift and drag coefficient curves progressively converge towards those of a single cylinder. An additional observation is that when the L/D = 5, the “multi-frequency” vibration phenomenon, which is presented at L/D = 3 and 4 at Ur = 7, is absent. This suggests that at a spacing ratio L/D = 5, the mutual influence between cylinders is considerably reduced, to the extent that it can be largely disregarded.
As depicted in Figure 10, when L/D is further increased to 6, the phase and amplitude fluctuation patterns of the force element component curves of lift and drag force as functions of Ur are similar to those observed at L/D = 5. The “multi-frequency” phenomena in the coefficient curves are attenuated. The amplitudes of surface acceleration forces and surface vorticity forces in lift and drag directions, generated by the inter-cylinder interactions, are further reduced. The lift, drag, and associated force element components of the cylinders converge even closer to those of a single cylinder.

3.3. Motion Trajectory

Tu et al. [42] classified the vibration trajectories into seven patterns, including “figure-eight”, “teardrop”, “elliptical”, “double-teardrop”, “double-figure-eight”, “double-elliptical”, and irregular shapes. In the present study, additional trajectory patterns emerge beyond those identified by Tu et al. [42], including “bounded elliptical”, “bounded teardrop”, and “bounded figure-eight”. As the motion trajectories of the two cylinders are completely symmetrical in this study, only the trajectory of cylinder C1 is presented in this section.

3.3.1. Trajectory Modes

As shown in Figure 11a,b, when Ur = 2 and 3, the cylinder C1 with L/D = 3 is in the initial branch of its vibration responses. During this phase, the cylinder motion exhibits a small amplitude, and the trajectory is relatively stable, taking on the appearance of an “asymmetric figure-eight”. This trajectory is characterized by being narrow on one side and wider on another side. According to Tu et al. [42], this is attributed to the asymmetric wake distribution around the cylinders at this stage. This study further analyzes the cause of this phenomenon from the perspective of force elements, as detailed below in Section 3.3.2. When Ur = 4, approaching the end of the initial branch, the trajectory transforms into an irregular shape, as seen in Figure 11c. This is characterized by a series of nested “elliptical” curves, which can be considered a form of “bounded elliptical”. When Ur = 5 and 6, the vibration enters the “lock-in” region, where the trajectories are stable elliptical curves throughout this phase, as depicted in Figure 11d,e. The cylinder vibration amplitude increases but remains stable. At Ur = 7, exiting “lock-in” region and entering the desynchronization zone, Figure 11f reveals a trajectory consisting of nested “elliptical” curves, also interpretable as a “bounded elliptical”. The cause of this phenomenon will be analyzed later in Section 3.3.2. As shown in Figure 11g–k, once entering the desynchronization region (for Ur > 7), the trajectories consistently manifest as “elliptical” curves.
The motion trajectories for cylinders with L/D = 4 are depicted in Figure 12. At a reduced velocity Ur = 2, the motion traces an “asymmetric figure-eight” pattern that is narrow at the top and wider at the bottom. When Ur = 3, this “figure-eight” pattern transitions to one that is broader at the top and narrower at the bottom. At Ur = 4, the trajectory shows a “teardrop” shape, which is relatively stable. Within the “lock-in” region, the trajectories consistently exhibit a “teardrop” shape that is broader at the top and narrower at the bottom, remaining relatively stable. At Ur = 7, once exiting the “lock-in” region and entering the desynchronization zone, the trajectory becomes a regular bounded motion curve, which can be described as a “bounded teardrop”. The causes behind this phenomenon will be discussed later in Section 3.3.2. For Ur = 8 and 9, the trajectory evolves into a “double-teardrop” shape. When Ur > 9, the trajectory once again reverts to a “teardrop” form.
Figure 13 displays the displacement trajectories of cylinders at various reduced velocities for a spacing ratio of L/D = 5. When Ur ranges from 2 to 4, the cylinders are in the initial branch, tracing “figure-eight” paths. As Ur increases, the “figure-eight” curves gradually become narrower. When Ur spans from 5 to 7 and the cylinders are within the “lock-in” region, the trajectories take on a “teardrop” shape that is broader at the top and pointed at the bottom. As Ur continues to increase and the cylinders exit the “lock-in” region into the desynchronization zone, the motion trajectories evolve, with the curves increasingly resembling a “figure-eight” pattern as Ur grows. At Ur = 8, it can be observed that over the course of the simulation, the trajectory transitions from a “teardrop” to a “figure-eight” shape. For Ur ≥ 9, the cylinder motion trajectories shift from dual “figure-eight” to “bounded figure-eight”. This transformation might be attributed to the presence of unstable frequencies in the streamwise vibration, a detailed explanation for which will be provided in Section 3.3.2.
Figure 14 presents the motion trajectories of cylinders at various Ur for a spacing ratio L/D = 6. Under this spacing condition, the trajectories resemble those of a single cylinder, predominantly a “figure-eight” pattern. Within the initial branch, relatively regular “figure-eight” shapes are observed, with the “figure-eight” becoming increasingly narrow as Ur increases, with the ends approaching near-coincidence. When in the “lock-in” region, the curves adopt “figure-eight” forms that are nearly “teardrop”-like, owing to the stronger transverse vibrations and the amplified effect of surface interaction forces, as observable in Figure 10d,e.
Once exiting the “lock-in” region at Ur = 8 and entering desynchronization zone, the trajectories initially manifest as disordered “figure-eight” curves. However, as reduced velocity continues to rise, the trajectories gradually settle into a more distinct “figure-eight” pattern. At Ur = 12, the cylinder motion traces a double “figure-eight” trajectory. According to Tu et al. [42], this is due to the second-order frequency in the power spectral density (PSD) of x/D and Cd being twice that of the primary frequency in y/D and Cl, generating an asymmetrical double “figure-eight” in the trajectory.
To have a clearer comparison of amplitude, the cylindrical motion trajectories at different Ur and L/D = 3, 4, 5, 6 are shown in Figure 15 under the same axis ranges for all subfigures in Figure 11, Figure 12, Figure 13 and Figure 14.

