Next Article in Journal
Emerging Contaminants: A Rising Threat to Urban Water and a Barrier to Achieving SDG-Aligned Planetary Protection
Previous Article in Journal
Seasonal Circulation Characteristics of Oceanic System in the Beibu Gulf Based on Observations and Numerical Simulations
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Hydrodynamic Loadings and Vorticity Distribution on a Circular Cylinder in a Narrow Channel

1
Department of Civil Engineering, National Central University, 300 Jhong-Da Road, Jhong-Li, Taoyuan 32001, Taiwan
2
Institute of Hydrological and Oceanic Science, National Central University, 300 Jhong-Da Road, Jhong-Li, Taoyuan 32001, Taiwan
*
Author to whom correspondence should be addressed.
Water 2025, 17(16), 2366; https://doi.org/10.3390/w17162366
Submission received: 19 June 2025 / Revised: 28 July 2025 / Accepted: 6 August 2025 / Published: 9 August 2025
(This article belongs to the Section Hydraulics and Hydrodynamics)

Abstract

A large eddy simulation (LES) model, integrated with the volume of fluid (VOF) method, was employed to investigate hydrodynamic forces and vorticity distribution around a circular cylinder in a narrow channel. The simulated surface pressure and drag coefficient closely matched the experimental results from flume testing. The ratio of cylinder diameter to channel width is defined as the blockage ratio (Br). The effects of blockage on hydrodynamic loadings and vortex structures around the cylinder were examined through a series of numerical simulations. The results reveal that blockage ratios exceeding 20% significantly alter key flow characteristics, including the upstream and circumferential pressure coefficients, drag coefficient, lateral force coefficient, and Strouhal number. Higher blockage ratios enhance near-wall vortex formation and intensify shear layers. The vertical (Ωy), streamwise (Ωx), and spanwise (Ωz) vorticity components all increase with Br, leading to stronger and more spatially extensive vortex structures near the bed, particularly in the form of horizontally elongated vorticity structures. These changes have important implications for structural stability and local scour. Overall, the findings contribute to the optimization of hydraulic structure design by highlighting the effects of channel confinement on flow-induced forces.

1. Introduction

Many bridges worldwide have suffered severe damage due to scour around bridge abutments, particularly local scour at piers and piles. According to the AASHTO LRFD Bridge Design Specifications (2010), scour is considered a critical factor in the design of bridge foundations, especially under extreme event limit states. In a comprehensive review, Brandimarte et al. [1] concluded that flooding and foundation scour are among the most frequent contributors to bridge collapse, as supported by several earlier studies [2,3,4,5]. This finding is also supported by a review conducted by Zhang et al. [6], who analyzed global studies on bridge failures from 1980 to 2016. As illustrated in Figure 1, their results confirm that hydraulic causes, including scour, represent one of the leading external causes of bridge failure globally.
Research on local scour has a long history, originating from field observations and laboratory experiments that systematically investigated the flow structures around the bridge piers and characterized the vortex patterns generated by cylindrical constructions in flume conditions. This is illustrated by Dargahi [7], who used flume experiments and hydrogen bubbles to visualize the vortex structures around a circular cylinder in a flat channel bed. The study demonstrated that the upstream horseshoe vortex system is the primary driver of scour around the pier. Simpson [8] reviewed the studies on junction flows that occur near an obstacle and a flat for laminar- and turbulent-approaching boundary layers. The review concluded that the pressure gradient surrounding the obstacle creates a three-dimensional separation, allowing development into horseshoe vortices. Das et al. [9] used an acoustic Doppler velocimeter to measure the vorticity fields around different pier shapes, including square, circular, and equilateral triangular piers, under clear-water equilibrium scour conditions with a uniform sand bed. The results demonstrated that the square and equilateral triangular pier produced stronger circulation and more concentrated vorticity within the horseshoe vortex, particularly at azimuthal planes of 0° and 45°, compared to the circular pier under equilibrium scour conditions. Gautam et al. [10] compared velocity measurements around a simple circular pier and a complex pier composed of a column, an elliptical pile cap, and multiple piles. They observed that the flow fields differed significantly between the two pier types, with reverse flow prevailing downstream of the simple pier at low Reynolds numbers (Re) and shifting upstream as the Re increased.
Alongside experimental studies, computational fluid dynamics (CFD) models have been applied to investigate the horseshoe vortex around a bed-mounted cylinder. Olsen and Melaaen [11] and Olsen and Kjellesvig [12] were among the first to use three-dimensional numerical models to simulate steady flow over a circular cylinder on a mobile bed, where sediment transport and bed evolution were considered by coupling the flow field with sediment continuity and utilizing the standard k-ε model and a finite volume approach. Their results agreed well with the physical model tests [13]. On the other hand, Tseng et al. [14] used a weakly compressible large eddy simulation (LES) model to investigate the three-dimensional flow field around square and circular cylinders on an immobile bed, i.e., a non-erodible rigid surface. Their findings showed that the down-welling flow in front of the cylinder generated a horseshoe vortex. Escauriaza and Sotiropoulos [15] applied the detached eddy simulation (DES) model to examine the effect of Re on the horseshoe vortex in the junction of the cylinder and channel bed. Their simulation results highlighted that the horseshoe vortex in front of the cylinder was composed of a train of several vortical structures and was affected by Re, as demonstrated by the three-dimensional visualization of coherent structures and instantaneous image of vorticity at the symmetry plane. In a combined numerical and experimental approach, Kirkil and Constantinescu [16] employed LES modeling and laboratory experiments to visualize the turbulent horseshoe vortex and the near-wake flow behind a circular cylinder on a flat bed under subcritical (Re = 16,000) and supercritical (Re = 5 × 105) shallow channel conditions. Their analysis indicated that a single primary necklace vortex dominated the horseshoe vortex region in both flow regimes, while bimodal oscillations within the vortex core enhanced turbulent kinetic energy and Reynolds stresses.
In addition to investigations of the horseshoe vortex formed in front of a cylindrical pier in an open channel, another area of interest is the analysis of hydrodynamic forces acting on a cylinder placed in narrow open channel flows. In a study by Ranga Raju et al. [17], laboratory experiments were combined with theoretical analysis to examine the afflux, i.e., the increase in the upstream water level resulting from an obstruction, caused by three circular cylinders in open channel flows with various blockage ratios under subcritical conditions. Their findings indicated that the drag coefficient increased due to the blockage effect, leading to a decrease in the upstream free surface level, known as afflux. Davis et al. [18] quantitatively examined the impact of channel walls on the flow around a two-dimensional square cylinder constrained in a channel. They concluded that the blockage effect considerably alters the flow field and increases the drag coefficient and Strouhal number. Chu et al. [19] used an LES model to investigate the free-surface flows over a submerged rectangular bridge deck. Their study revealed that the drag coefficient remained constant when the blockage ratio was less than 14% but increased as the blockage ratio became larger than 14%. Furthermore, Chu et al. [20] and Nguyen and Lei [21] utilized CFD simulations to analyze the flow around circular cylinders, emphasizing the amplification of dynamic pressure and drag forces, particularly under confined flow conditions with varying Reynolds numbers. The blockage ratio refers to the ratio of the projected area of the cylinder to the cross-sectional area of the flow domain. These studies enhance the current understanding of how blockage ratio, cylinder geometry, and flow characteristics interact across diverse hydrodynamics environments.
More recently, Hurtado-Herrera et al. [22] conducted a numerical investigation using an LES model to examine the effects of blockage ratio (Br) and approach velocity on local scour around a vertical cylinder under clear-water and live-bed conditions. They compared two blockage ratios, Br = 0.0625 and 0.1, and found that both the scour rate and depth increased in the narrower channel due to stronger flow contraction. In their study, a channel was considered narrow when Br ≥ 0.1, a condition that led to significant flow confinement and altered wake structures. Their results were validated against experimental data from [23,24], confirming the accuracy of the modeled horseshoe vortex-induced scour mechanisms. The LES-derived flow patterns and near-bed shear stress distributions revealed modifications of vortex structures induced by blockage. These findings contribute to the understanding of morphodynamic responses under confined conditions. However, there remains a need to quantify hydrodynamic loadings (e.g., drag and lift coefficients) and vorticity components (Ωx, Ωy, and Ωz) across a range of Br to gain a deeper understanding of the interaction between flow characteristics and confinement effects in narrow-channel conditions.
In summary, previous experimental and numerical studies have contributed to the understanding of horseshoe vortex formation and hydrodynamic forces acting on cylindrical structures in open channel flows. However, most of these efforts have focused on wide channels and horizontally oriented cylinders, with limited attention being paid to confined flows and the three-dimensional vorticity components (Ωx, Ωy, and Ωz) around vertically mounted cylinders. Additionally, while local scour and flow separation have been widely studied, the systematic effects of Br on hydrodynamic loads and vortex development remain insufficiently explored. The present study addresses these gaps by employing an LES model to investigate the influence of Br (ranging from 10% to 30%) on vorticity distribution and hydrodynamic forces in narrow-channel flows. In this context, a narrow channel refers to a flow domain in which lateral sidewalls significantly affect the local hydrodynamics, typically corresponding to Br > 10%. The blockage ratio is defined as follows:
B r = D W
where D and W are the cylinder diameter and the channel width, respectively. The findings of this study aim to support the design and safety assessment of hydraulic structures operating under confined flow conditions, where flow constriction can substantially modify velocity fields, vortex dynamics, and hydrodynamic loading.

