Next Article in Journal
Review of Prediction Models for Chloride Ion Concentration in Concrete Structures
Next Article in Special Issue
Interpretation of the Jiangnan Landscape and Countryside (Shan-Shui) Pattern: Evidence from the Classification and Spatial Form of Traditional Settlements in the Nanxi River Basin
Previous Article in Journal
Life Cycle Assessment of a Structural Insulated Panel Modular House in New Zealand
Previous Article in Special Issue
Automated Generation of Urban Spatial Structures Based on Stable Diffusion and CoAtNet Models
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Optimization of Anti-Scour Device Combined with Perforated Baffle and Ring-Wing Plate Based on a Multi-Factor Orthogonal Experiment

1
School of Civil Engineering and Architecture, Zhejiang Sci-Tech University, Hangzhou 310018, China
2
China Merchants Chongqing Communications Technology Research & Design Institute Co., Ltd., Chongqing 400067, China
*
Author to whom correspondence should be addressed.
Buildings 2025, 15(1), 148; https://doi.org/10.3390/buildings15010148
Submission received: 29 November 2024 / Revised: 19 December 2024 / Accepted: 4 January 2025 / Published: 6 January 2025
(This article belongs to the Special Issue Advanced Technologies for Urban and Architectural Design)

Abstract

:
In this paper, a new anti-scour device combined with a perforated baffle and ring-wing plate is proposed to enhance the traditional method for better protection of bridge piers from local scour. Based on computational fluid dynamics (CFD), the orthogonal experiments investigated the general laws of the influence of the main factors, such as the ratio of baffle perforated, the position of baffle, and the height of ring-wing plate on the anti-scour effect. Under the protection of the combined device, the maximum scour depth reduction rate in front of the pier is between 65.18% and 81.01%, while that at the side of the pier is between 52.63% and 68.42%. Especially when the perforated ratio is 20%, the baffle is 2d (d is diameter of the pier) away from the pier, and the ring-wing plate is located at 1/3 of water depth, the anti-scour effect is the best. Also, the flow field around the pier under the protection of the combined device is further investigated. The results show that the structure blocks the down-flow actively and diverts and dissipates the flow energy to decrease flow below the critical velocity of sediment. Thus, the device combined with perforated baffle and ring-wing plate has a prominent anti-scour effect and provides a basis for further studies and engineering application.

1. Introduction

As vital transportation infrastructure, sea-crossing bridges play a key role in the comprehensive development of coastal regions. Therefore, it is essential to secure the safety of hydraulic structures such as piers. However, numerous cases and theoretical studies show that the complex ocean environment will lead to scouring around the pier [1,2]. Through investigation and research on the existing bridge monitoring systems, as well as the bridge condition rating systems, scholars like Matos [3] and Brighenti [4] have discovered that the existing bridge scour monitoring systems have difficulties in effectively monitoring the scour situation. Thus, Ayoubloo et al. [5], Pang et al. [6], and Jiawang Zhan et al. [7] have proposed modern technology to predict scour depth. Similarly, based on the scour mechanisms [8,9], proposing reasonable measures to reduce local scour is of enormous economic and social significance.
Existing research shows that the anti-scour measures can be divided into active protection and passive protection [10]. The first category is carried out by changing the flow structure around the pier, while passive protection is carried out by improving the anti-scour capacity of the riverbed through riprap countermeasure [11] and so on.
As discussed and concluded by Mohsen et al. [12], Liang and Wang [13], Fouli and Elsebaie [14], Shan Wang et al. [15], and many other scholars, compared with passive protection, active protection is more economical and effective, with better application prospects [16].
Numerous studies have been carried out on flow-altering devices. Zarrati et al. [17] tested the effectiveness of collars on rectangular piers, and a 74% reduction in maximum scour depth was obtained. Ghorbani and Kells [18] observed that the plate diverted the flow from the pier and reduced the strength of horseshoe vortex. Cheng et al. [19] figured out that the plate was most effective when located at 1/3 water depth, and the length of the plate was the same as the radius of the pier. Under clear-water conditions and uniform bed materials, Karami et al. [20] investigated the anti-scour effect of collar on abutment experimentally. Park et al. [21,22] studied the effect of sacrificial piles on scour through experiments and empirical formulas. And Oral Yagci et al. [23] studied scour and deposition patterns around hexagonal arrays of circular cylinders; the results showed that this method reduced 27% scour volume and 22% scour depth compared to a single solid cylinder counterpart. Tsung-chow et al. [16] tested a small detached ring and obtained a 70% reduction in the initial stage of scouring, which was effective for all angles of attack.
However, all of the approaches used to decrease scour have certain shortcomings, it is indeed difficult to choose a device that can counter the scour effectively under all adverse and unpredictable field conditions. Thus, researchers have been studying combinations of these devices so that the combinations are able to reduce the local scour more effectively. Vittal et al. [24] tested a combination of a full-depth pier group with a collar plate; the results showed that the full pier group was seen to be as effective as a solid cylinder with a collar 3.5 times its diameter. And Mashahir et al. [25] examined a combination of continuous collars and a riprap in clear-water scour conditions. Moncada-M et al. [26] found a combination of collar and piercing of a rectangular slot was able to almost entirely eliminate the scour for the considered flow conditions. Considering the advantages of active countermeasures and the above research experience, the current study proposes an new anti-scour device combined with two active protection devices: a perforated baffle and a ring-wing plate. To the best of our knowledge, there is no report on the on the combination of these two devices.
The aim of the present work is two-fold: (1) to explore the best combination of the new device through an orthogonal experiment to achieve the best anti-scour effect; (2) to verify the effectiveness of the combined optimization device and investigate the underlying mechanisms by a detailed description of the flow field on the horizontal and vertical planes, as well as thoroughly analyzing the flow velocity changes around the pier before and after the installation of the combined device with the help of empirical formulas.
Different from the optimal design of a single device, since the overall efficiency of a combined device is not the linear sum of the individual efficiency of each device [12], an orthogonal experiment of multiple factors and two observation indicators can more effectively show the impact of each parameter on scour depth. In addition, the study on the variation of the flow field on the vertical plane and the variation of the sediment starting velocity on the horizontal plane can further verify the effectiveness of the optimal design of the combined device. The results show that the anti-scour effect of the combined device is better than most of the measures, and it is a reasonable choice for an existing pier to reduce local scour. This paper provides a method and theoretical basis for the design and mechanism analysis of similar types of combined anti-scour devices.