3.3.2. Trajectory Patterns Analyses

In this section, further analysis will be conducted on the motion trajectories of interest discussed in the previous section.
For gaining deeper insight into the genesis of the asymmetric “figure-eight” pattern, time-history plots of the surface acceleration force coefficient Cda1 and streamwise displacement relative to x/D have been delineated for a cylinder with a spacing ratio L/D = 3 at reduced velocities Ur = 2 and 3. These plots, presented in Figure 16, reveal a distinct “modulation period” within the curves, serving as the principal mechanism behind the formation of the “asymmetric figure-eight” trajectory. It is noteworthy that the side characterized by elevated peak values manifests a constricted “figure-eight” configuration.
At L/D = 3 and Ur = 7, the time-history curves of the streamwise and cross-flow positions of the cylinder is depicted in Figure 17. Here, the phenomenon of “flutter” is evident in each direction, characterized by periodic simultaneous amplification and diminution of displacements in the x and y directions, respectively. This results in the formation of superimposed “bounded elliptical” trajectories in the motion paths. Similarly, at L/D = 4 and Ur = 7, the time-history curves for the streamwise and cross-flow positions of the cylinder are illustrated in Figure 18. In this case, the displacements in the x and y directions exhibit a “multi-periodic” behavior, leading to “bounded teardrop-shaped” trajectories. It is noteworthy that these “bounded” trajectories predominantly occur at the boundaries of branches for spacing ratios L/D ≤ 4, particularly when transitioning into or out of the “lock-in” region; this is consistent with the conclusion reached in the previous section.
To understand the underlying mechanism of the specific trajectory “bounded figure-eight” at L/D = 5, Ur = 10 in Figure 13i, as depicted in Figure 19a, the Fast Fourier Transform (FFT) analysis of drag coefficient Cd unveils a multitude of sub-harmonics within the PSD, contributing to a subtle transverse displacement of “figure-eight” trajectory. In conjunction, the time-series plot of x/D and Cda1 are shown in Figure 19b, the component of the acceleration force coefficient associated with the motion. It is evident that Cda1 exhibits a “modulation period” correlating with the development of a “figure-eight” trajectory, which is distinctly constricted at the lower portion and expanded at the upper section. This observation aligns seamlessly with the conclusions drawn earlier in the study.
These observations provide profound insights into the complex dynamics of VIV for cylinders in side-by-side configurations under varying spacing ratios and reduced velocities. The identification of distinct patterns and the underlying mechanisms driving the transition between different vibration modes is essential for the development of predictive models and mitigation strategies in engineering applications involving VIV phenomena.