2. Numerical Model

This study combined a three-dimensional LES model with the VOF method to simulate the hydrodynamic loadings and vorticity distributions in confined channel flows with a range of widths. The numerical solver was modified from the open-source code Truchas, which was developed by Los Alamos National Laboratory [25]. The same numerical simulation has been employed to predict the free surface flow over a rectangular deck [19], dambreak flows over cylinder arrays [26], and the flow field around a cross-river rubber dam [27].
The filtered continuity equation and Navier–Stokes equations can be used to model fluid motion [28,29]:
u i ¯ x i = 0
ρ   u i ¯ t + ρ u i u j ¯ x j = P ¯ x i + ρ g δ i 3 + x j μ e f f u i ¯ x j + u j ¯ x i
where u, P, ρ, and g are the fluid velocity, pressure, density of fluid, and gravitational acceleration, respectively, and the over-bar indicates that the quantity is the spatially filtered value [30,31]. The subscripts i, j = 1, 2, and 3 represent the x, y, and z directions, respectively; t is the time; δ is the Kronecker delta; and μeff is the effective viscosity, which is represented by
μeff = μ + μSGS
where µ is the dynamic viscosity of the water and μSGS is the viscosity of the sub-grid scale turbulence. The sub-grid scale turbulence was modeled using the Smagorinsky formula [32]:
μ SGS = ρ ( C s Δ s ) 2 2 S ¯ ij S ¯ ij   1 / 2
where Cs and Sij are the Smagorinsky coefficient and the filtered rate of strain, respectively [33]:
S ¯ i j = 1 2 u i ¯ x j + u j ¯ x i
s is the characteristic length of the spatial filter. The characteristic length is calculated as follows:
Δ s = ( Δ x Δ y Δ z ) 1 / 3
In the numerical simulation, the value of the Smagorinsky coefficient was set to Cs = 0.15 based on our previous research [20,27,34,35]. The projection approach [36] was employed to solve the Poisson pressure equation and to decouple the velocity and pressure in the Navier–Stokes equations. The VOF method [37] was used to resolve the kinematics of the water surface. The volume fraction fm of water in a grid cell is governed by the following equation:
f m t + x j f m u j ¯ = 0
When fm = 1, fm = 0, or 0 < fm < 1, the computational cell is full of water, the cell is fully occupied by air, or the cell is partially occupied by water, respectively. The convergence criterion for the momentum equation was set to 10−6, and the Courant number was Cr = 0.85. The average time step was 1 × 10−3 s, and the output time step was ∆t = 0.1 s. The simulation results before dimensionless time τ = tV1/D = 30 were discarded to exclude the initial transient results of the simulation, where V1 and D are the upstream velocity and the cylinder diameter, respectively. The time-averaged velocity and pressure were computed from the simulation results when the dimensionless time τ = 30–64.