2. Numerical Model

2.1. Governing Equations and Turbulence Model

2.1.1. Governing Equations

In this paper, the mixture model, a simplified two-fluid model that assumes local equilibrium at small spatial scales, is used to solve the equations for the continuity, momentum, and energy conservation of the mixed phase and the volume fraction equation for the second phase, slip velocity, and drift velocity formulas.
Considering incompressible and unsteady flow, the Reynolds averaged N-S equations are as follows:
ρ u i t + ρ u j u i x j p x i + μ u i + ρ τ i j x j , i , j = 1 , 2 , 3
The continuity equations are as follows:
u i x i = 0
in which xi (i = 1, 2, 3) are three directions in Cartesian coordinates; ui (i = 1, 2, 3) are corresponding velocity components; ρ is density of fluid; p is pressure; t is flow time; Re(=u0D0/v) is Reynolds number, defined by the flow velocity u0 and the characteristic length D0; μ is dynamic viscosity; and τij is Reynolds effect force. Boussinesq proposed the eddy viscosity μt, which relates the Reynolds stress to the partial derivative of velocity.
τ i j = μ t ρ u i x j + u j x i 2 3 δ i j , k , i , j = 1 , 2 , 3
in which k is turbulent kinetic energy, and δij is the sign of Kronecker.

2.1.2. Turbulence Model

In view of the fact that the k-ε turbulence model, as a two-equation model based on Reynolds averaging, has the ability to effectively describe the key physical quantities of turbulent flow, by solving the equations of turbulent kinetic energy and turbulent kinetic energy dissipation rate, it can better capture the basic characteristics of turbulence. Especially for the complex flow conditions around bridge piers, including wake regions, boundary layer separation, and vortices, it can reasonably describe the turbulent kinetic energy and its dissipation situation and effectively characterize the turbulent behavior. Moreover, compared with more complex models, such as the Reynolds Stress Model (RSM), the k-ε turbulence model requires less computational resources and time. In addition, the k-ε turbulence model is relatively mature, with stable mathematical expressions and solution methods. The ANSYS-FLUENT used in this paper provides good support for it, and there are relatively complete setting options in terms of boundary condition treatment and discretization format selection. Therefore, in order to obtain relevant simulation results efficiently and accurately, this paper selects the k-ε turbulence model as the finite element calculation model.
This study adopts the standard k-ε turbulence model. The model parameters are shown in Table 1. And the governing equations are as follows:
ρ k t + ρ u i k x i   = x i μ + μ t σ k k x i + G k ρ ε
ρ ε t + ρ u i ε   x i   = x i μ + μ t σ ε ε x i + G ε 1 ε k G k ρ G ε 2 ε 2 k
μ t = ρ G μ k 2 ε
in which Gk = 2μtSijSij is source term of turbulent kinetic energy and Sij = ½(∂uj/∂xi + ∂ui/∂xj) is average strain rate. Gu, Gε1, and Gε2 are empirical constants; σk and σε are Prandtl numbers corresponding to turbulent kinetic energy and dissipation rate, respectively.
It should be pointed out that the particle dispersion motion theory caused by uniform turbulence is used to describe the characteristics of solid phase turbulence. The relevant equations, turbulent kinetic energy, and dispersion coefficient of the particle phase are given by two different time scales. The first is the characteristic example relaxation time scales associated with particle inertial effects.
τ F . s f = α s ρ f K s f 1 ρ s ρ f + C V
Lagrangian integral time scale computed along particle trajectories:
τ t . s f = τ t , f 1 1 + C β ξ 2
in which C V is the additional quality factor, and the value is 0.5; the turbulent vortex characteristic time scale is τ t , f = 3 2 C μ K f ε f ; ξ = V r 2 3 k f ; V r is the average value of liquid–solid relative velocity; C β = 1.8 1.35 c o s 2 θ 0 ; and θ0 is the angle between the average particle velocity and the relative velocity of the group.

2.1.3. Sediment Transport Model

Sediment transported by water above the river bed can be divided into suspended load and bed load. suspended load means sediment carried by flow settles down sufficiently slowly without any touching on river bed, If the silt carried by the flow settles down slowly enough without contacting the river bed, it is referred to as suspended load; otherwise, it is referred to as bed load. The sediment transport form depends on the size of sediments and flow velocity, which can be judged by the critical diameter
d = U 2 360 · g
where dc is the critical diameter to distinguish suspended load and bed load, and U is the mean approach flow velocity. The average flow velocity is calculated by Equation (9), and the incipient velocity of sediment with the particle size in this research is U = 0.25 m/s, and the d = 1.77 × 10−5 m, smaller than the critical diameter dc = 0.8 × 10−4 m. Therefore, the sediment will be transported in the form of bed load.
The accuracy of the model calculation largely depends on the accuracy of bed load transport rate, which can be calculated by the following equation:
q b , i = q b τ i τ C q b h x i
where q b , i is the component of time-averaged volumetric sediment transport rate per unit width; τi (i = x, y, z) is the bed shear stress vector; C = 1.5 is an empirical constant; h is the bed elevation; and q b is the time-averaged volumetric sediment transport rate per unit width on the horizontal bed, which can be calculated by the following equations:
q b = 0.053 s 1 0.5 g 0.5 d 50 1.5 D 0.3 T 2.1 T < 2.5
q b = 0.100 s 1 0.5 g 0.5 d 50 1.5 D 0.3 T m 1.5 T 2.5
where s is the relative density of sediment; d50 is the median sediment grain size; D = d 50 [ ( s 1 ) g / v 2 ] 1 / 3 is the dimensionless sediment parameter; and v is the viscosity coefficient. T and Tm can be recovered from the following equations:
T = τ τ b , c r
T m = λ τ τ b , c r τ b , c r
where τ is the shearing stress of riverbed bottom; τ b , c r is the critical bed shear stress; and λ is the modification factor, which can be obtained through
λ = e 0.45 α + 0.2 β
α = 1 d 84 / d 50 + d 50 / d 16 / 2
β = d 50 / d 90 d 10 / d 50
where di is the sediment size.

2.2. Computational Domain and Mesh

As shown in Figure 1, the numerical model is a K-epsilon two-phase flow model, adopting the standard wall function. Given that the flow fields near the cylindrical pier and baffle are rather complex, the grids in these parts have been refined locally. In this paper, the grid independence was numerically verified by establishing different grid sizes, and it was found that the result errors were within a controllable range. It is known from the meshing software that the number of grids is approximately 2.5 × 10⁵, and the average quality of grids is above 0.85 (the closer to 1, the better). The maximum skew-ness is 0.84, while the user manual requires that it should not be greater than 0.95.
The numerical model in this paper is established by using ANSYS-FLUENT. The model is 1.2 m long and 0.8 m wide. And the pier diameter is 0.04 m, which locates at the center of the model, 0.6 m away from the water inlet and outlet and 0.4 m away from the two side walls. All the flume properties are the same as Wang’s [27] flume in the laboratory. And Melville and Coleman [28] indicated that the flume width should be at least 10 times greater than the pier diameter to minimize any contraction effects on the scour depth. In this study, the flume width is 0.8 m, and a 0.04 m pier diameter was selected. Therefore, the resultant ratio of the flume width to the pier diameter is 20, and it satisfies this criterion. As indicated in Figure 2, the water depth of the upper layer is 0.1 m, and the sand in the lower layer is 0.15 m. As stated by Melville [29], the ratio of the diameter of the bridge pier (d) to the average size of the sediment particles (d50) should be greater than 25. In this research, the mean grain size was set as d50 = 0.80 mm, so d/d50 = 50 is evaluated.

2.3. Boundary Conditions and Method

In order to apply the boundary conditions, the upstream is set as Specified Velocity (u0 = 0.25 m/s) and downstream boundary is set as Outflow. The top boundary is set as Symmetry. Also, the Wall boundary condition is defined for left and right walls, pier, and baffle, which act as virtual frictionless walls. Applications of boundary conditions to this model are shown as Figure 2.
The mixture model obtains the flow field information and the concentration distribution of each phase by solving the governing equations of the mixture and the volume fraction and relative velocity of the second phase. And it determines the scour surface by defining the water–sand interface with the sand volume fraction in CFD-Post. Using the SIMPLE algorithm, which iteratively solves the discrete form of the N-S equation, a pressure field is given and corrected, and the next iterative calculation is started with the corrected pressure field and velocity field until convergence. And the residuals of the continuity equation, velocity, turbulent kinetic energy, and other variables are 1 × 10−6. Standard initialization from the velocity inlet and the time step size is 0.01 s. According to Melville [30], scour mainly occurs in the first 30 min of the test and is almost fully developed; after that, the growing process is not intensive and goes very slow. Therefore, in order to save time and improve efficiency, the model in this paper only calculates the scour in the first 30 min.

3. Test Design Using Orthogonal Method

3.1. Definition of Variables

As shown in Figure 3, for the aim of determining the optimal location and dimension of the combined device to achieve the least scour depth, the variables of the combined device considered are the perforated ratio of the baffle (S), the distance between baffle and pier (L), and the height of ring-wing plate (H).
Based on the outcomes of Cheng et al. [19], the extension length of the ring-wing plate is consistent with the radius of the pier, and three values were considered for H (i.e., H = 1/2, 1/3, 1/6 h; h is the water depth). Huang et al. [31] reported that the optimal parameters of the surface guide panels are an interior angle θ = 60°, L/D = 2–2.5, and PD/h = 0.7. Similarly, the interior angle of perforated baffle is 60°, and the height is 3/4 h. Furthermore, considering the diversion and permeability of the baffle, three values of S are tested (i.e., S = 10, 20, 30%). And this structure is located at three different spacings from the pier (i.e., L = 2, 3, and 4 d). It should be noted that, first of all, an oblique flow may affect the efficiency of the device, and the device proposed in this study is mainly intended to reduce the down-flow in front of the pier and the flow intensity near the bed surface in front of the pier. In addition, changes in the angle of attack seem to have an unsafe effect on piers as the approach flow cannot pass smoothly around the pier, forming stronger down-flow and a horseshoe vortex [10]. This design is the same as the ideas of Huang et al. [31] and Lauchlan [32] on similar structures, such as surface guide panels and permeable sheet tiles.

3.2. Test Plan

In order to avoid the influence of human factors on laboratory experiments and improve the experimental efficiency, this paper selected ANSYS-FLUENT finite element software to conduct the simulation experiments. The experimental parameters are shown in Table 2. There are three independent variables, and for each of them, three levels are considered. Moreover, the typical parameters of the three levels are taken out and designed as three grades for test grouping. In this experiment, the interaction among these three variables is ignored, and an orthogonal experimental scheme L9(3⁴) with 3 factors and 3 levels is designed. In addition, the single pier and the single ring-wing plate are set as the control tests. The configurations of all the tests are listed in Table 2. Compared with the full test, only 11 tests have been carried out in this paper, which has greatly improved the work efficiency.

4. Results and Discussion

4.1. Model Verification

4.1.1. Flow Field Verification

Local scour is a complex process in the flow field, so whether the flow field around the pier can be accurately simulated is very important to ensure the calculation accuracy of local scour. Thus, to verify the accuracy of the finite element model in this paper, the simulated results, including the streamlines and scour depth, are fully compared with laboratory experiment. The upper part of Figure 4 shows the simulation results of the streamlines, which are around the pier and 1 mm above the riverbed.
The experimental results of streamlines at the same section are also given in Figure 4 [29,32]. As concluded by Pang et al. [6], a flow constriction occurs at approximately a 45° angle to the pile, showing good correlation between the two results.
Velocity contour and streamlines on the y-axis plane in front of the pier are given in Figure 5. Upstream of the pier, part of the flow becomes down-flow, which impinges on the bed materials and results in the scour around the pier. The simulations show that the standard k-ε turbulence model is capable of capturing the key flow features, down-flow and flow amplification, which are the main driving force for the sediment transport around the pier and the formation of scour hole. This is consistent with the finding of Roulund et al. [33].

4.1.2. Validation of Scour Pit Model

As known, scour hole is the most direct characterization of local scour, and it can be described by scour depth. Taking the nonprotected pier as an example, the numerical simulation results in this paper are compared with the physical experiment. The overall riverbed surface after equilibrium scour is shown in Figure 6. Obviously, the scour patterns on both sides of the pier are symmetrical, and the sand accumulates at back surface of the pier. Figure 7a,b show the local scour around the pier in the numerical simulation test and physical experiment, respectively. It can be observed that the scour holes are approximately circular, and their shapes are basically the same. The maximum scour depth in front of the pier is greater than that on both sides of the pier. The maximum scour depths of the simulated value and physical experiment value are 3.16 cm and 3.1 cm, respectively, with an error of 1.9%.
In terms of flow field verification, the model established in this paper was compared and verified with the streamline diagrams obtained by Pang et al. [6], and it was found that the flow field patterns were similar. Regarding the scouring pit, by establishing a laboratory flume scouring experiment and comparing it with the scouring pit obtained from the numerical model, the error was found to be around 1.9%. The results indicate that the CFD model in this paper has high reliability and accuracy and can be used for subsequent simulations.

4.2. Analysis of Orthogonal Experimental Results

The maximum scour depth can directly reflect the anti-scour effect of the combined device. Therefore, it has a high reliability to be the evaluation index of the orthogonal experiment. The output responses from the orthogonal experimental tests are recorded in Table 3. In order to show the change rule of the index (scour depth) more clearly when the level of each factor changes, the main-effect plots are depicted.
According to the simulation results in Table 3, Figure 8 shows the reduction rate of maximum scour depth at the front and side end of the pier. It can be seen that under the protection of the combined device, the scour depth is greatly reduced. More specifically, the reduction rate of the maximum scour depth in front of the pier is between 65.18% and 81.01%, while that at side of the pier is between 52.63% and 68.42%.
Referring to Table 4, when the maximum scour depth in front of the pier is used as the evaluation index, the variable ranked first is the ratio of perforated area (S); the sequence of the impact of all other variables on the maximum scour depth at the front end of pier can be expressed as H > L. While when the index is the maximum scour depth at the side of the pier, the significance ranks as H > S > L. Although the results in Table 4 show that the factors are arranged differently, L is in the last rank for reducing the scour depth. In other words, the distance from the perforated baffle to the pier has little effect on the reduction of the scour depth.
Furthermore, as shown in Figure 9, when the indexes are different, the values of RS are very different, while those of RH are very close, indicating that the influence of S on the scour depth at front end and side end of the pier is stronger than that of H. On the whole, the significance of the three factors on the scour depth can be ranked as S > H > L.
Grimaldi et al. [34] reached a scour reduction in front of the pier of about 45% with a best combination of downstream bed-sill and slot. Odgaard and Wang [35] obtained 64% and 45% scour depth reduction by using combinations of submerged vanes and collar or pier plates and collar, respectively. Zarrati et al. [36] examined a combination of continuous collar and riprap, and a 50% and 60% scour depth reduction in the vicinity of the front and rear piers was achieved. Compared with the above conclusions, the combined device proposed in this paper has a good anti-scour effect, especially at the front side of the pier. However, it is necessary to emphasize that there are oscillations in the results of different cases of combined devices versus single plate, such as in tests 3 and 10. In other words, the maximum scour depth reduction (52.63%) of these combinations is lower than that of the single ring-wing plate (57.89%). The reason for this phenomenon is that combining two devices may change their combined behavior for better or worse, and the mechanisms of both devices are not mutually reinforcing, which is consistent with the findings of Garg et al. [37]. In summary, it is necessary to obtain the best combination through range analysis.

4.2.1. Range Analysis for Factors

Range analysis is an important analysis technique in orthogonal experiments. The range of a factor is the difference between its average maximum and lowest values at different levels; the larger the value of the range, the more important this factor is. Thus, the primary and secondary relationship of the factors and the best dimensions can be obtained. The range can be expressed as Equation (7), and the results of range analysis are illustrated in Table 4.
R = K m a x K m i n

4.2.2. Influence of Factors on Maximum Scour Depth

In order to obtain the best level of each factor, the influence of different levels of each factor on the maximum scour depth is needed to be further investigated. As indicated in Figure 10, the increase of the perforated ratio will reduce the scour depth at the side of the pier; this is because the larger perforated ratio allows more water to flow through the baffle to the front end of the pier; correspondingly, the scour depth increases. However, there is still a question regarding efficiency when the perforated ratio is lower than 20%. Accordingly, the optimal ratio of the perforated area is 20%.
As shown in Figure 11, the trend of distance between the pier and the perforated baffle shows that, the farther the baffle is, the deeper the scour at front end of the pier is, and the depth at the side of the pier is opposite. It can also be clearly seen from Figure 11 that when the baffle is 2d from the pier, the combined device has the same anti-scour effect on the front and side of the pier. Instead, the longer distance provides the flow with enough space to recover parallel to the incoming flow and becomes down-flow after meeting the pier, and the scour depth in front of the pier is also increased. However, it should be pointed out that increasing the distance between the baffle and the pier has little effect on reducing the scour depth at side of the pier. Accordingly, 2d can be considered as the best installation location for the perforated baffle.
As illustrated in Figure 12, both lines show that the optimal height for the ring-wing plate is 1/3 of the water depth. Increasing H may reduce the effectiveness of this structure. However, there is still a question regarding the efficiency of the ringwing plate when its height is lower than 1/3 of water depth. This finding is consistent with the conclusion of Cheng et al. [19].
Under comprehensive consideration, the best dimensions of a combined device to achieve the maximum reduction of the scour depth are equal to S = 20%, L = 2 d, and H = 1/3 h.

4.2.3. Effect of the Optimal Combined Device on Flow Field

In order to see how the optimized combined device impacts the flow field and reduces the scour depth, the flow field is further studied. Shown in Figure 13 are the streamlines near the bed for the two conditions. And the background of this figure shows the velocity of flow (Va = (u2 + v2)0.5).
It can be clearly seen from Figure 13a that there is separation and acceleration of the water flow on the sides of the pier; more specifically, the velocity on the pier side is 1.2 times the inlet velocity, which is almost the same as the magnification of 1.3 obtained by Pang et al. [4]. On the contrary, with the protection of a combined device, the velocity at both sides of the pier is 0.22 m/s, as shown in Figure 13b. It is also seen that the velocity around the perforated baffle is smaller than the inlet velocity.
To better illustrate the anti-scour mechanism of the combined device, two empirical equations proposed by Melville [29] and Dou Guoren [38] are used to calculate the critical velocity of the sand, given as Equations (19)–(21).
V c U = 5.75 log 5.53 y d 50
U = 0.0115 + 0.012 d 50 0.14 0.1 m m < d 50 < 1 m m
V c = y d 50 0.14 0.000000605 10 + y d 50 0.72 0.5
where V c is the critical velocity of sediment, U is the critical shear velocity, y is the water depth, and d50 is the median particle size of sand. The critical velocity obtained in this paper and calculated by the above equations is equal to 0.34 and 0.31 m/s. According to the results, the flow at both sides of the unprotected pier almost reaches the critical velocity of sediment. As shown in Figure 13b, due to the diversion and shielding effect of the perforated baffle, the velocity of the accelerated region (0.28 m/s) is reduced, and the region is far from the pier. These findings interpret the reduction of local scour at the vicinity of the pier.
Figure 14 shows the streamlines and velocity contour on the vertical plane. As concluded by Aslani-Kordkandi and Beheshti [39] and Keshavarzi et al. [40], Figure 14a shows the streamlines impinging on the sediment, with the maximum Vz = 0.16 m/s. And as expected and shown in Figure 14b, the flow filed upstream of the pier changes significantly after installing the combined device. Although there is still down-flow above the ring-wing plate, due to the shielding effect of the perforated baffle, its vertical velocity is very small, only 0.06 m/s. It is also worth mentioning that the flow under the ring-wing plate is basically parallel to the incoming flow. These changes in the flow field interpret the reduction of local scour at the vicinity of the pier, and together with the other results in this study, very clearly show the anti-scour effect of the combined device.
Based on the above conclusions, it can be deduced that the optimized combined device is able to change the vortex systems, including down-flow and the horseshoe vortex upstream of the pier, and it is considered a very effective countermeasure against scour around the pier. The combined device can be applied for both new and existing bridge piers, especially when the direction of flow is perpendicular to the arrangement of the bridge piers. It should be pointed out that the above conclusions apply exclusively for the flow and sand conditions considered in the present research.

5. Conclusions

This study proposed a new anti-scour device combined with perforated baffle and ring-wing plate. By using ANSYS-FLUENT, with the maximum scour depth at the front and side ends of the pier as evaluation indices, the effect of the ratio of baffle perforated, relative position of baffle, and height of ring-wing plate on local scour was analyzed. The best dimensions of the combined device were recommended by orthogonal testing, and the flow filed around pier under the protection of the optimal combined device was further studied. The results allow us to draw the following conclusions:
(1) The comparisons between the numerical and the experimental results show that the numerical model adopted in this paper is of high accuracy and advantageous for investigating the shape of scour hole and flow field characteristics around the bridge pier.
(2) After installing the combined device, the maximum scour depth reduction rates in front of the pier are between 65.18% and 81.01%, while those at the side of the pier are between 52.63% and 68.42%. However, some combinations (no. 2 and no. 10, with 52.63%) have a low scour reduction, which is less than the scour reduction of the ring-wing plate alone (57.89%). These results indicate that the efficiency of a combined device is not the linear sum of the individual efficiency of each device.
(3) Based on the orthogonal experiments, the significance of the three factors on the scour depth can be ranked as S > H > L. And the best dimensions of a combined device to achieve the maximum reduction of the scour depth are equal to S = 20%, L = 2d, and H = 1/3 h.
(4) The study of the flow field shows that the ring-wing plate blocks most of the down-flow, so the bed is protected from impinging. In addition, on the horizontal plane, the effect of diversion and water energy dissipation provided by the perforated baffle reduce the velocity of the flow below the critical velocity of sediment. With the two protection mechanisms to reinforce each other, the combined device provides excellent performance.

6. Methods

In order to better control the specific parameters of the test and reduce human influence, ANSYS-FLUENT finite element software was mainly used to conduct simulation tests of the device, and laboratory flume tests were used to verify the numerical model to prove its effectiveness. In terms of test parameters, a variety of test parameters were de-signed, and the orthogonal test method was adopted to select the optimal device parameters. Through the finite element software, the protection effect of the device was better presented through the scouring pit morphology and flow field morphology. After obtaining the specific data, the scour reduction rate of the protection device around the bridge pier was calculated and analyzed to evaluate and analyze the scour reduction effect.

Author Contributions

Y.W.: Funding acquisition, Resources, Supervision, Writing—original draft. R.L.: Conceptualization, Methodology, Software, Investigation, Formal analysis, Writing—original draft. P.Y.: Funding acquisition, Resources. J.C.: Data curation. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

This work was supported by the Science and technology projects of the National Key R&D Program of China (2024YFC3015203).

Informed Consent Statement

Not applicable for studies not involving humans.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

Author Pei Yuan was employed by the company China Merchants Chongqing Communications Technology Research & Design Institute Co., Ltd. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Myrhaug, D.; Rue, H. Scour around group of slender vertical piles in random waves. Appl. Ocean Res. 2005, 27, 56–63. [Google Scholar] [CrossRef]
  2. Yuksel, Y.; Tan, R.I.; Celikoglu, Y. Propeller jet flow scour around a pile structure. Appl. Ocean Res. 2018, 79, 160–172. [Google Scholar] [CrossRef]
  3. Matos, J.C.; Nicoletti, V.; Kralovanec, J.; Sousa, H.S.; Gara, F.; Moravcik, M.; Morais, M.J. Comparison of Condition Rating Systems for Bridges in Three European Countries. Appl. Sci. 2023, 13, 12343. [Google Scholar] [CrossRef]
  4. Brighenti, F.; Caspani, V.F.; Costa, G.; Giordano, P.F.; Limongelli, M.P.; Zonta, D. Bridge management systems: A review on current practice in a digitizing world. Eng. Struct. 2024, 321, 118971. [Google Scholar] [CrossRef]
  5. Ayoubloo, M.; Etemad-Shahidi, A.; Mahjoobi, J. Evaluation of regular wave scour around a circular pile using data mining approaches. Appl. Ocean Res. 2010, 32, 34–39. [Google Scholar] [CrossRef]
  6. Pang, A.L.J.; Skote, M.; Lim, S.Y.; Gullman-Strand, J.; Morgan, N. A numerical approach for determining equilibrium scour depth around a mono-pile due to steady currents. Appl. Ocean Res. 2016, 57, 114–124. [Google Scholar] [CrossRef]
  7. Zhan, J.; Wang, Y.; Zhang, F.; An, Y.; Liu, K. Scour depth evaluation of highway bridge piers using vibration measurements and finite element model updating. Eng. Struct. 2022, 253, 113815. [Google Scholar] [CrossRef]
  8. Xiang, Q.; Wei, K.; Li, Y.; Zhang, M.; Qin, S. Experimental and Numerical Investigation of Local Scour for Suspended Square Caisson under Steady Flow. KSCE J. Civ. Eng. 2020, 24, 2682–2693. [Google Scholar] [CrossRef]
  9. Xiang, Q.; Wei, K.; Qiu, F.; Yao, C.; Li, Y. Experimental Study of Local Scour around Caissons under Unidirectional and Tidal Currents. Water 2020, 12, 640. [Google Scholar] [CrossRef]
  10. Tafarojnoruz, A.; Gaudio, R.; Dey, S. Flow-altering countermeasures against scour at bridge piers: A review. J. Hydraul. Res. 2010, 48, 441–452. [Google Scholar] [CrossRef]
  11. Valela, C.; Whittaker, C.N.; Rennie, C.D.; Nistor, I.; Melville, B.W. Novel Riprap Structure for Improved Bridge Pier Scour Protection. J. Hydraul. Eng. 2022, 148. [Google Scholar] [CrossRef]
  12. Ranjbar-Zahedani, M.; Keshavarzi, A.; Khabbaz, H.; Ball, J.E. Optimizing flow diversion structure as an effective pier-scour countermeasure. J. Hydraul. Res. 2021, 59, 963–976. [Google Scholar] [CrossRef]
  13. Fayun, L.; Chen, W. Review on Countermeasures to Bridge Piers from Local Scour. Struct. Eng. 2014, 5, 130–138. [Google Scholar]
  14. Fouli, H.; Elsebaie, I.H. Reducing local scour at bridge piers using an upstream subsidiary triangular pillar. Arab. J. Geosci. 2016, 9, 598. [Google Scholar] [CrossRef]
  15. Wang, S.; Garbatov, Y.; Chen, B.; Soares, C.G. Dynamic structural response of perforated plates subjected to water impact load. Eng. Struct. 2016, 125, 179–190. [Google Scholar] [CrossRef]
  16. Su, T.C.; Chen, B.; Zhang, H.; Hu, T.; Yang, Z.Y. On reducing scouring by controlling junction flow. In Proceedings of the Twenty-third International Offshore and Polar Engineering Conference, Anchorage, Alaska, 30 June–5 July 2013. [Google Scholar]
  17. Zarrati, A.R.; Gholami, H.; Mashahir, M.B. Application of collar to control scouring around rectangular bridge piers. J. Hydraul. Res. 2004, 42, 97–103. [Google Scholar] [CrossRef]
  18. Ghorbani, B.; Kells, J.A. Effect of submerged vanes on the scour occurring at a cylindrical pier. J. Hydraul. Res. 2008, 46, 610–619. [Google Scholar] [CrossRef]
  19. Cheng, L.Y.; Mou, X.Y.; Wen, H.; Hao, L.Z. Experimental Research on Protection of Ring-Wing Pier Against Local Scour. Adv. Sci. Technol. Water Resour. 2012, 3. Available online: https://api.semanticscholar.org/CorpusID:112450695 (accessed on 1 November 2024).
  20. Karami, H.; Hosseinjanzadeh, H.; Hosseini, K.; Ardeshir, A. Scour and three-dimensional flow field measurement around short vertical-wall abutment protected by collar. KSCE J. Civ. Eng. 2018, 22, 141–152. [Google Scholar] [CrossRef]
  21. Park, J.H.; Sok, C.; Park, C.K.; Kim, Y.D. A study on the effects of debris accumulation at sacrificial piles on bridge pier scour:, I. Experimental results. KSCE J. Civ. Eng. 2016, 20, 1546–1551. [Google Scholar] [CrossRef]
  22. Park, J.H.; Sok, C.; Park, C.K.; Kim, Y.D. A study on the effects of debris accumulation at sacrificial piles on bridge pier scour:, I.I. Empirical formula. KSCE J. Civ. Eng. 2016, 20, 1552–1557. [Google Scholar] [CrossRef]
  23. Yagci, O.; Yildirim, I.; Celik, M.F.; Kitsikoudis, V.; Duran, Z.; Kirca, V.O. Clear water scour around a finite array of cylinders. Appl. Ocean Res. 2017, 68, 114–129. [Google Scholar] [CrossRef]
  24. Vittal, N.; Kothyari, U.C.; Haghighat, M. ClearWater Scour around Bridge Pier Group. J. Hydraul. Eng. 1994, 120, 1309–1318. [Google Scholar] [CrossRef]
  25. Mashahir, M.B.; Zarrati, A.R.; Mokallaf, E. Application of Riprap and Collar to Prevent Scouring around Rectangular Bridge Piers. J. Hydraul. Eng. 2010, 136, 183–187. [Google Scholar] [CrossRef]
  26. Moncada-M, A.T.; Aguirre-Pe, J.; Bolivar, J.C.; Flores, E.J. Scour protection of circular bridge piers with collars and slots. J. Hydraul. Res. 2009, 47, 119–126. [Google Scholar] [CrossRef]
  27. Wang, Y.; Shen, X.; Chen, J.; Chen, Z.; Liu, J. Research on the Wing-Type Antiscour Device of Pier Based on Scour Test and Numerical Simulation. Adv. Civ. Eng. 2021, 2021, 2491707. [Google Scholar] [CrossRef]
  28. Melville, C.; Bruce, W.; Coleman, S.E. Bridge Scour; Water Resources Publication, LLC: Highlands Ranch, CO, USA, 2000; ISBN 9781887201186. [Google Scholar]
  29. Bruce, W.M. Pier and Abutment Scour: Integrated Approach. J. Hydraul. Eng. 1997, 123, 125–136. [Google Scholar]
  30. Melville, B.W.; Raudkivi, A.J. Flow characteristics in local scour at bridge piers. J. Hydraul. Res. 2010, 15, 373–380. [Google Scholar] [CrossRef]
  31. Huang, C.K.T.C. Use of surface guide panels as pier scour countermeasures. Int. J. Sediment Res. 2005, 2, 119–130. [Google Scholar]
  32. Lauchlan, C.S. Pier Scour Countermeasures; The University of Auckland: Auckland, New Zealand, 1999. [Google Scholar]
  33. 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]
  34. Grimaldi, C.; Gaudio, R.; Calomino, F.; Cardoso, A.H. Countermeasures against Local Scouring at Bridge Piers: Slot and Combined System of Slot and Bed Sill. J. Hydraul. Eng. 2009, 135, 425–431. [Google Scholar] [CrossRef]
  35. Odgaard, A.J.; Wang, Y. Scour prevention at bridge piers. In Hydraulic Engineering; ASCE: Reston, VA, USA, 1987; pp. 523–527. [Google Scholar]
  36. Zarrati, A.R.; Nazariha, M.; Mashahir, M.B. Reduction of Local Scour in the Vicinity of Bridge Pier Groups Using Collars and Riprap. J. Hydraul. Eng. 2006, 132, 154–162. [Google Scholar] [CrossRef]
  37. Garg, V.; Setia, B.; Verma, D.V.S. Combination of scour protection devices around oblong bridge pier. ISH J. Hydraul. Eng. 2008, 14, 56–68. [Google Scholar] [CrossRef]
  38. Guoren, D. Incipient Motion of Coarse and Fine Sediment. J. Sediment Res. 1999, 6, 1–9. [Google Scholar]
  39. Ataie-Ashtiani, B.; Beheshti, A.A. Experimental Investigation of Clear-Water Local Scour at Pile Groups. J. Hydraul. Eng. 2006, 132, 1100–1104. [Google Scholar] [CrossRef]
  40. Alireza, K. Experimental study of flow structure around two in-line bridge piers. Proc. Inst. Civ. Eng.-Water Manag. 2017, 171, 311–327. [Google Scholar]
Figure 1. Grid element.
Figure 1. Grid element.
Buildings 15 00148 g001
Figure 2. Boundary conditions of CFD model.
Figure 2. Boundary conditions of CFD model.
Buildings 15 00148 g002
Figure 3. Schematic diagram of combined device.
Figure 3. Schematic diagram of combined device.
Buildings 15 00148 g003
Figure 4. Comparisons of flow fields near river bed. (The arrows in the figure represent the direction of the flow lines. The same below).
Figure 4. Comparisons of flow fields near river bed. (The arrows in the figure represent the direction of the flow lines. The same below).
Buildings 15 00148 g004
Figure 5. Vertical Water Flow Streamlines and Flow Velocities in Front of the Pier.
Figure 5. Vertical Water Flow Streamlines and Flow Velocities in Front of the Pier.
Buildings 15 00148 g005
Figure 6. Three-dimensional pattern of scour hole around pier.
Figure 6. Three-dimensional pattern of scour hole around pier.
Buildings 15 00148 g006
Figure 7. Pattern of local scour hole. (a) Local scour hole of numerical simulation; (b) local scour hole of physical experiment.
Figure 7. Pattern of local scour hole. (a) Local scour hole of numerical simulation; (b) local scour hole of physical experiment.
Buildings 15 00148 g007
Figure 8. Maximum scour depth reduction rate.
Figure 8. Maximum scour depth reduction rate.
Buildings 15 00148 g008
Figure 9. Relationship between factors and range.
Figure 9. Relationship between factors and range.
Buildings 15 00148 g009
Figure 10. Influence of S on maximum scour depth.
Figure 10. Influence of S on maximum scour depth.
Buildings 15 00148 g010
Figure 11. Influence of L on maximum scour depth.
Figure 11. Influence of L on maximum scour depth.
Buildings 15 00148 g011
Figure 12. Influence of H on maximum scour depth.
Figure 12. Influence of H on maximum scour depth.
Buildings 15 00148 g012
Figure 13. The streamlines and velocity contour near the bed: (a) single pier, (b) pier with combined device.
Figure 13. The streamlines and velocity contour near the bed: (a) single pier, (b) pier with combined device.
Buildings 15 00148 g013
Figure 14. The streamlines and velocity contour on the vertical plane: (a) single pier, (b) pier with combined device.
Figure 14. The streamlines and velocity contour on the vertical plane: (a) single pier, (b) pier with combined device.
Buildings 15 00148 g014
Table 1. Model parameters for the standard k-ε model.
Table 1. Model parameters for the standard k-ε model.
GuσkσεGε1Gε2
0.091.01.31.441.92
Table 2. Factors and levels of orthogonal experiment.
Table 2. Factors and levels of orthogonal experiment.
LevelsVariables
SLH
110%2 d1/2 h
220%3 d1/3 h
330%4 d1/6 h
Table 3. Schemes and results of orthogonal experiment.
Table 3. Schemes and results of orthogonal experiment.
Test NumberFactors Scour Depth
SLHEmpty ColumnFront End of Pier (cm)Side End of Pier (cm)
1----3.161.9
2--1/3 h-1.20.8
310%2d1/2 h10.80.9
410%3d1/3 h20.90.7
510%4d1/6 h31.10.8
620%2d1/6 h30.80.8
720%3d1/2 h10.60.7
820%4d1/3 h20.60.8
930%2d1/3 h20.70.6
1030%3d1/6 h30.90.9
1130%4d1/2 h10.90.6
Table 4. Results of range analysis.
Table 4. Results of range analysis.
Analysis ResultsMaximum Scour Depth
at Front End of Pier
Maximum Scour Depth
at Side End of Pier
SLHSLH
K10.9330.7670.7670.8330.7670.733
K20.6670.8000.7330.7670.7670.700
K30.8330.8670.9330.7000.7330.867
R0.2660.1000.2000.1330.0340.167
Rank132231
Optimal levelS2L1H2S3L3H2
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

Wang, Y.; Liao, R.; Yuan, P.; Chen, J. Optimization of Anti-Scour Device Combined with Perforated Baffle and Ring-Wing Plate Based on a Multi-Factor Orthogonal Experiment. Buildings 2025, 15, 148. https://doi.org/10.3390/buildings15010148

AMA Style

Wang Y, Liao R, Yuan P, Chen J. Optimization of Anti-Scour Device Combined with Perforated Baffle and Ring-Wing Plate Based on a Multi-Factor Orthogonal Experiment. Buildings. 2025; 15(1):148. https://doi.org/10.3390/buildings15010148

Chicago/Turabian Style

Wang, Yan, Rongjun Liao, Pei Yuan, and Jinchao Chen. 2025. "Optimization of Anti-Scour Device Combined with Perforated Baffle and Ring-Wing Plate Based on a Multi-Factor Orthogonal Experiment" Buildings 15, no. 1: 148. https://doi.org/10.3390/buildings15010148

APA Style

Wang, Y., Liao, R., Yuan, P., & Chen, J. (2025). Optimization of Anti-Scour Device Combined with Perforated Baffle and Ring-Wing Plate Based on a Multi-Factor Orthogonal Experiment. Buildings, 15(1), 148. https://doi.org/10.3390/buildings15010148

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