3.4. Wake-Stream Pattern

In the flow past side-by-side dual circular cylinders, the interaction between them varies with spacing ratio L/D, leading to several distinct wake patterns [43]. Investigations by Singh and Mittal [44] and Prasanth and Mittal [45] revealed that the vortex shedding modes corresponding to IB and LB are denoted as 2S and C(2S), respectively. The C(2S) mode shares similarities with the 2S mode, but it is characterized by the coalescence of vortices in the wake.
For L/D = 3, the vortex shedding patterns around cylinders are illustrated in Figure 20. When Ur ≤ 4, within the primary branch, the vortex shed from both cylinders exhibits a uniform single-type pattern, with small amplitudes and minimal influence on the generation of vortices. Consequently, the 2S shedding mode, akin to that observed in the flow around an isolated cylinder, where pairs of vortices of opposite rotation directions are shed alternately from the upper and lower sides of the cylinder during each oscillation cycle, is prevalent. As Ur increases to five, the VIV enters the lower branch’s “lock-in” region. With enhanced transverse vibrations, the shedding pattern transforms; pairs of counter-rotating vortices form on either side of the cylinders’ wake. During each oscillation cycle, two rows of vortices are shed, one stronger than the other, resulting in a symmetrically opposed double-row C(2S) shedding mode. The vortices dissipate energy more rapidly in this configuration. At Ur = 6, the wake transitions from the initially antiphase synchronous 2S mode towards the symmetrically opposed C(2S) mode. Within this stage, the shedding length shortens, vortex shedding accelerates, and the shedding frequency significantly increases. The cylinders exhibit larger transverse amplitudes, with vigorous vortex shedding and a pronounced lateral displacement of the vortices. The wake widens, and the vortices detach further from the cylinder surface. When Ur ≥ 7, the vortex shedding pattern reverts to 2S, similar to that at Ur = 3. It can also be observed from Figure 6 that the amplitudes of side-by-side cylinders are suppressed in the primary branch Ur = 2 to 4 and non-synchronous range Ur = 7 to 12, dominated by 2S mode throughout.
When L/D = 4, the vortex shedding patterns around cylinders are depicted in Figure 21. Both within the primary branch Ur ≤ 4 and in the non-synchronous regime Ur ≥ 7, the wake exhibits a pair of antiphase synchronized 2S mode vortices. Once entering the “lock-in” zone, at Ur = 5, the vortices shed follows a counter-rotating double-row C(2S) mode. At Ur = 6, the near-wake region demonstrates an antiphase synchronized 2S mode, while in the far-wake region, the 2S mode vortices merge and weaken, gradually transitioning toward C(2S) mode. For an increased spacing ratio of L/D = 5, the vortex shedding behavior of cylinders across different reduced velocities adheres to the same pattern as observed for L/D = 4; thus, these observations will not be reiterated here.
As L/D between the cylinders further increases to six, the vortex shedding patterns around cylinders are illustrated in Figure 22. When the reduced velocity Ur ≤ 4, the vortices shed from both cylinders display an anti-phase synchronized shedding of two sets in 2S mode. As Ur increases and the system’s vibration frequency reaches the “lock-in” region, the cylinder amplitudes are substantial, and the vortex shedding pattern assumes a double-row C(2S) mode. At a reduced velocity Ur = 6, despite the increase in Ur, the significant spacing between cylinders causes a reduction of inter-cylinder interactions, and as a result, no vortex merging is observable in the wake. When Ur ≥ 7 and the system enters the non-synchronous zone, the vibration response amplitudes further decrease, and the vortex shedding pattern returns to the 2S mode, similar to that in the primary branch.
It is noteworthy that, in the present study, all spacing conditions exhibited anti-phase synchronized vortex streets, even when the spacing ratio was further increased to L/D = 6. No in-phase synchronized vortex streets, as mentioned by Williamson [17] and Sumner [43], were observed in this investigation. This finding highlights the dominance of anti-phase synchronization in the vortex shedding patterns for the examined range of spacing ratios, suggesting that other factors may be influential in achieving in-phase synchronization reported in previous studies.

4. Conclusions

This study presents a force mechanism, from the perspective of the multi-body force element formulations, of the vortex-induced vibration (VIV) phenomenon of the side-by-side dual cylinders at the Reynolds number Re = 100. The main findings are summarized as follows:
(1)
The VIV phenomenon for side-by-side dual cylinders exhibits three vibration regions with different spacing ratios, similar to those observed for a single cylinder, namely the initial branch (IB), the lower branch (LB), and the desynchronization branch (DB). When the spacing ratio L/D is greater than or equal to five, the range of reduced velocities Ur corresponding to the lower branch is wider and more closely resembles that of a single cylinder. For side-by-side cylinders at different spacing ratios, the maximum displacements in the streamwise and cross-flow directions occur at a reduced velocity Ur = 5. Cross-flow displacements are comparable to those of a single cylinder. In the streamwise direction, the amplitude decreases gradually with increasing spacing ratio, approaching the behavior of a single cylinder. Notably, there is no sudden increase in streamwise amplitude at a Ur = 8, unlike the case for a single cylinder.
(2)
During the VIV phenomenon, the lift and drag forces acting on side-by-side cylinders are primarily due to the volume vorticity forces associated with the cylinders themselves, combined with surface vortices (including friction) forces. When entering the “lock-in” zone, as cylinder vibrations intensify increase, the acceleration forces on the cylinder surface also increase. The volume vorticity lift force phase opposes the total lift force, thus inhibiting (negatively contributing) to the lift force. As the reduced velocity increases and the cylinder amplitude reaches its maximum, the surface acceleration forces begin to suppress changes in lift force, decreasing with further increases in reduced velocity. At lower reduced velocities, side-by-side cylinders are mainly influenced by the surface vorticity forces from the surface of the other cylinder. As reduced velocity increases, the influence of inter-cylinder surface acceleration forces grows, leading to increased amplitudes. The effect of surface vorticity forces in lift direction stabilizes over time. As the spacing ratio between cylinders increases, the influence between cylinders diminishes, and the force variation trends approach those of a single cylinder.
(3)
Once in the “lock-in” zone and experiencing significant vibrations, the wake vortices of side-by-side cylinders transition from a 2S mode to a C(2S) mode. The smaller the spacing ratio, the more pronounced this transformation becomes. At the onset and end of the wake mode transition, the displacement trajectories of the cylinders exhibit irregular “bounded” motions. During this period, the lift and drag coefficients display a “multi-frequency” behavior. Simultaneously, the vibration trajectories of side-by-side cylinders form “bounded elliptical”, and “bounded teardrop” shapes.