3. Model Validation

The accuracy of the present numerical model was validated by comparing its simulation results with the laboratory experiments conducted by Dargahi [7]. A circular cylinder with a diameter and a height of D = 0.15 m and 0.50 m, respectively, was installed along the centerline of a water flume. The upstream and downstream water depth was h1 = 0.2 m. Figure 2 presents the schematic of the computational domain and flow conditions. The domain was 3.15 m, 1.5 m, and 0.36 m in length, width, and height and was discretized using the finite volume method. The channel bed and sidewalls were set as no-slip boundary conditions, and the downstream boundary was defined with a zero-velocity gradient, following practices adopted in previous OpenFOAM-based simulations [38,39]. The domain was initialized with water, and a velocity field based on the inlet log-law profile was applied as the initial condition to represent a fully developed flow throughout.
A logarithmic profile was used to describe the distribution of streamwise velocity in the vertical direction:
V ( z ) u * = 1 κ l n z z 0
where the von Karman constant, shear velocity, and roughness length are κ = 0.41, u* = 0.012 m/s, and z0 = 2.5 × 10−5, respectively [7]. Note that the grid size of ∆z = 2–3 mm near the bottom wall can be normalized as z+ =zu*/ν = 24–36, which is slightly larger than the thickness of the viscous sublayer. Since the viscous sublayer is not resolved and the study focuses on large-scale flow dynamics, this wall treatment is considered appropriate, in accordance with previous LES studies [40,41,42]. The reference velocity at z = 0.18 m, y = 0, and 0.3 m from the inlet (from the cylinder face) was V1 = 0.26 m/s; this velocity value will be used for validation grids. The upstream Froude number was Fr1 = V1/(gh1)0.5 = 0.186; the Reynolds number, based on the cylinder diameter, was Re = V1D/ν = 39,000.
A three-dimensional view of the computational mesh is shown in Figure 3a. The entire computational domain was divided into ten zones. The circular cylinder was located in Zone VI, with a non-structured grid (as depicted in Figure 3b). Zone V used a uniform grid, while the other zones utilized non-uniform grids with a stretching ratio of 1.05. The grid arrangement in the z direction was divided into two layers: finer grid cells (∆z = 3 mm for Grid 1, and ∆z = 2 mm for Grid 2) in the lower layer (z = 0–30 mm) and a non-uniform grid in the upper layer (z = 30–200 mm). The vertical grid size was gradually increased (with a stretching ratio of 1.05). In the horizontal direction, the smallest grid size around the cylinder was 5 mm = D/30.
Figure 4 compares the simulated velocity profiles at the upstream boundary (x/D = 2, y/D = 0) with the measured velocity distribution from Dargahi [7]. The curve was determined using the logarithmic profile in Equation (9). The average relative difference between the simulated and experimental results was 4.9% and 4.7% for Grid 1 and Grid 2, respectively. For the validation case, the Br was 10%. The simulated and measured pressure coefficients right in front of the cylinder (at x/D = 5.46, y/D = 0) are shown in Figure 5.
The pressure coefficient (Cp) is defined as follows:
C p = P P s 1 2 ρ V 2
where P is the total pressure, Ps is the hydrostatic pressure, ρ is the water density, and V is the streamwise velocity at height z = 0.18 m.
Figure 5 shows the distribution of Cp in front of the cylinder face (at x/D = 5.46, y/D = 0), which is directly related to the non-uniform velocity profile upstream of the cylinder. Cp is calculated using Equation (10). The average differences between the simulated and measured Cp values from Dargahi [7] are 2.18% and 2.1% for Grid 1 and Grid 2, respectively, while the discrepancy between the earlier study by Tseng [14] and Dargahi [7] is 3%. The good agreement between the measured and simulated Cp on the frontal side of the cylinder demonstrates the accuracy and reliability of the current LES model.
The time-averaged drag force and lateral force were calculated from the simulated pressure on the surface of the cylinder:
F D = j = 1 m i = 1 n P ( θ i , j ) cos θ i , j ( R Δ θ ) Δ z
F L = j = 1 m i = 1 n P ( θ i , j ) sin θ i , j ( R Δ θ ) Δ z
where θi,j is the angle; R is the radius of the cylinder; n is the number of grid points on the cylinder surface; and m is the number of grid points in the z-direction. The current computation uses n = 90 (90 grid points around circumference of the cylinder) and m = 9 (9 different heights in the vertical direction). The drag coefficient, CD, is defined as
C D = F D 1 2 ρ V 2 A
where FD is the drag force acting on the cylinder and A = hD is the projected area of the circular cylinder. The lateral force coefficient, CL, can be written as
C L = F L 1 2 ρ V 2 A
where FL is the lateral force acting on the cylinder.
Table 1 compares the drag coefficients and the Strouhal number computed by two different computational grids. The time-averaged drag coefficients CD = 1.099 and 1.199 for Grid 1 and Grid 2, respectively, are very close to CD = 1.10–1.20 of the two-dimensional circular cylinder in the range 103 < Re < 105 [43,44,45]. The time-averaged lateral force coefficients are CL = 0.06 and 0.04 for Grid 1 and Grid 2, respectively. They are close to the value CL = 0 for a circular cylinder in uniform, symmetric flow. This comparison verifies the accuracy of the present LES model.
The Strouhal number (St) is computed as follows:
S t = f s D V
where fs is the shedding frequency. The shedding frequency was computed from the time series of lateral forces at a normalized height of z/h1 = 0.6, resulting in frequencies of fs = 33.6 Hz and 34.0 Hz for Grids 1 and 2, respectively. These values are in good agreement with previous results (St = 0.20–0.21) for two-dimensional circular cylinders at Re = 103–104 [46,47,48].
Figure 6 presents the simulated pressure coefficients around the cylinder at z/h1 = 0.6 using two computational grids, along with experimental data for a two-dimensional circular cylinder (Re = 4.4 × 104), as reported by Surry [49], which closely matches the Reynolds number of the present study. The good agreement between both datasets suggests that the flow fields at this height approximate two-dimensional behavior. According to potential flow theory, the theoretical Cp at the stagnation point (θ = 0°) is Cp = 1.0. The simulated values at this point are Cp = 0.991 and Cp = 0.986 for Grids 1 and 2, corresponding to deviations of 1% and 1.4%, respectively. Additionally, the experimental study [7] reported Cp = 0.965 at θ = 0°, z/h1 = 6, while the simulation by Tseng [14] yielded Cp = 0.97 (see Figure 5). This discrepancy arises because the velocity near the cylinder base is reduced by boundary layer effects, causing Cp to deviate from the theoretical value. Dargahi [7] also noted that the pressure gradients may be a result of the non-uniform velocity distribution in the approaching flow. Furthermore, the average total errors between the simulated results and the measurement of Surry [49] were 1.97% and 2.01% for Grids 1 and 2, respectively.
As shown in the numerical validation model, the results obtained using Grid 1 and Grid 2 show good agreement with the experimental data. This demonstrates that the present LES model can accurately simulate the flow field around the circular cylinder. It is worth noting that although both grids yielded comparable accuracy, the CPU time required for Grid 2 was longer than that of Grid 1 (see Table 1). Therefore, Grid 1 was selected for the remaining simulations in order to reduce computational cost.

4. Results and Discussion

4.1. Hydrodynamic Loading Effect

This section investigates the blockage effect on the hydrodynamic loadings and the vorticity distribution around a circular cylinder. The upstream and downstream water depths are identical, with h1 = 0.2 m and the cylinder diameter D = 0.15 m. The upstream velocity profile is set to the same logarithmic profile as in the validation case, with a shear velocity of u* = 0.012 m/s and a reference velocity V1 = 0.26 m/s at height z = 0.18 m. The channel widths are set as W = 1.5 m, 1.05 m, 0.75 m, 0.6 m, and 0.5 m, resulting in Br = 10%, 14.3%, 20%, 25%, and 30%, respectively.
Figure 7 illustrates the evolution of time-averaged velocity vectors around the circular cylinder at z/h1 = 0.04 for various Br values. As the Br increases, the flow exhibits enhanced separation and a more pronounced recirculation zone in the cylinder’s wake. Notably, the velocity in the gap between the cylinder and the sidewall increases significantly, and the velocity vectors become increasingly aligned with the channel walls. This behavior suggests a higher risk of erosion near the side boundaries and greater turbulence intensity downstream due to the confined flow. The observed changes in vector magnitude and direction underscore the impact of blockage on flow dynamics in narrow channels.
The time-averaged Cp on the frontal face of the cylinder (at x/D = 5.46, y/D = 0) under varying Br conditions is presented in Figure 8. As the channel width decreases, notable changes are observed in the Cp distribution along the cylinder surface. The minimum Cp consistently occurs at z/h1 = 0.1125, and its magnitude decreases with increasing Br. The interaction between the incoming turbulent boundary layer and the cylinder surface is enhanced under more confined flow conditions. This implies that the Br influences the separation-induced vortex flow at the junction of the channel bed and the cylinder. Compared to the baseline case (Br = 10%), the Cp value exhibits progressive deviations of 1.18%, 2.26%, 3.38%, and 5.61% at blockage ratios of 14.3%, 20%, 25%, and 30%, respectively. These variations provide quantitative benchmarks for assessing the impact of Br on surface pressure characteristics in hydraulic structures.
The time-averaged Cp around the circumference of the cylinder at z/h1 = 0.6 for different Br values is presented in Figure 9. As previously discussed in Figure 5, the frontal Cp (≈1.0) at θ = 0° remains consistent for Br = 10% and 14.3% but becomes slightly greater than 1.0 when Br ≥ 20%. Beyond this threshold, the overall distribution of Cp around the cylinder undergoes a notable change. The values of Cp deviate from the potential flow theory, due to flow separation, at other locations along the cylinder circumference. In addition, the minimum Cp (θ = 68–80°) and the base pressure Cpb= 120–180°) decrease with increasing Br, resulting from the high velocity in the gap region (see Figure 7). The location of minimum Cp is approximately at θ = 68° for Br = 10% and 14.3%, shifting toward the leeward side of the cylinder (θ = 76–80°) as Br increases. These results indicate that the blockage effect considerably alters the flow characteristics around the circular cylinder in an open channel.
The time-averaged drag coefficients are plotted against Br in Figure 10. It shows that the CD remains constant at CD = 1.12 when Br ≤ 14.3%, indicating that Br has a negligible effect on drag under low-blockage conditions. This behavior is consistent with the results reported for circular cylinders in unconfined domains [14]. As Br increases, the blockage effect accelerates the incoming flow and increases the dynamic pressure on the front face of the cylinder, thereby amplifying the pressure difference between upstream and downstream surfaces and resulting in a higher CD. This phenomenon was also observed by Chu et al. [19] for a horizontal rectangular cylinder in free surface flows. The contraction of the channel width not only contributed to the increase in CD but also exerted a notable influence on CL.
Figure 11 reveals the time histories of the lateral force coefficient for Br values ranging from 10% to 30%, computed using Equations (12) and (14). The results exhibit periodic fluctuations induced by vortex shedding around the cylinder. Higher blockage ratios, particularly Br = 25% and 30%, show significantly larger amplitudes and higher frequencies compared to lower-blockage conditions. This figure indicates that flow confinement intensifies the lateral forces acting on the cylinder, potentially leading to stronger dynamic responses and structural vibrations. These findings underscore the sensitivity of vortex-induced forces to geometric confinement in the flow domain. This is due to the enhanced vortex shedding, which makes the alternating low- and high-pressure zones become more prominent and amplifies the lateral force oscillations. Consequently, an increase in the root mean square of CL and St values can be anticipated, as shown in Figure 12 and Figure 13.
The amplitude of the fluctuation is expressed as the root mean square value of the lateral force coefficient CL_rms:
C L _ rms = 1 N i = 1 N C L ( t i ) C ¯ L 2
where N = 228 is the data number and C L ¯ is the time-averaged lateral force coefficient.
Figure 12 shows the variation in the root mean square of the lateral force coefficient for different blockage ratios. The values are derived from the temporal data shown in Figure 11 and calculated using Equation (16). A pronounced upward tendency in CL_rms is observed as the Br increases, particularly within the range of 10% to 30%, indicating a stronger unsteady lateral force response under higher-confinement conditions. This trend confirms that greater flow confinement enhances vortex shedding intensity, which strengthens the alternating pressure field around the cylinder. As a result, the amplitude and persistence of unsteady lateral forces increase, leading to higher CL_rms values. The diagram also compares the present results with previous experimental and computational studies [21,50,51], and the close agreement confirms the reliability of the current approach in capturing the effects of blockage on lateral force fluctuations. These results demonstrate that stronger and more coherent vortices, which are generated under high-blockage conditions, impose greater pressure asymmetry over time, thereby amplifying the overall lateral force acting on the cylinder.
The dimensionless St is plotted against Br in Figure 13. At Br = 10%, St = 0.206, which closely agrees with the experimental results [7,46]. In addition, the simulated value of St = 0.228 at Br = 25% is consistent with the findings of Richter and Naudascher [50], based on experiments involving a horizontal circular cylinder in a narrow rectangular duct. These comparisons support the validity of the LES model in investigating the effects of Br on hydrodynamic loading on circular obstacles in open channel flows. Moreover, Figure 13 shows that the St value increases when Br > 20%. This pattern results from the acceleration of flow through the narrowed gap caused by the reduced channel width, which amplifies the vortex shedding frequency and consequently elevates the St value.
As shown in the simulation results, variations in channel width under identical inlet flow conditions led to changes in hydrodynamic loadings. These findings are consistent with previous studies, underscoring the critical role of blockage in amplifying flow-induced scour and structural loading in narrow channels. Modifications in the pressure coefficient along the upstream surface and around the circumference of the cylinder at z/h1 = 0.6, along with elevated drag coefficients, demonstrate the significant influence of Br on flow behavior. Furthermore, the time history of CL reveals a noticeable growth in oscillation amplitude at Br = 20–30%, resulting in higher values of both the CL_rms and St values. These enhancements reflect intensified unsteady lateral forces acting on the structure. The diagrams illustrating these changes allow engineers to assess the impact of Br on structural performance and adapt their design plans accordingly.