Author Contributions

Conceptualization, M.S. and H.X.; Data curation, S.G.; Investigation, M.S. and S.G.; Methodology, C.-C.C.; Resources, M.S.; Software, S.G. and W.T.; Supervision, M.S., H.X., J.L. and C.-C.C.; Writing—original draft, S.G. and W.T.; Writing—review and editing, M.S., S.G., H.X., J.L. and C.-C.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Guangxi Science and Technology Base and Special Talents through Grant No. guike AD21220101, and the National Council of Science and Technology (Taiwan) through Grant No. NSTC 113-2221-E-002-193-MY3.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request.

Conflicts of Interest

The authors declare no conflicts of interest.

Nomenclature

ReReynolds number
UUniform incoming fluid velocity
DDiameter of the cylinder
ρDensity of the fluid
PNon-dimensional pressure
P*Pressure
UrReduced velocity
StStrouhal number
fsFrequency of the vortex shedding
fnNatural frequency
f *Frequency ratio
kSpring constant
cDamping coefficient
ζDamping ratio
m*Mass ratio
FDDrag force
FLLift force
CdDrag force coefficient
ClLift force coefficient
C ¯ d Time-averaged drag force coefficient
C ¯ l Time-averaged lift force coefficient
CdaSurface acceleration drag force coefficient
CdvVolume vorticity drag force coefficient
CdsSurface vorticity drag force coefficient
ClaSurface acceleration lift force coefficient
ClvVolume vorticity lift force coefficient
ClsSurface vorticity lift force coefficient
A x * In-line amplitude
A y * Cross-flow amplitude
h*Nondimensional mesh spacing
hfirstNear-wall first grid height
LCenter-to-center spacing of cylinders
L/DSpacing ratio
CtotalTotal hydrodynamic force coefficient
FtotalTotal hydrodynamic force