4.2. Vorticity Distribution Effect

As well as the hydrodynamic loadings, this analysis also examines the vorticity distribution around the cylinder near the channel bed. Figure 14 is a schematic diagram of three-dimensional vorticity around the circular cylinder. The vorticity in Cartesian coordinates is
Ω =   w   y   v   z i +   u   z   w   x j +   v   x   u   y k
Vorticity upstream of a cylinder in unconfined channels has been the primary focus of previous studies. In contrast, the present study examines the influence of blockage on vertical vorticity in front of the cylinder and the streamwise vorticity component within the region between the sidewall and the cylinder, referred to hereafter as the gap region. Figure 15 shows the distribution of vertical vorticity (Ωy) upstream of a wall-mounted cylinder for Br = 10–30%. The contours are plotted on the symmetrical plane (y = 0), highlighting the formation of near-wall vortical structures. In all cases, a strong concentration of Ωy appears near the cylinder base, indicating intensified spanwise shear layers caused by the deflection of the incoming boundary layer. As the Br increases, both the peak vorticity and its closeness to the cylinder surface grow, suggesting stronger flow confinement and adverse pressure gradients. These changes imply that local scouring in front of circular piers may increase under narrow-channel conditions with erodible beds.
The relationship between the blockage ratio and the peak Ωy on the symmetry plane (y = 0) is shown in Figure 16. The results show a consistent increase in peak vertical vorticity as Br increases from 10% to 30%, supporting the interpretation that geometric confinement intensifies upstream vortex generation by increasing adverse pressure gradients and boundary layer deflection near the cylinder base. The evolution in Ωy_max values indicates the development of stronger near-bed shear layers and more complex three-dimensional vortex interactions at higher blockage ratios. These findings are consistent with previous studies [7,14,22,52] which reported that the obstacle in confined channel flows exhibits stronger shear-induced vorticity and more pronounced vortex deformation due to higher pressure gradients and velocity contraction near the junction between the channel bed and the cylinder surface.
Figure 17 illustrates the distribution of streamwise vorticity (Ωx) on the y-z plane across the cylinder center at x/D = 6 under varying Br conditions. The vortex rotates counterclockwise when the vorticity Ωx > 0 (represented by red color), whereas the vorticity Ωx < 0 (represented by blue color) indicates the vortex rotating clockwise. At Br = 10% and 14.3%, the vorticity structures appear compact and symmetric, indicating stable vortex formation. As Br increases from 20% to 30%, the intensity and spatial extent of Ωx increase, resulting in laterally expanded and more prominent vortical structures that reflect stronger rotational motion and elevated near-bed turbulence. These findings underscore the influence of blockage effects on near-bed vortex dynamics (Ωx), with important implications for sediment transport, scour formation, and hydrodynamic forces on structures in confined channel environments.
Building on these observations, Figure 18 summarizes the correlation between the maximum (red circles) and minimum (blue square) streamwise vorticity values (Ωx). These values are measured at the cross-section through the cylinder center (y-z plane) at x/D = 6. The vorticity magnitudes exhibit a slight increase with Br values, ranging from Ωx_max = 13.1 to 14.6 s−1 and Ωx_min = −14 to −15 s−1. The vorticity distribution remains symmetric on both sides of the cylinder, with the vortex cores (locations of peak vorticity) consistently located around y/D = ±1.02 and z/D = 0.03 across all cases. Although the channel side walls have a limited impact on the magnitude of Ωx in the gap region, the emergence of secondary vortices opposing the main vortex becomes more evident as the channel narrows. This effect is particularly pronounced when Br ≥ 20%, as shown in Figure 17. These intensified vortices expand their influence area, leading to erosion not only at the base of bridge piers but also along the channel banks, contributing to bank instability and lateral scour.
In addition to the streamwise (Ωx) and vertical (Ωy) vorticity components shown in the preceding figures, Figure 19 presents the time-averaged distribution of the spanwise vorticity (Ωz) at a height of z/D = 0.05 for various Br. The Ωz component is concentrated in the gap region between the cylinder and the channel wall, originating from the cylinder surface and extending downstream into the wake. The magnitude of Ωz increases with Br, indicating intensified vortex shedding under more confined flow conditions. At Br = 25% and 30%, two secondary vortices emerge and align downstream along the channel wall, forming a nearly straight structure. In addition, due to the no-slip boundary condition at the lateral wall, the velocity at the channel wall is zero, while the adjacent flow in the gap region remains non-zero. This velocity gradient generates an additional shear-induced vortex near the channel wall, which enhances local spanwise rotation and contributes to the three-dimensional complexity of the wake. This near-wall vortex may further intensify lateral momentum transport and promote local scour along the channel boundary.
Figure 20 shows the variation in the maximum and minimum spanwise vorticity (Ωz) at z/D = 0.05 across a range of Br. As Br increases from 10% to 30%, both Ωz_max and Ωz_min exhibit a consistent rise from approximately ±15.0 s−1 to ±18.2 s−1. The locations of these peak values remain nearly constant at x/D = 6.1 and y/D = ±0.54, indicating a persistent vortex formation zone near the cylinder. This trend reflects the intensification of shear layers within the narrowing gap caused by increased blockage, which enhances vorticity generation. The symmetrical growth of positive and negative vorticity suggests the development of stronger and more coherent rotational structures on both sides of the wake, thereby contributing to more complex flow dynamics at higher Br values.
As shown in the results above, although the effects of Br on vorticity values appear minor, their cumulative impact could influence long-term structural performance. Therefore, understanding the subtle variations in vorticity induced by blockage is essential for improving the reliability and durability of structures subjected to confined flow conditions. Overall, this insight highlights the importance of incorporating Br considerations into the design process to optimize performance and prevent potential structural failures over time, particularly in cases where Br ≥ 20%.

5. Conclusions

This study employed the LES model and VOF method to investigate the hydrodynamic forces and the vorticity distribution around a circular cylinder in a confined channel. The numerical model was validated against experimental data from [7], as detailed in Section 3. The numerical model was then utilized to investigate the impact of blockage on the vortex structure near the channel and hydrodynamic loading through a series of numerical simulations. The simulated results are as follows:
-
The blockage ratio has a significant influence on the pressure and force characteristics around the cylinder. The changes in the upstream and circumferential pressure distributions, along with elevated drag coefficients, reflect the effect of flow confinement. At Br = 20–30%, amplified oscillations in the lateral force coefficient lead to higher CL_rms and St values, highlighting the impact of blockage on hydrodynamic loading in confined channels.
-
The vorticity analysis indicates that increasing Br intensifies near-wall vortex structures, particularly through elevated vertical (Ωy), streamwise (Ωx), and spanwise (Ωz) vorticity components. As Br increases, the vorticity field becomes more dominant and spatially extensive, with Ωx and Ωz forming elongated structures and enhancing near-bed turbulence on the leeward side of the cylinder. These changes suggest that higher blockage ratios (Br = 20–30%) promote strengthening vortex activity near the bed, increasing the potential for local scour in narrow-channel conditions with erodible beds.
In conclusion, the present study confirms that Br ≥ 20% significantly affects flow characteristics around a circular cylinder in narrow channels, including the upstream and circumferential pressure coefficients, drag coefficient, lateral force coefficient, and Strouhal number. The enhanced near-wall vortex structures and shear layers at higher Br contribute to increased vorticity magnitudes, with important implications for structural performance and erosion. These findings emphasize key design considerations for engineers, particularly regarding hydrodynamic forces acting on obstacles in confined environments. However, this study was conducted at a fixed Reynolds number, which limits its ability to capture Reynolds number-dependent flow phenomena such as transition to turbulence, scale separation in vortical structures, and increased flow unsteadiness. As such, future studies are recommended to investigate the effects of varying Reynolds numbers on hydrodynamic forces and vorticity distributions around a circular cylinder in confined channels, particularly at Br = 30%.

Author Contributions

Methodology, investigation, formal analysis, visualization, writing—original draft preparation, T.T.T.T.; conceptualization, funding acquisition, supervision, writing and editing the manuscript, C.-R.C.; methodology, software, T.-R.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the National Science and Technology Council, Republic of China, Taiwan, with grant no. 112-2221-E-008-024.