References

  1. Williamson, C.H.K.; Govardhan, R. Vortex-induced vibrations. Annu. Rev. Fluid Mech. 2004, 36, 413–455. [Google Scholar] [CrossRef]
  2. Zhao, M. Flow induced vibration of two rigidly coupled circular cylinders in tandem and side-by-side arrangements at a low Reynolds number of 150. Phys. Fluids 2013, 25, 123601. [Google Scholar] [CrossRef]
  3. Zdravkovich, M.M. Review of flow interference between two circular cylinders in various arrangements. ASME J. Fluids Eng. 1977, 99, 618–633. [Google Scholar] [CrossRef]
  4. Zdravkovich, M.M. Flow induced oscillations of two interfering circular cylinders. J. Sound Vib. 1985, 101, 511–521. [Google Scholar] [CrossRef]
  5. Zdravkovich, M.M. The effects of interference between circular cylinders in cross flow. J. Fluids Struct. 1987, 1, 239–261. [Google Scholar] [CrossRef]
  6. Williamson, C.H.K. Vortex dynamics in the cylinder wake. Annu. Rev. Fluid Mech. 1996, 28, 477–539. [Google Scholar] [CrossRef]
  7. Alam, M.M.; Zhou, Y. Flow around two side-by-side closely spaced circular cylinders. J. Fluids Struct. 2007, 23, 799–805. [Google Scholar] [CrossRef]
  8. Zhao, M.; Cheng, L. Two-dimensional numerical study of vortex shedding regimes of oscillatory flow past two circular cylinders in side-by-side and tandem arrangements at low Reynolds numbers. J. Fluid Mech. 2014, 751, 1–37. [Google Scholar] [CrossRef]
  9. Bai, X.D.; Zhang, W.; Guo, A.X.; Wang, Y. The flip-flopping wake pattern behind two side-by-side circular cylinders: A global stability analysis. Phys. Fluids 2016, 28, 044102. [Google Scholar] [CrossRef]
  10. Feng, C.C. The Measurement of Vortex Induced Effects in Flow Past Stationary and Oscillating Circular and D-section Cylinders. Ph.D. Thesis, University of British Columbia, Vancouver, BC, Canada, 1968. [Google Scholar]
  11. Bearman, P.W. Vortex shedding from oscillating bluff bodies. Annu. Rev. Fluid Mech. 2012, 16, 195–222. [Google Scholar] [CrossRef]
  12. Williamson, C.H.K.; Govardhan, R. A brief review of recent results in vortex-induced vibrations. J. Wind Eng. Ind. Aerod. 2008, 96, 713–735. [Google Scholar] [CrossRef]
  13. Sarpkaya, T. A critical review of the intrinsic nature of vortex-induced vibrations. J. Fluid Struct. 2004, 19, 389–447. [Google Scholar] [CrossRef]
  14. Gabbai, R.D.; Benaroya, H. An overview of modeling and experiments of vortex induced vibration of circular cylinders. J. Sound Vib. 2005, 282, 575–616. [Google Scholar] [CrossRef]
  15. Zhao, M.; Cheng, L. Numerical simulation of two-degree-of-freedom vortex-induced vibration of a circular cylinder close to a plane boundary. J. Fluid Struct. 2011, 27, 1097–1110. [Google Scholar] [CrossRef]
  16. Wu, X.; Ge, F.; Hong, Y. A review of recent studies on vortex-induced vibrations of long slender cylinders. J. Fluid Struct. 2012, 28, 292–308. [Google Scholar] [CrossRef]
  17. Williamson, C.H.K. Evolution of a single wake behind a pair of bluff bodies. J. Fluid Mech. 1985, 159, 1–18. [Google Scholar] [CrossRef]
  18. Kim, H.J.; Durbin, P.A. Investigation of the flow between a pair of cylinders in the flopping regime. J. Fluid Mech. 1988, 196, 431–448. [Google Scholar] [CrossRef]
  19. Alam, M.M.; Moriya, M.; Sakamoto, H. Aerodynamic characteristics of two side-by-side circular cylinders and application of wavelet analysis on the switching phenomenon. J. Fluids Struct. 2003, 18, 325–346. [Google Scholar] [CrossRef]
  20. Cui, Z.; Zhao, M.; Teng, B. Vortex-induced vibration of two elastically coupled cylinders in side-by-side arrangement. J. Fluids Struct. 2014, 44, 270–291. [Google Scholar] [CrossRef]
  21. Chen, W.; Ji, C.; Wang, R.; Xu, D.; Campbell, J. Flow-induced vibrations of two side-by-side circular cylinders: Asymmetric vibration, symmetry hysteresis and near-wake patterns. Ocean Eng. 2015, 110, 244–257. [Google Scholar] [CrossRef]
  22. Rahmanian, M.; Cheng, L.; Zhao, M.; Zhou, T. Lock-in study of two side-by-side cylinders of different diameters in close proximity in steady flow. J. Fluids Struct. 2014, 49, 386–411. [Google Scholar] [CrossRef]
  23. Xu, W.; Wu, H.; Sha, M.; Wang, E. Numerical study on the flow-induced vibrations of two elastically mounted side-by-side cylinders at subcritical Reynolds numbers. Appl. Ocean Res. 2022, 124, 103191. [Google Scholar] [CrossRef]