Data Availability Statement

The simulation data will be available on reasonable request to the first author.

Acknowledgments

The financial support from the Ministry of Science and Technology (MOST) of Republic of China, Taiwan, under grant no. MOST 112-2221-E-008-024, is gratefully appreciated.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Brandimarte, L.; Paron, P.; Baldassarre, G.D. Bridge Pier Scour: A Review of Processes, Measurements and Estimates. Environ. Eng. Manag. J. 2012, 11, 975–989. [Google Scholar] [CrossRef]
  2. Melville, B.; Coleman, S.E. Bridge Scour; Water Resource, LLC.: Highlands Ranch, CO, USA, 2000; ISBN 9781887201186. [Google Scholar]
  3. Richardson, E.V.; Davis, S.R. Evaluating Scour at Bridges, 4th ed.; FHWA NHI 01-001 HEC-18; Federal Highway Administration (FHWA), U.S. Department of Transportation: Washington, DC, USA, 2001. [Google Scholar]
  4. Wardhana, K.; Hadipriono, F.C. Analysis of Recent Bridge Failures in the United States. J. Perform. Constr. Facil. 2003, 17, 144–150. [Google Scholar] [CrossRef]
  5. Imhof, D. Risk Assessment of Existing Bridge Structures; University of Cambridge: Cambridge, UK, 2004. [Google Scholar]
  6. Zhang, G.; Liu, Y.; Liu, J.; Lan, S.; Yang, J. Causes and Statistical Characteristics of Bridge Failures: A Review. J. Traffic Transp. Eng. (Engl. Ed.) 2022, 9, 388–406. [Google Scholar] [CrossRef]
  7. Dargahi, B. Experiments in Fluids The Turbulent Flow Field around a Circular Cylinder. Exp. Fluids 1989, 8, 1–12. [Google Scholar] [CrossRef]
  8. Simpson, R.L. Junction Flows. Annu. Rev. Fluid Mech. 2001, 33, 415–443. [Google Scholar] [CrossRef]
  9. Das, S.; Das, R.; Mazumdar, A. Vorticity and Circulation of Horseshoe Vortex in Equilibrium Scour Holes at Different Piers. J. Inst. Eng. (India) Ser. A 2014, 95, 109–115. [Google Scholar] [CrossRef]
  10. Gautam, P.; Eldho, T.I.; Mazumder, B.S.; Behera, M.R. Experimental Study of Flow and Turbulence Characteristics around Simple and Complex Piers Using PIV. Exp. Therm. Fluid Sci. 2019, 100, 193–206. [Google Scholar] [CrossRef]
  11. Olsen, N.R.B.; Melaaen, M.C. Three-Dimensional Calculation of Scour around Cylinders. J. Hydraul. Eng. 1993, 119, 1048–1054. [Google Scholar] [CrossRef]
  12. Olsen, N.R.B.; Kjellesvig, H.M. Three-Dimensional Numerical Flow Modeling for Estimation of Maximum Local Scour Depth. J. Hydraul. Res. 1998, 36, 579–590. [Google Scholar] [CrossRef]
  13. Torsethaugen, K. Lokalerosjon ved Store Konstruksjoner; SINTEF Report No. STF60A75055; The Norwegian Hydrotechnical Laboratory: Trondheim, Norway, 1975. (In Norwegian) [Google Scholar]
  14. Tseng, M.H.; Yen, C.L.; Song, C.C.S. Computation of Three-Dimensional Flow around Square and Circular Piers. Int. J. Numer. Meth. Fluids 2000, 34, 207–227. [Google Scholar] [CrossRef]
  15. Escauriaza, C.; Sotiropoulos, F. Reynolds Number Effects on the Coherent Dynamics of the Turbulent Horseshoe Vortex System. Flow Turbul. Combust. 2011, 86, 231–262. [Google Scholar] [CrossRef]
  16. Kirkil, G.; Constantinescu, G. Effects of Cylinder Reynolds Number on the Turbulent Horseshoe Vortex System and near Wake of a Surface-Mounted Circular Cylinder. Phys. Fluids 2015, 27, 075102. [Google Scholar] [CrossRef]
  17. Ranga Raju, K.G.; Rana, O.P.S.; Asawa, G.L.; Pillai, A.S.N. Rational Assessment of Blockage Effect in Channel Flow Past Smooth Circular Cylinders. J. Hydraul. Res. 1983, 21, 289–302. [Google Scholar] [CrossRef]
  18. Davis, R.W.; Moor, E.F.; Purtell, L.P. A Numerical Experimental Study of Confined Flow around Rectangular Cylinders. Phys. Fluids 1984, 27, 40–59. [Google Scholar] [CrossRef]
  19. Chu, C.R.; Chung, C.H.; Wu, T.R.; Wang, C.Y. Numerical Analysis of Free Surface Flow over a Submerged Rectangular Bridge Deck. J. Hydraul. Eng. 2016, 142, 108567. [Google Scholar] [CrossRef]
  20. Chu, C.R.; Lin, Y.A.; Wu, T.R.; Wang, C.Y. Hydrodynamic Force of a Circular Cylinder Close to the Water Surface. Comput. Fluids 2018, 171, 154–165. [Google Scholar] [CrossRef]
  21. Nguyen, Q.D.; Lei, C. Hydrodynamic Characteristics of a Confined Circular Cylinder in Cross-Flows. Ocean Eng. 2021, 221, 108567. [Google Scholar] [CrossRef]
  22. Hurtado-Herrera, M.; Zhang, W.; Hammouti, A.; Pham, V.B.; Nguyen, K.D. Numerical Study of the Flow and Blockage Ratio of Cylindrical Pier Local Scour. Appl. Sci. 2023, 13, 11501. [Google Scholar] [CrossRef]
  23. Lachaussée, F. Érosion et Transport de Particules au Voisinage d’un Obstacle. Ph.D. Thesis, Université Paris-Saclay, Paris, France, 2018. [Google Scholar]
  24. Roulund, A.; Sumer, B.M.; Fredsøe, J.; Michelsen, J. Numerical and Experimental Investigation of Flow and Scour around a Circular Pile. J. Fluid Mech. 2005, 534, 351–401. [Google Scholar] [CrossRef]
  25. Korzekwa, D.A. Truchas—A Multi-Physics Tool for Casting Simulation. Int. J. Cast Met. Res. 2009, 22, 187–191. [Google Scholar] [CrossRef]
  26. Wu, T.R.; Vuong, T.H.N.; Lin, J.W.; Chu, C.R.; Wang, C.Y. Three-Dimensional Numerical Study on the Interaction between Dam-Break Wave and Cylinder Array. J. Earthq. Tsunami 2018, 12, 1840007. [Google Scholar] [CrossRef]
  27. Chu, C.R.; Tran, T.T.T.; Wu, T.R. Numerical Analysis of Free-Surface Flows over Rubber Dams. Water 2021, 13, 1271. [Google Scholar] [CrossRef]
  28. Deardorff, J.W. A Numerical Study of Three-Dimensional Turbulent Channel Flow at Large Reynolds Numbers. J. Fluid Mech. 1970, 41, 453–480. [Google Scholar] [CrossRef]
  29. Pop, S.B. Turbulent Flows; Cambridge University Press: Cambridge, UK, 2000; ISBN 978-0-5215-9886-6. [Google Scholar]
  30. O’Neil, J.; Meneveau, C. Subgrid-Scale Stresses and Their Modelling in a Turbulent Plane Wake. J. Fluid Mech. 1997, 349, 253–293. [Google Scholar] [CrossRef]
  31. Cabot, W.; Moin, P. Approximate Wall Boundary Conditions in the Large Eddy Simulation of High Reynolds Number Flow. Flow Turbul. Combust. 1999, 63, 269–291. [Google Scholar] [CrossRef]
  32. Smagorinsky, J. General Circulation Experiments with the Primitive Equation I: The Basic Experiment. Mon. Weather Rev. 1963, 91, 99–164. [Google Scholar] [CrossRef]
  33. Rodi, W.; Constantinescu, G.; Stoesser, T. Large-Eddy Simulation in Hydraulics; CRC Press: Boca Raton, FL, USA, 2013; ISBN 978-0-3675-7638-7. [Google Scholar]
  34. Huynh, L.E.; Chu, C.R.; Wu, T.R. Hydrodynamic Loads of the Bridge Decks in Wave-Current Combined Flows. Ocean Eng. 2023, 270, 113520. [Google Scholar] [CrossRef]
  35. Truong, N.M.; Wu, T.R.; Chu, C.R.; Wang, C.Y. A Numerical Study of Plunging Breakers in the Nearshore Area under the Influence of Wind. Ocean Eng. 2024, 312, 119171. [Google Scholar] [CrossRef]
  36. DeLong, M. Two Examples of the Impact of Partitioning with Chaco and Metis on the Convergence of Additive-Schwarz Preconditioned FGMRES; Los Alamos National Laboratory: Los Alamos, NM, USA, 1997. [Google Scholar]
  37. Hirt, C.W.; Nichols, B.D. Volume of Fluid (VOF) Method for the Dynamics of Free Boundaries. J. Comput. Phys. 1981, 39, 201–225. [Google Scholar] [CrossRef]
  38. Iqbal, S.; Dissanayaka, K.D.C.R.; Tanaka, N. Numerical Modeling and Validation of Dike-Induced Water Flow Dynamics Using OpenFOAM. ISH J. Hydraul. Eng. 2024, 30, 559–573. [Google Scholar] [CrossRef]
  39. Anjum, N.; Iqbal, S.; Pasha, G.A.; Tanaka, N.; Ghani, U. Optimizing Coastal Forest Arrangements for Tsunami Flow Dynamics Using a Three-Dimensional Approach. Phys. Fluids 2025, 37, 035197. [Google Scholar] [CrossRef]
  40. Piomelli, U.; Chasnov, J.R. Chapter 7 Large-Eddy Simulations: Theory and Applications; Kluwer Academic Publishers: Dordrecht, The Netherlands, 1996. [Google Scholar]
  41. Salim, S.M.; Cheah, S.C. Wall Y+ Strategy for Dealing with Wall-Bounded Turbulent Flows. In Proceedings of the International MultiConference of Engineers and Computer Scientists, Hong Kong, China, 18–20 March 2009; Newswood Ltd.: Hong Kong, China; International Association of Engineers: Hong Kong, China, 2009; Volume II, p. 2210. [Google Scholar]
  42. Larsson, J.; Kawai, S. Wall-Modeling in Large Eddy Simulation: Length Scales, Grid Resolution and Accuracy. Cent. Turbul. Res. Annu. Res. Briefs 2010, 39–46. [Google Scholar]
  43. Hoerner, S.F. Fluid Dynamic Drag: Theoretical, Experimental and Statistical Information. In Hoerner Fluid Dynamics; Hoerner Fluid Dynamics: Bakersfield, CA, USA, 1965. [Google Scholar]
  44. Achenbach, E. Distribution of Local Pressure and Skin Friction around a Circular Cylinder in Cross-Flow up to Re = 5 × 106. J. Fluid Mech. 1968, 34, 625–639. [Google Scholar] [CrossRef]
  45. Blevins, R.D. Applied Fluid Dynamics Handbook; Van Nostrand Reinhold Co.: New York, NY, USA, 1984. [Google Scholar]
  46. Roshko, A. On the Drag and Shedding Frequency of Two-Dimensional Bluff Bodies; Government Printing Office: Washington, DC, USA, 1954. [Google Scholar]
  47. Schlichting, H. Boundary Layer Theory, 7th ed.; McGraw-Hill Book Company: New York, NY, USA, 1979. [Google Scholar]
  48. Achenbach, E.; Heinecke, E. On Vortex Shedding from Smooth and Rough Cylinders in the Range of Reynolds Numbers 6 × 103 to 5 × 106. J. Fluid Mech. 1981, 109, 239–251. [Google Scholar] [CrossRef]
  49. Surry, D. Some Effects of Intense Turbulence on the Aerodynamics of a Circular Cylinder at Subcritical Reynolds Number. J. Fluid Mech. 1972, 52, 543–563. [Google Scholar] [CrossRef]
  50. Richter, A.; Naudascher, E. Fluctuating Forces on a Rigid Circular Cylinder in Confined Flow. J. Fluid Mech. 1976, 78, 561–576. [Google Scholar] [CrossRef]
  51. Norberg, C. Fluctuating Lift on a Circular Cylinder: Review and New Measurements. J. Fluids Struct. 2003, 17, 57–96. [Google Scholar] [CrossRef]
  52. Baker, C.J. The Laminar Horseshoe Vortex. J. Fluid Mech. 1979, 95, 347–367. [Google Scholar] [CrossRef]
Figure 1. Bridge failure statistics (1980–2016) by region and cause: (a) documented bridge failures by region (USA, China, and worldwide); (b) distribution of typical bridge failure causes [6].
Figure 1. Bridge failure statistics (1980–2016) by region and cause: (a) documented bridge failures by region (USA, China, and worldwide); (b) distribution of typical bridge failure causes [6].
Water 17 02366 g001
Figure 2. Schematic diagram of the flow condition, where the cylinder diameter D = 0.15 m.
Figure 2. Schematic diagram of the flow condition, where the cylinder diameter D = 0.15 m.
Water 17 02366 g002
Figure 3. Computational domain and mesh arrangement. (a) Three-dimensional view where the yellow dashed line represents the still water depth. (b) Top view of the mesh around the cylinder.
Figure 3. Computational domain and mesh arrangement. (a) Three-dimensional view where the yellow dashed line represents the still water depth. (b) Top view of the mesh around the cylinder.
Water 17 02366 g003
Figure 4. Velocity profile at the inlet boundary (at x/D = 2 and y/D = 0) for the experimental case [7].
Figure 4. Velocity profile at the inlet boundary (at x/D = 2 and y/D = 0) for the experimental case [7].
Water 17 02366 g004
Figure 5. Comparison of simulated and measured pressure coefficients on the frontal side of the circular cylinder [7,14].
Figure 5. Comparison of simulated and measured pressure coefficients on the frontal side of the circular cylinder [7,14].
Water 17 02366 g005
Figure 6. Validation of simulated pressure coefficients against experimental data on the cylinder surface at z/h1 = 0.6 [49].
Figure 6. Validation of simulated pressure coefficients against experimental data on the cylinder surface at z/h1 = 0.6 [49].
Water 17 02366 g006
Figure 7. Time-averaged velocity vectors (at z/h1 = 0.04) around the cylinder. (a) Br = 10%; (b) Br = 14.3%; (c) Br = 20%; (d) Br = 25%; (e) Br = 30%.
Figure 7. Time-averaged velocity vectors (at z/h1 = 0.04) around the cylinder. (a) Br = 10%; (b) Br = 14.3%; (c) Br = 20%; (d) Br = 25%; (e) Br = 30%.
Water 17 02366 g007
Figure 8. Time-averaged pressure coefficient on the frontal side of the circular cylinder (at x/D = 5.46 and y/D = 0).
Figure 8. Time-averaged pressure coefficient on the frontal side of the circular cylinder (at x/D = 5.46 and y/D = 0).
Water 17 02366 g008
Figure 9. Time-averaged pressure coefficient on the cylinder surface at height z/h1 = 0.6.
Figure 9. Time-averaged pressure coefficient on the cylinder surface at height z/h1 = 0.6.
Water 17 02366 g009
Figure 10. Relationship between the drag coefficient of circular cylinder and blockage ratios [7,14].
Figure 10. Relationship between the drag coefficient of circular cylinder and blockage ratios [7,14].
Water 17 02366 g010
Figure 11. Time histories of lateral forces on the cylinder across a range of Br.
Figure 11. Time histories of lateral forces on the cylinder across a range of Br.
Water 17 02366 g011
Figure 12. Effect of blockage ratio on the root mean square of the lateral force coefficient [21,50,51].
Figure 12. Effect of blockage ratio on the root mean square of the lateral force coefficient [21,50,51].
Water 17 02366 g012
Figure 13. Relationship between the Strouhal number and blockage ratios [7,14,50].
Figure 13. Relationship between the Strouhal number and blockage ratios [7,14,50].
Water 17 02366 g013
Figure 14. Schematic diagram of the horseshoe vortex around a circular cylinder in an open channel flow.
Figure 14. Schematic diagram of the horseshoe vortex around a circular cylinder in an open channel flow.
Water 17 02366 g014
Figure 15. Distribution of the vorticity Ωy on the symmetric plane (y = 0) of the cylinder at z/D = 0–0.2. (a) Br = 10%; (b) Br = 14.3%; (c) Br = 20%; (d) Br = 25%; (e) Br = 30%.
Figure 15. Distribution of the vorticity Ωy on the symmetric plane (y = 0) of the cylinder at z/D = 0–0.2. (a) Br = 10%; (b) Br = 14.3%; (c) Br = 20%; (d) Br = 25%; (e) Br = 30%.
Water 17 02366 g015aWater 17 02366 g015b
Figure 16. Maximum vorticity Ωy in front of the cylinder on the symmetric plan (y = 0) of the cylinder.
Figure 16. Maximum vorticity Ωy in front of the cylinder on the symmetric plan (y = 0) of the cylinder.
Water 17 02366 g016
Figure 17. Vorticity distribution (Ωx) on the y-z plane (x/D = 6.0) of the cylinder. (a) Br = 10%; (b) Br = 14.3%; (c) Br = 20%; (d) Br = 25%; (e) Br = 30%.
Figure 17. Vorticity distribution (Ωx) on the y-z plane (x/D = 6.0) of the cylinder. (a) Br = 10%; (b) Br = 14.3%; (c) Br = 20%; (d) Br = 25%; (e) Br = 30%.
Water 17 02366 g017
Figure 18. Maximum and minimum vorticity Ωx on the y-z plane (x/D = 6.0) versus blockage ratio.
Figure 18. Maximum and minimum vorticity Ωx on the y-z plane (x/D = 6.0) versus blockage ratio.
Water 17 02366 g018
Figure 19. Distribution of the vorticity Ωz on the x-y plane of the cylinder at height z/D = 0.05. (a) Br = 10%; (b) Br = 14.3%; (c) Br = 20%; (d) Br = 25%; (e) Br = 30%.
Figure 19. Distribution of the vorticity Ωz on the x-y plane of the cylinder at height z/D = 0.05. (a) Br = 10%; (b) Br = 14.3%; (c) Br = 20%; (d) Br = 25%; (e) Br = 30%.
Water 17 02366 g019
Figure 20. Maximum and minimum vorticity Ωz on the x–y plane (at height z/D = 0.05) for different Br.
Figure 20. Maximum and minimum vorticity Ωz on the x–y plane (at height z/D = 0.05) for different Br.
Water 17 02366 g020
Table 1. Simulation results of the validation case using different computational grids.
Table 1. Simulation results of the validation case using different computational grids.
GridDargahi’s
Experiment
Grid 1Grid 2
Total grid no.-949,3261,562,615
Grid no. on the cylinder surface-90120
Smallest grid size-Δx = 6 mmΔx = 5 mm
Δy = 6 mmΔy = 5 mm
Δz = 3 mmΔz = 2 mm
Strouhal no.0.200.2060.208
Relative error-3.0%4.0%
CD1.151.0991.199
Relative error-4.43%4.26%
CL00.060.04
CPU time-255 h416 h
Note: Flow parameters: D = 0.15 m, h1 = 0.2 m, V1 = 0.26 m/s.
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

Tran, T.T.T.; Chu, C.-R.; Wu, T.-R. Effect of Hydrodynamic Loadings and Vorticity Distribution on a Circular Cylinder in a Narrow Channel. Water 2025, 17, 2366. https://doi.org/10.3390/w17162366

AMA Style

Tran TTT, Chu C-R, Wu T-R. Effect of Hydrodynamic Loadings and Vorticity Distribution on a Circular Cylinder in a Narrow Channel. Water. 2025; 17(16):2366. https://doi.org/10.3390/w17162366

Chicago/Turabian Style

Tran, Truc Thi Thu, Chia-Ren Chu, and Tso-Ren Wu. 2025. "Effect of Hydrodynamic Loadings and Vorticity Distribution on a Circular Cylinder in a Narrow Channel" Water 17, no. 16: 2366. https://doi.org/10.3390/w17162366

APA Style

Tran, T. T. T., Chu, C.-R., & Wu, T.-R. (2025). Effect of Hydrodynamic Loadings and Vorticity Distribution on a Circular Cylinder in a Narrow Channel. Water, 17(16), 2366. https://doi.org/10.3390/w17162366

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