  24. Chen, W.; Ji, C.; Xu, D. Vortex-induced vibrations of two side-by-side circular cylinders with two degrees of freedom in laminar cross-flow. Comput. Fluids 2019, 193, 104288. [Google Scholar] [CrossRef]
  25. Chen, W.; Ji, C.; Xu, D.; An, H.; Zhang, Z. Flow-induced vibrations of two side-by-side circular cylinders at low Reynolds numbers. Phys. Fluids 2020, 32, 023601. [Google Scholar] [CrossRef]
  26. Chang, C.C.; Yang, S.H.; Chu, C.C. A many-body force decomposition with applications to flow about bluff bodies. J. Fluid Mech. 2008, 600, 95–104. [Google Scholar] [CrossRef]
  27. Chang, C.C. Potential flow and forces for incompressible viscous flow. Proc. R. Soc. A-Math. Phy. 1992, 437, 517–525. [Google Scholar]
  28. Lee, J.J.; Hsieh, C.T.; Chang, C.C.; Chu, C.C. Vorticity forces on an impulsively started finite plate. J. Fluid Mech. 2012, 694, 464–492. [Google Scholar] [CrossRef]
  29. Lin, Y.S.; Tzeng, Y.T.; Hsieh, C.T.; Chang, C.C.; Chu, C.C. A mechanism of thrust enhancement on a heaving plate due to flexibility at moderately low Reynolds numbers. J. Fluid Struct. 2018, 76, 573–591. [Google Scholar] [CrossRef]
  30. Menon, K.; Mittal, R. On the initiation and sustenance of flow-induced vibration of cylinders: Insights from force partitioning. J. Fluid Mech. 2020, 907, A37. [Google Scholar] [CrossRef]
  31. Menon, K.; Kumar, S.; Mittal, R. Contribution of spanwise and cross-span vortices to the lift generation of low-aspect-ratio wings: Insights from force partitioning. Phys. Rev. Fluids 2022, 7, 114102. [Google Scholar] [CrossRef]
  32. Yin, G.; Janocha, M.J.; Ong, M.C. Estimation of hydrodynamic forces on cylinders undergoing flow-induced vibrations based on modal analysis. J. Offshore Mech. Arct. 2022, 144, 060904. [Google Scholar] [CrossRef]
  33. Luo, X.G.; Gao, A.K.; Lu, X.Y. Enhanced performance of a self-propelled flexible plate by a uniform shear flow and mechanism insight. Phys. Fluids 2023, 35, 021903. [Google Scholar] [CrossRef]
  34. Xu, H.L.; Lei, J.C.; Chang, C.C.; Wang, C.Y. Lifting Stokes’ paradox by accelerating flow past a circular cylinder and extension of the analysis to the sphere. Phys. Fluids 2023, 35, 033608. [Google Scholar] [CrossRef]
  35. Moriche, M.; Sedky, G.; Jones, A.R.; Flores, O.; García-Villalba, M. Characterization of aerodynamic forces on wings in plunge maneuvers. AIAA J. 2021, 59, 751–762. [Google Scholar] [CrossRef]
  36. Carini, M.; Auteri, F.; Giannetti, F. Secondary instabilities of the in-phase synchronized wakes past two circular cylinders in side-by-side arrangement. J. Fluids Struct. 2015, 53, 70–83. [Google Scholar] [CrossRef]
  37. Han, X.; Lin, W.; Wang, D.; Qiu, A.; Feng, Z.; Tang, Y.; Wu, J. Numerical simulation of super upper branch of a cylindrical structure with a low mass ratio. Ocean Eng. 2018, 168, 108–120. [Google Scholar] [CrossRef]
  38. Vu, H.C.; Ahn, J.; Hwang, J.H. Numerical simulation of flow past two circular cylinders in tandem and side-by-side arrangement at low Reynolds numbers. KSCE J. Civ. Eng. 2016, 20, 1594–1604. [Google Scholar] [CrossRef]
  39. De Langre, E. Frequency lock-in is caused by coupled-mode flutter. J. Fluid Struct. 2006, 22, 783–791. [Google Scholar] [CrossRef]
  40. Mittal, S.; Kumar, V. Finite element study of vortex-induced cross-flow and in-line oscillations of a circular cylinder at low Reynolds numbers. Int. J. Numer. Methods Fluid. 1999, 31, 1087–1120. [Google Scholar] [CrossRef]
  41. Jiang, H.; Cheng, L.; Draper, S.; An, H.; Tong, F. Three-dimensional direct numerical simulation of wake transitions of a circular cylinder. J. Fluid Mech. 2016, 801, 353–391. [Google Scholar] [CrossRef]
  42. Tu, J.; Tan, X.; Deng, X.; Han, Z.; Zhang, M.; Li, Z.; Xu, J.; Zhang, P. Dynamic responses and flow-induced vibration mechanism of three tandem circular cylinders in planar shear flow. Ocean Eng. 2020, 199, 107022. [Google Scholar] [CrossRef]
  43. Sumner, D. Two circular cylinders in cross-flows: A review. J. Fluids Struct. 2010, 26, 849–899. [Google Scholar] [CrossRef]
  44. Singh, S.P.; Mittal, S. Vortex- Induced Oscillations at Low Reynolds Numbers: Hysteresis and Vortex-Shedding Modes. J. Fluid Struct. 2005, 20, 1085–1104. [Google Scholar] [CrossRef]
  45. Prasanth, T.K.; Mittal, S. Vortex- Induced Vibrations of a Circular Cylinder at Low Reynolds Numbers. J. Fluid Mech. 2007, 594, 463–491. [Google Scholar] [CrossRef]
Figure 1. Schematic illustration of the flow field physical model.
Figure 1. Schematic illustration of the flow field physical model.
Jmse 12 01529 g001
Figure 2. The grid diagram of the flow field.
Figure 2. The grid diagram of the flow field.
Jmse 12 01529 g002
Figure 3. Comparison of the average lift and drag coefficients under different Ur: (a) average lift coefficient; (b) average drag coefficient [24].
Figure 3. Comparison of the average lift and drag coefficients under different Ur: (a) average lift coefficient; (b) average drag coefficient [24].
Jmse 12 01529 g003
Figure 4. Comparison of lift and drag coefficients obtained by pressure method (PM) and Force Element Analysis (FEA) at Re = 100 and Ur = 5.
Figure 4. Comparison of lift and drag coefficients obtained by pressure method (PM) and Force Element Analysis (FEA) at Re = 100 and Ur = 5.
Jmse 12 01529 g004
Figure 5. The variation in the vibration frequency ratio of cylinder C1 under different Ur, here, the dashed line represents the shedding frequency (St = 0.185) for the cylinder wake [41].
Figure 5. The variation in the vibration frequency ratio of cylinder C1 under different Ur, here, the dashed line represents the shedding frequency (St = 0.185) for the cylinder wake [41].
Jmse 12 01529 g005
Figure 6. The displacement of cylinder C1 at different L/D and different Ur: (a) in-line flow direction; (b) cross-flow direction.
Figure 6. The displacement of cylinder C1 at different L/D and different Ur: (a) in-line flow direction; (b) cross-flow direction.
Jmse 12 01529 g006
Figure 7. The variations in the lift and drag coefficients together with their components (Cda, Cdv, Cds, Cla, Clv, Cls) with time under different Ur at L/D = 3.0: (I) forces on cylinder C1; (II) forces on cylinder C2; (III) the interaction forces between cylinder C1 and C2.
Figure 7. The variations in the lift and drag coefficients together with their components (Cda, Cdv, Cds, Cla, Clv, Cls) with time under different Ur at L/D = 3.0: (I) forces on cylinder C1; (II) forces on cylinder C2; (III) the interaction forces between cylinder C1 and C2.
Jmse 12 01529 g007aJmse 12 01529 g007b
Figure 8. The variations in the lift and drag coefficients together with their components (Cda, Cdv, Cds, Cla, Clv, Cls) with time under different Ur at L/D = 4: (I) forces on cylinder C1; (II) forces on cylinder C2; (III) the interaction forces between cylinder C1 and C2.
Figure 8. The variations in the lift and drag coefficients together with their components (Cda, Cdv, Cds, Cla, Clv, Cls) with time under different Ur at L/D = 4: (I) forces on cylinder C1; (II) forces on cylinder C2; (III) the interaction forces between cylinder C1 and C2.
Jmse 12 01529 g008
Figure 9. The variations in the lift and drag coefficients together with their components (Cda, Cdv, Cds, Cla, Clv, Cls) with time under different Ur at L/D = 5: (I) forces on cylinder C1; (II) forces on cylinder C2; (III) the interaction forces between cylinder C1 and C2.
Figure 9. The variations in the lift and drag coefficients together with their components (Cda, Cdv, Cds, Cla, Clv, Cls) with time under different Ur at L/D = 5: (I) forces on cylinder C1; (II) forces on cylinder C2; (III) the interaction forces between cylinder C1 and C2.
Jmse 12 01529 g009aJmse 12 01529 g009b
Figure 10. The variations in the lift and drag coefficients together with their components (Cda, Cdv, Cds, Cla, Clv, Cls) with time under different Ur at L/D = 6: (I) forces on cylinder C1; (II) forces on cylinder C2; (III) the interaction forces between cylinder C1 and C2.
Figure 10. The variations in the lift and drag coefficients together with their components (Cda, Cdv, Cds, Cla, Clv, Cls) with time under different Ur at L/D = 6: (I) forces on cylinder C1; (II) forces on cylinder C2; (III) the interaction forces between cylinder C1 and C2.
Jmse 12 01529 g010aJmse 12 01529 g010b
Figure 11. Cylindrical motion trajectories under different Ur at L/D = 3.
Figure 11. Cylindrical motion trajectories under different Ur at L/D = 3.
Jmse 12 01529 g011
Figure 12. Cylindrical motion trajectories under different Ur at L/D = 4.
Figure 12. Cylindrical motion trajectories under different Ur at L/D = 4.
Jmse 12 01529 g012
Figure 13. Cylindrical motion trajectories under different Ur at L/D = 5.
Figure 13. Cylindrical motion trajectories under different Ur at L/D = 5.
Jmse 12 01529 g013
Figure 14. Cylindrical motion trajectories under different Ur at L/D = 6.
Figure 14. Cylindrical motion trajectories under different Ur at L/D = 6.
Jmse 12 01529 g014
Figure 15. Cylindrical motion trajectories under the same axis ranges at different Ur and L/D = 3, 4, 5, 6.
Figure 15. Cylindrical motion trajectories under the same axis ranges at different Ur and L/D = 3, 4, 5, 6.
Jmse 12 01529 g015
Figure 16. The variations in displacement and surface acceleration at L/D = 3: (a) Ur = 2; (b) Ur = 3.
Figure 16. The variations in displacement and surface acceleration at L/D = 3: (a) Ur = 2; (b) Ur = 3.
Jmse 12 01529 g016
Figure 17. The variations in the inline and transverse positions of the cylinder in the flow at L/D = 3, Ur = 7.
Figure 17. The variations in the inline and transverse positions of the cylinder in the flow at L/D = 3, Ur = 7.
Jmse 12 01529 g017
Figure 18. The variations in the inline and transverse positions of the cylinder in the flow at L/D = 4, Ur = 7.
Figure 18. The variations in the inline and transverse positions of the cylinder in the flow at L/D = 4, Ur = 7.
Jmse 12 01529 g018
Figure 19. The variations in (a) Power Spectral Density (PSD) of Cd1 with frequency ratio; (b) x-amplitude and surface acceleration drag force Cda1 with time; here L/D = 5, Ur = 10.
Figure 19. The variations in (a) Power Spectral Density (PSD) of Cd1 with frequency ratio; (b) x-amplitude and surface acceleration drag force Cda1 with time; here L/D = 5, Ur = 10.
Jmse 12 01529 g019
Figure 20. Vorticity contours of wake vorticity distributions under different Ur at L/D = 3: (a) Ur = 3; (b) Ur = 5; (c) Ur = 6; (d) Ur = 8.
Figure 20. Vorticity contours of wake vorticity distributions under different Ur at L/D = 3: (a) Ur = 3; (b) Ur = 5; (c) Ur = 6; (d) Ur = 8.
Jmse 12 01529 g020
Figure 21. Vorticity contours of wake vorticity distributions under different Ur at L/D = 4: (a) Ur = 3; (b) Ur = 5; (c) Ur = 6; (d) Ur = 8.
Figure 21. Vorticity contours of wake vorticity distributions under different Ur at L/D = 4: (a) Ur = 3; (b) Ur = 5; (c) Ur = 6; (d) Ur = 8.
Jmse 12 01529 g021
Figure 22. Vorticity contours of wake vorticity distributions under different Ur at L/D = 6: (a) Ur = 3; (b) Ur = 5; (c) Ur = 6; (d) Ur = 8.
Figure 22. Vorticity contours of wake vorticity distributions under different Ur at L/D = 6: (a) Ur = 3; (b) Ur = 5; (c) Ur = 6; (d) Ur = 8.
Jmse 12 01529 g022
Table 1. Grid and time step independence test: the results of side-by-side dual circular cylinders at Re = 100, m* = 2, L/D = 3, Ur = 6, ζ = 0.
Table 1. Grid and time step independence test: the results of side-by-side dual circular cylinders at Re = 100, m* = 2, L/D = 3, Ur = 6, ζ = 0.
MeshCell Numberh*tU/D C ¯ d C ¯ l A x * A y *
M156,4900.010.00251.9010.1910.0690.514
M280,5500.0050.00251.872 (1.5%)0.186 (2.7%)0.066 (4.5%)0.511 (0.6%)
M3146,4000.00250.00251.863 (0.5%)0.184 (1.1%)0.065 (1.5%)0.503 (1.6%)
M280,5500.0050.001251.860 (0.6%)0.187 (0.5%)0.066 (0%)0.508 (0.6%)
Chen et al. [24] ---1.8280.1870.0600.536
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Song, M.; Guo, S.; Xu, H.; Tao, W.; Lei, J.; Chang, C.-C. Force Element Analysis in Vortex-Induced Vibrations of Side-by-Side Dual Cylinders: A Numerical Study. J. Mar. Sci. Eng. 2024, 12, 1529. https://doi.org/10.3390/jmse12091529

AMA Style

Song M, Guo S, Xu H, Tao W, Lei J, Chang C-C. Force Element Analysis in Vortex-Induced Vibrations of Side-by-Side Dual Cylinders: A Numerical Study. Journal of Marine Science and Engineering. 2024; 12(9):1529. https://doi.org/10.3390/jmse12091529

Chicago/Turabian Style

Song, Mengtian, Suxiang Guo, Hailong Xu, Weijian Tao, Jiechao Lei, and Chien-Cheng Chang. 2024. "Force Element Analysis in Vortex-Induced Vibrations of Side-by-Side Dual Cylinders: A Numerical Study" Journal of Marine Science and Engineering 12, no. 9: 1529. https://doi.org/10.3390/jmse12091529

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop