Research article Special Issues

Laboratory and numerical investigation of the 2-array submerged vanes in meandering open channel


  • In the case of flooding in rivers, river regulation structures are important since scours occur on the outer meander due to high flow velocities. In this study, 2-array submerged vane structures were investigated which is a new method in the meandering part of open channels, both laboratory and numerically with an open channel flow discharge of 20 L/s. Open channel flow experiments were carried out by using a submerged vane and without a vane. The flow velocity results of the computational fluid dynamics (CFD) models were compared to the experimental results and the results were found compatible. The flow velocities were investigated along with depth using the CFD and found that the maximum velocity was reduced by 22–27% along the depth. In the outer meander, the 2-array submerged vane with a 6-vane structure was found to affect the flow velocity by 26–29% in the region behind the vane.

    Citation: Bestami TAŞAR, Fatih ÜNEŞ, Ercan GEMİCİ. Laboratory and numerical investigation of the 2-array submerged vanes in meandering open channel[J]. Mathematical Biosciences and Engineering, 2023, 20(2): 3261-3281. doi: 10.3934/mbe.2023153

    Related Papers:

    [1] Huu Thuan Nguyen, Tu Anh Do, Benoît Cosson . Numerical simulation of submerged flow bridge scour under dam-break flow using multi-phase SPH method. Mathematical Biosciences and Engineering, 2019, 16(5): 5395-5418. doi: 10.3934/mbe.2019269
    [2] Aslak Tveito, Glenn T. Lines, Pan Li, Andrew McCulloch . Defining candidate drug characteristics for Long-QT (LQT3) syndrome. Mathematical Biosciences and Engineering, 2011, 8(3): 861-873. doi: 10.3934/mbe.2011.8.861
    [3] Chii-Dong Ho, Jr-Wei Tu, Hsuan Chang, Li-Pang Lin, Thiam Leng Chew . Optimizing thermal efficiencies of power-law fluids in double-pass concentric circular heat exchangers with sinusoidal wall fluxes. Mathematical Biosciences and Engineering, 2022, 19(9): 8648-8670. doi: 10.3934/mbe.2022401
    [4] Peter W. Bates, Jianing Chen, Mingji Zhang . Dynamics of ionic flows via Poisson-Nernst-Planck systems with local hard-sphere potentials: Competition between cations. Mathematical Biosciences and Engineering, 2020, 17(4): 3736-3766. doi: 10.3934/mbe.2020210
    [5] Virginia González-Vélez, Amparo Gil, Iván Quesada . Minimal state models for ionic channels involved in glucagon secretion. Mathematical Biosciences and Engineering, 2010, 7(4): 793-807. doi: 10.3934/mbe.2010.7.793
    [6] Ziqiang Cheng, Jin Wang . Modeling epidemic flow with fluid dynamics. Mathematical Biosciences and Engineering, 2022, 19(8): 8334-8360. doi: 10.3934/mbe.2022388
    [7] Alexander N. Ostrikov, Abdymanap A. Ospanov, Vitaly N. Vasilenko, Nurzhan Zh. Muslimov, Aigul K. Timurbekova, Gulnara B. Jumabekova . Melt flow of biopolymer through the cavities of an extruder die: Mathematical modelling. Mathematical Biosciences and Engineering, 2019, 16(4): 2875-2905. doi: 10.3934/mbe.2019142
    [8] Xiangqun Li, Hu Chen, Dong Zheng, Xinzheng Xu . CED-Net: A more effective DenseNet model with channel enhancement. Mathematical Biosciences and Engineering, 2022, 19(12): 12232-12246. doi: 10.3934/mbe.2022569
    [9] Hamid Mofidi . New insights into the effects of small permanent charge on ionic flows: A higher order analysis. Mathematical Biosciences and Engineering, 2024, 21(5): 6042-6076. doi: 10.3934/mbe.2024266
    [10] Yi Wang, Shicheng Zhong, Guo Wang . Dynamic selection of clarification channels in rumor propagation containment. Mathematical Biosciences and Engineering, 2023, 20(8): 14995-15017. doi: 10.3934/mbe.2023672
  • In the case of flooding in rivers, river regulation structures are important since scours occur on the outer meander due to high flow velocities. In this study, 2-array submerged vane structures were investigated which is a new method in the meandering part of open channels, both laboratory and numerically with an open channel flow discharge of 20 L/s. Open channel flow experiments were carried out by using a submerged vane and without a vane. The flow velocity results of the computational fluid dynamics (CFD) models were compared to the experimental results and the results were found compatible. The flow velocities were investigated along with depth using the CFD and found that the maximum velocity was reduced by 22–27% along the depth. In the outer meander, the 2-array submerged vane with a 6-vane structure was found to affect the flow velocity by 26–29% in the region behind the vane.



    In the meander section of the rivers, secondary flow develops when flow crosses the tip due to a combination of transverse pressure gradient and centrifugal force [1]. The outer meander of the river is subject to erosion due to the high flow velocity, while the low velocity of the flow near the bed towards the inner shore causes accumulation [2]. Also, induced velocity field variations are the reason for shear stress changes that cause erosion and sediment transport. [3]. In order to reduce these effects of erosion and sediment, studies with a submerged vane on meander curves by Odgaard and Kennedy [4,5] led to the introduction of submerged vane into the literature as a new method. Ghorbani and Kells [6] examined the effect of a submerged vane on local scour in a cylindrical leg. Davoodi and Bejestan [7] experimented with four different longitudinal distances between submerged vane, equal to 4, 6, 8, and 10 H, under four different flow conditions. Gemici [8] investigated the effects of different Froude numbers, vane arrangement, vane length/vane height, and approach angle combinations on flow velocity. Mohammadiun et al. [9] conducted their research to reduce sedimentation and erosion potential at the junction of curved and straight channels. Fathi and Zomorodian [10] presented the results of laboratory experiments investigating how submerged vanes affect scrubbing around a vertical wall and overflow abutments under clean water conditions. Kalathil et al. [11] used submerged vanes to control sediment entry into the inlet channel with a physical model. Zarei et al. [12] investigated the effects of submerged vanes around the bridge pier and on bed topography changes. Karami Moghadam [13] used a 55° intake channel branched from a rectangular channel for shear stress effect under submerged vanes. Azizi and Bajestan [14] investigated the rectangular vane and five other modified vanes with tapered leading edges with angles of θ = 30°, 45°, 60°, 70°, and 73.3°. Baltazar et al. [15] examined the effect of a submerged vane field on the flow pattern and bed morphology near and inside the entrance reach of a movable bed 90° lateral diversion. Lake et al. [16] investigated the effect of river-based, submerged vane structures in Australia. When the previous studies on the submerged vane are examined, especially the models made with computational fluid dynamics (CFD) are limited. It is difficult to analyze while the basic equations of fluid mechanics were applied to parameters such as complex geometries, viscosity, flow velocity, and temperature. CFD could be applied to the solution of problems in fluid mechanics [17]. In the past, some researchers have used 3D numerical models to determine flow characteristics [18,19,20,21,22,23,24].

    When the past studies were examined, it has been determined that there were very few or limited studies to reduce/prevent scours caused by high flow velocities in the meandering of the rivers. This study focuses on the effect of the submerged vane, which is accepted as a new method [25], on the meandering open channel. For this purpose, experiments were carried out where in the meandering part of the open channel which is of 2-array with 6 submerged vanes, and without a vane using a channel discharge of 20 L/s. The submerged vane experiments conducted in this study will contribute to the literature by researching the originality of the river meanders and the effect of the vane. In addition, flow velocities in the open channel were modeled with the CFD method and the cross-sectional flow velocity changes were investigated along the depth.

    The performance of no vane (V0 model) and 2-array with 6 submerged vanes (V1) models in meandering open channel flows has been investigated by using a flow discharge of 20 L/s in this study. Detailed information about the experimental setup, experimental conditions, and mathematical model for the V0 and V1 experiments conducted in the open channel are given below.

    At Bartın University hydro-mechanic laboratory, a 30 cm wide curvilinear channel was built within a 50 cm wide rectangular channel inside the existing open channel system. The initial conditions for the experiment are given in Table 1. Q, B, d, A, T, Rh, Fr, and Vmean denote the discharge flow, channel width, depth of flow, wetted area, wetted perimeter, hydraulic radius, flow Froude number and inlet mean flow velocity in Table 1. The subcritical flow has occurred in the open channel where close to the energy-reducing tank and the upstream because the Froude number is less than 1.

    Table 1.  The initial conditions for the experiment.
    Q (m3/s) B (m) d (m) A (m2) T (m) Rh (m) Vmean (m/s) Fr
    0.020 0.300 0.085 0.026 0.470 0.054 0.785 0.860

     | Show Table
    DownLoad: CSV

    In order to reduce the turbulence of the flow of water entering the channel with the pump, the entering water is taken to the settling tank and the energy of the entering water is reduced. The open channel assembly in which the submerged vane experiment was performed had a base slope of 0.0003. A meandering channel was constructed along the 50 × 50 cm sectioned rectangular channel with approximately 3 meters long and 30 cm wide. The curved 30° meander is connected between upstream and downstream rectangular channels. The meandering channel has a center radius R = 3.60 m. The submerged vane is constructed using 2 mm thick sheet metal. In order to determine the flow discharge, the flowmeter ultrasonic device was used at the entrance to the open channel system from the water tank. On the bottom of the submerged vane, 5 mm thick sheet of material was mounted in accordance with the slope/dimensions of the meandering channel and was removed and installed during each V0, V1 experiment. These dimensions have also been considered in the numerical simulations.

    The flow discharges were measured by an Acoustic Doppler Velocimetry (ADV) flowmeter. Measurements were made at 60% of the depth (0.6 d). Discharge flow was recorded with an ultrasonic flow meter. The experimental setup and all sections were shown in Figure 1.

    Figure 1.  Submerged vane experimental set-up for V1 case: a) Channel top (plan) view, b) Cross section view, c) Section view.

    For the flow occurrence in river conditions, the bottom of the channel was laid with 15 cm of sediment (gravel) material and the base of the vane was placed on the sand of 5 cm. The experiment conditions prepared for the V1 case were also provided for the V0 case. With the start of the pump, the water drawn from the water tank passed through the gravel filters and reached the flow conditions to stabilize. It took approximately 15 minutes for the flow conditions to stabilize. After flow conditions were stabilized, flow discharges were recorded and flow discharge adjustments were made with an ultrasonic flow meter.

    During the design phase, the boundary conditions and flow properties of the experiment were defined using the Gambit program. CFD was applied with the Fluent software after the design. Flow velocity simulations at the measurement points of the open channel were examined as experimental and CFD results separately. Finally, the experimental results of the V1 case were analyzed based on hydraulic terminology and its comparative effect was investigated with the V0 case.

    For the V1 case, measurements were taken in the middle of the 10 cm gap after each vane (5 cm) (see Figure 2). Flow velocity measurements were taken perpendicular to the flow at 2.5 cm intervals at 8 post-vane measurement points. The section measured after the vane in the first row was called the 1-1 section. The section after the 6th vane was called the 3-3 section. For the V1 case, the experimental measurement points are given in Figure 2. The velocity changes were investigated in 2.5 cm intervals in each section. Velocity variations are discussed in V0 and V1 cases.

    Figure 2.  According to the V1 case, distances (cm) of all sections and points measured in the channel.

    The flow conditions of modeling, are turbulent flow and free surface flow which is directed by gravity. According to the literature, this kind of free surface flow can be simulated by the volume of fluid (VOF) method as water-air 2 phase flow problems. The standard k-ε turbulence model was used in the three-dimensional numerical simulation of flow. The k-ε turbulence model was a possible method for the free surface flow simulations [26,27].

    The investigated open channel flow was a 3D, turbulent, unsteady free surface flow. For the k-ε turbulence model given by Launder and Spalding [28], the continuity equation, the momentum equation, and the equations for k-ε were given Eqs (1)–(6).

    Continuity equation:

    ρt+ρuixi=0 (1)

    Momentum equation:

    ρuit+xj(ρuiuj)=Pxi+xj[(μ+μt)(uixj+ujxi)] (2)

    Turbulence kinetic energy (k) equation:

    ρkt+ρuikxi=xi[(μ+μtσk)kxi]+Gρε (3)

    Turbulence dissipation rate energy (ε) equation:

    (ρε)t+(ρuiε)xi=xi[(μ+μtσε)εxi]+C1εεkGC2ερε2k (4)

    where, t is the time; ui is the velocity components; xi is the coordinate components; ρ is the density; µ is the molecular viscosity coefficient; P is the correct pressure; µt is the turbulent viscosity coefficient, which can be derived from the turbulent kinetic energy k and turbulent dissipation rates:

    μt=ρCμk2ε (5)
    G=μt(μixj+ujxi)uixj (6)

    σk and σε are turbulence Prandtl numbers for k and ε equation respectively, σk = 1.0, σε = 1.3, C and C are ε equation constants, C = 1.44, C = 1.92, Cμ = 0.09 is a constant, determined by experimentally.

    VOF model

    In this study, the VOF method was used to calculate the water-air interface. This method was used for 2-phase air-water flow simulation [29] and to compute the free surface of the flow [30]. The VOF method essentially determines whether the element volumes in the computational mesh are empty, partially filled, or completely filled with water. Representing the volumetric filling ratio of the mesh elements, the mesh element is fully filled for F = 1, empty (filled with air) for F = 0, and partially filled with water for 0 < F < 1 [31]. In this approach, the tracking interface between air and water is accomplished by the solution of a continuity equation for the volume fraction of water:

    αwt+αwuixi=0 (7)

    where, αw is volume fraction of water. In each cell, the sum of the volume fractions of air and water is unity. Volume fractions of air denote αa can be given as

    αw+αa=1;0αw1 (8)

    Flow field boundary conditions must be specified differently on walls, channel inlet, outlet and free surface; due to unsteady and channel flow. Since the boundary and inlet conditions are different for each variable, they must be selected separately. Velocities on free surfaces and at side-bottom channels, velocities on submerged vane faces were obtained using the standard wall function based on the recommendation of Launder and Spalding [28]. This wall function accepts a log-law velocity profile close to the wall and was determined as in Eq (9):

    upu=1Kln(Euypv) (9)

    where, "up" is the average stream flow velocity at the "p" point; "K" von Karman constant (0.418); "yp" is the distance from point p to the wall; empirical constant "E" has the value of 9.79; "u*" is the friction velocity. The "u" uniform velocity distribution was given to the horizontal velocity component in the x-direction at the inflow boundary. The vertical velocity component "v" in the y-direction was set to zero. The inlet velocity field to the channel consists of a forward 'u' horizontal velocity and zero 'v' vertical velocities at all points except points close to the channel.

    The wall y+ is a dimensionless distance similar to the local Reynolds number often used in CFD to indicate how the mesh is for a particular flow. It determines whether the effects in cells adjacent to the wall are laminar or turbulent [32].

    y+=uτyν (10)
    uτ=τwρ (11)

    where uτ is the friction velocity, y is the height from the wall to the mid-point of the wall-adjacent cells, v is the kinematic viscosity, τw is the wall shear stress and ρ is the fluid density at the wall. Values of y+ close to the lower bound (y+ ≈ 30) are most desirable for wall functions, whereas values of y+ ≈ 1 are better for near-wall modelling [33].

    In this paper, the 3D analysis was built up for V0 and V1 structures. CFD model was created according to the open channel experiment (Figures 1 and 2). Inlet conditions and boundary layer conditions were established according to the experimental study. 1,280,602 meshes for V0 submerged vane case and for the V1 case, 1,293,691 meshes were assigned. Also, tetrahedron-type meshes were used in the design. The mesh for the design of V1 case is given in Figure 3.

    Figure 3.  Mesh in the three-dimensional submerged vane experiment.

    The proposed model was used to simulate the variation of submerged vane open channel flow. The biggest problem during the implementation of the submerged vane in the Fluent model was the skewness error of the network grid. The mesh qualities were controlled with equiangular skewness, equisize skewness and aspect ratio in the Gambit program. Skewness and other problems were solved by getting a smaller mesh dimension grid line. Since the real shape of the vane was considered in the model, the skewness error was increased due to the shape of the open channel. Those errors were defined and corrected by increasing the amount of mesh with Gambit. However, when the mesh gets smaller, the number of mesh was increased. Furthermore, the computational domains for all cases were approximately divided 1,280,602–1,293,691 tetrahedral mesh volumes to obtain acceptable skewness coefficients, as seen in Figure 3. This represents approximately 235,000 nodes. In the meshing cases, the convergence criterion of flow parameters was taken at 0.01 for all runs.

    The 3D analysis was built up for V0 and V1 cases. The CFD model was created according to the open channel experiment setup consisting of 1 main meander channel (Figures 1 and 2). According to the experimental inlet measure (Table 1), inlet conditions and boundary layer conditions were established. The boundary conditions of the flow formed with V0 and V1 cases have been defined. Figure 4a), b) show the surfaces for the CFD model solution. According to Figure 4, the water inlet height = velocity inlet, submerged vane - open channel surfaces = wall and outlet section (in downstream) = outflow were accepted. Also, for the flow discharge of 20 L/s, the water inlet height was defined as 5.5 cm.

    Figure 4.  Boundary conditions for V1 case: a) downstream view; b) top view).

    The mesh independence test was performed by doubling the mesh numbers in the lateral and vertical directions. However, no significant difference in flow velocity results was observed due to mesh improvement. This finding showed that the flow velocity results were independent of the mesh.

    When the velocity fields into the open channel have complex flows such as circulation flow, turbulent flow and secondary flow problems arise because the open channel flows are nonlinear and velocity—the pressure field interdependent. These problems were solved using the "Coupled" procedure approach. This procedure was the iteration method and it was based on the prediction-corrector approach. Fluent provides the option to choose "Coupled" pressure-velocity coupling algorithms. The full implicit scheme was used in the model, since the flow was unstable and the analysis time is very long [19]. As an initial state, the inlet channel is first filled with air and water. Then, water at a certain height was released into the free flow at a certain rate from the inlet channel to the upstream end. The channel flow continues until it reaches the downstream limit. The calculation continued for about 600 seconds, at which point the front had already crossed the downstream boundary and any change in flow area was ignored. A time step of 1 second was chosen after preliminary trials to have the desired converged solution. Discretization methods and solver settings are presented in Table 2.

    Table 2.  Numerical model details.
    Solver set Solver Pressure based
    Space-Time 3D, unsteady
    Model Multiphase Model VOF
    Viscous Model k-ε
    Phase Primary Phase Air
    Secondary Phase water
    Discretization Pressure Presto
    Momentum Second order upwind
    Pressure-Velocity Coupling Method Coupled
    Convergence Criterion Residuals 0.001 (Continuity)
    0.01 (Momentum)

     | Show Table
    DownLoad: CSV

    The flow velocities were measured in the open channel for the V0 and V1 cases and it was simulated CFD. Then, the experimental and numerical results were compared for V0 and V1 cases. In order to see the velocity change after the section, the measured and simulated flow velocity results were given as dimensionless based on the average velocity.

    CFD results are given with the measured points according to the 1-1, 2-2, and 3-3 sections as seen in Table 3, Figures 5 and 6. It was observed that high flow velocities are seen at inner points where the channel meandering first starts (Figure 5a), b)). The flow velocity increased from the inner meander to the outer meander. In Figure 5a), b), it was observed that the flow velocity around the vane reduced and between the vanes it had a flow velocity of 0.50–0.75 m/s. It has been seen that flow has a velocity between 0.75–0.90 m/s, near the middle outer meander region and in channel parts without vane. In order to see the changes at the 0.6 d, it was given as a velocity contour in Figure 5c), d) with V0 and V1 cases. As it was presented in the figures, it was observed that the flow was directed around the vane and the flow velocity between the vanes decreased. It was seen that the vane structures were effective in directing the flow and reducing the flow velocity, and the flow reaches equilibrium in the inner meander and outer meander in the channel. Figure 6 shows the shear stress variation for V0 and V1. It has been determined that shear stress decreases behind and around the submerged vane.

    Table 3.  CFD and experimental dimensionless velocity results for V0 case in Sections 1-1, 2-2 and 3-3.
    Section No. Point No. Experiment results Fluent CFD results Error (%)
    1-1 1 0.96 0.97 0.62
    2 0.99 0.97 2.18
    3 1.00 0.97 3.24
    4 1.01 0.97 4.42
    5 1.01 0.98 2.94
    6 1.02 1.00 2.02
    7 0.99 1.06 6.68
    8 1.01 1.09 7.59
    2-2 1 0.98 0.97 1.31
    2 0.98 0.97 0.56
    3 0.97 0.97 0.67
    4 0.96 0.97 1.40
    5 0.97 0.98 1.23
    6 0.96 1.00 4.78
    7 1.10 1.05 3.99
    8 1.10 1.09 1.38
    3-3 1 0.98 0.98 0.21
    2 0.98 0.98 0.57
    3 0.98 0.98 0.35
    4 0.99 0.98 1.26
    5 1.01 0.98 2.96
    6 1.01 1.00 1.11
    7 1.00 1.04 4.35
    8 1.05 1.06 0.98

     | Show Table
    DownLoad: CSV
    Figure 5.  Velocity vectorial and contour representations at 0.6 d for, respectively: a) V0 case (vectorial), b) V1 case (vectorial), c) V0 case (contour), d) V1 case (contour).
    Figure 6.  Wall shear at channel base: a) V0 case) b) V1 case.

    The flow velocity at sections was accepted as dimensionless velocity [34] "u/umean". The 1st point represents the point close to the inner meander, and the 8th point represents the point close to the outer meander. According to the velocity inlet; 1, 1 meters (x1), 1, 5 meters (x2) and 2, 1 meters (x3) were accepted as a 1-1, 2-2 and 3-3 sections, respectively (see Figure 1). The experimental and numerical model dimensionless velocity values obtained in all sections were shown in Figure 7. Table 3 shows the errors by comparing the experimental and CFD analysis results. In Figure 8a)c), velocity contours of the 1-1, 2-2 and 3-3 cross-sections for V0 case were given, respectively. Table 3, Figures 7 and 8 were examined, and it was seen that the error percentage of the velocities was less at the midpoints than at the edges. Maximum error was observed in the outer meander (7–8th points) for V0 case 1-1 section (6–8%). In general, it was found that the experiment and Fluent-CFD results were in good agreement. According to Figure 8a)c), it was seen that the velocity increased in the outer meander and decreased in the inner meander. Also, as can be seen in Figure 8, the water depth decreased in the inner meander, while the water depth increases in the outer meander [34] (the water depth changes with the blue contour lines). For the V0 case, it was seen that the experiment and CFD results were generally compatible and the errors were at a reasonable level (see Table 3). The experimental and numerical model (V1 case) velocity values obtained in all sections were shown in Figure 9. Table 4 shows the errors by comparing the experimental and CFD analysis results. In Figure 10a)c), velocity contours of the 1-1, 2-2 and 3-3 cross-sections (V1 case) were given, respectively.

    Figure 7.  Dimensionless flow velocity variation for V0 case at: a) Section 1-1, b) Section 2-2 and c) Section 3-3.
    Figure 8.  Velocity value (m/s) contours obtained as a result of V0 case CFD analysis for: a) Section 1-1, b) Section 2-2 and c) Section 3-3.
    Figure 9.  Dimensionless flow velocity variation for V1 case at: a) Section 1-1, b) Section 2-2 and c) Section 3-3.
    Table 4.  CFD and experimental dimensionless velocity results for V1 case at Sections 1-1, 2-2 and 3-3.
    Section No. Point No. Experiment results Fluent CFD results Error
    (%)
    1-1 1 1.01 1.06 4.62
    2 1.00 1.05 5.78
    3 1.02 1.05 2.65
    4 1.05 1.00 4.48
    5 0.88 0.82 7.40
    6 1.04 1.02 1.06
    7 0.91 0.88 2.96
    8 1.10 1.12 1.68
    2-2 1 1.13 1.11 2.36
    2 1.13 1.10 2.62
    3 1.02 1.10 7.83
    4 0.91 0.99 9.23
    5 0.81 0.78 2.99
    6 1.09 0.99 9.03
    7 0.88 0.81 7.67
    8 1.05 1.13 7.86
    3-3 1 1.09 1.10 0.98
    2 1.07 1.10 3.17
    3 1.06 1.06 0.50
    4 1.05 1.03 1.31
    5 0.89 0.81 8.22
    6 1.01 0.94 6.20
    7 0.83 0.84 0.96
    8 1.01 1.11 8.99

     | Show Table
    DownLoad: CSV
    Figure 10.  Velocity value (m/s) contours obtained as a result of CFD V1 case analysis. a) Section 1-1, b) Section 2-2 and c) Section 3-3.

    Table 4, Figures 9 and 10 were examined, it was seen that the error percentage of the velocities is less at the midpoints than at the edges. Maximum error was observed in 4th point) for V1 case 2-2 section (9%). In general, it was found that the experiment and Fluent-CFD results are in good agreement.

    According to Figures 9 and 10, decreases in flow velocities were seen behind the vane. Compared to the 1-1 section without the vane (Figures 7a) and 8a)), it was understood that the flow velocity decreased from 0.759 to 0.590 m/s (5. point). It has been determined that there was a nearly 26% decrease in flow velocity. Compared to the 2-2 section without the vane (Figures 7b) and 8b)), it was observed that the flow velocity decreased from 0.820 to 0.583 m/s (7. point). It has been extracted that there is a nearly 29% decreased in flow velocity. Compared to the 3-3 section without the vane (Figures 7c) and 8c)), it was observed that the flow velocity decreased from 0.829 to 0.697 m/s (7. point). It has been determined that there was a nearly 16% decreased in flow velocity. Also, as can be seen in Figure 10a)c), the flow water depth in the inner meander and the water depth in the outer meander were balanced (the water depth changes with the up blue contour lines).

    In Figure 11, CFD -velocity variations across water depth in V0 and V1 cases with different measuring points. Here, the 5th and 7th points represent the back of the vane, and the 3rd point represents the V0 case inner meander. Figure 11a)c) are the velocity changes depending on the depth at the 3rd, 5th and 7th points at 1.1, 1.5 and 2.1 meters for V0 case. Figure 11d)f) at the 3rd, 5th and 7th points at 1.1, 1.5 and 2.1 meters for V1 case.

    Figure 11.  CFD-Velocity changes (m/s) across water depth (m) in V0 case points at a) 1-1, b) 2-2, c) 3-3 sections and V1 case points at d) 1-1, e) 2-2, f) 3-3 sections.

    In Figure 11, it was observed that while the velocities were approximately 0.8 m/s in the vaneless outer meander, these velocities decreased to an average of 0.6 m/s in the vaneless outer meander. For case V1, all sections were after the vane (5 cm) in the direction perpendicular to the flow. According to the after-vane structure's measurements, the submerged vanes reduced flow velocity by 30% in section 1-1, and decreased the flow velocity by 38% in sections 2-2 and 3-3.

    According to the V0 and V1 case, while the velocity decreased on the inner meander, velocity increased on the outer meander. When the results of the experiments were examined in accordance with the CFD simulation results, a 12% error has been detected. Considering that most of the flows in the rivers are turbulent, an error of close to 10% is considered an acceptable error.

    Flow velocity profiles made without a submerged vane are compatible with numerical model or experiment results [34,35]. Reported laboratory experiments show flow patterns of submerged vanes. Scour in the outer meander was caused by the increase in flow velocity and shear stress. The submerged vane cuts off secondary circulation or cross flow cuts off thereby reducing high velocities around the outer meander (Figures 10 and 12).

    Figure 12.  Velocity (m/s) changes for V0 and V1 cases: section of a) 1-1; b) 2-2; c) 3-3.

    In the meander section of the open channel, velocity distribution diagrams of the submerged vaneless streamflow bed (Figures 7 and 8) show that it slides out of the meander and reaches the maximum level near the outer meander before the outlet of the meander. This experimental pattern was consistent with experimental measurements in the other channel meander [34]. The centrifugal force increases proportionally with the square of the flow velocity and the upper region with high flow velocity reaches more centrifugal force than the lower region [35]. Experimental flow velocity results measured at 0.6 d, modeled both with submerged vane and without submerged vane, showed that the model with submerged vane decreased the flow velocity by 26 and 29% and directed the flow (Figures 79). When the literature studies were examined, Gemici [8] experimentally found that submerged vane flow velocities were reduced by 7–19%. The flow models obtained with the 3D model were examined along with the depth (Figure 11), and it was observed that the maximum velocity was reduced by 22–27% (Figure 12). The result of this study will be useful for understanding the flow phenomenon, validation of numerical work and subsequent design of meander protection measures.

    In this study, the performance of V0 and V1 cases in open channel flows was investigated by using a flow discharge of 20 L/s. The effect of the vane structure on the flow velocity measured in the open channel was determined and the experiments performed were modeled by CFD. Dimensionless average velocity values were used to compare the measured and simulated velocity values. It was seen that the experimental and CFD flow velocity results were compatible in V0 and V1 cases, and the CFD model showed a low error for the dimensionless mean velocity. In addition, it is thought that this numerical model can be used in the next stages, in the future projection phase with simulation.

    According to the flow velocity results, it was determined that the vanes significantly reduced the flow velocity, and the V1 case affected the channel flow velocity by 26–29%. The flow results with the CFD were examined along the depth, it was seen that the maximum velocity decreased by 22–27%. Submerged vane structures balanced the water depth between the inner and outer meander open channel flows. The flow velocity was reduced by flow directed from the outer meander to the inner meander. Submerged vane structures were found to be effective in regulating flow. In future studies, the effect of straight/curve, angle and dimensions of the submerged vane can be investigated. The applicability of submerged vane structures in real river meanders and especially the submerged vane effect in flood situations can be investigated. It is also recommended that submerged vane structures can be applied as an effective method of reducing flow velocities and directing flows. It has been known that submerged vane contributes to the protection of the environment by regulating the flow in the inner and outer meanders.

    The authors declare there is no conflict of interest.



    [1] S. Chamran, G. A. Barani, M. S. Sardo, Experimental investigation of submerged vanes shape effect on river-bend stability, J. Hydraul. Struct., 1 (2013), 37–43. https://doi.org/10.22055/JHS.2013.10073 doi: 10.22055/JHS.2013.10073
    [2] A. Voisin, R. D. Townsend, Model testing of submerged vanes in strongly curved narrow channel bends, Can. J. Civil Eng., 29 (2002), 37–49. https://doi.org/10.1139/L01-078 doi: 10.1139/l01-078
    [3] K. Blanckaert, W. H. Graf, Momentum transport in sharp open-channel bends, J. Hydraul. Eng., 130 (2004), 186–198. https://doi.org/10.1061/(ASCE)0733-9429(2004)130:3(186) doi: 10.1061/(ASCE)0733-9429(2004)130:3(186)
    [4] A. J. Odgaard, J. F. Kennedy, River‐bend bank protection by submerged vanes, J. Hydraul. Eng., 109 (1983), 1161–1173. https://doi.org/10.1061/(ASCE)0733-9429(1983)109:8(1161) doi: 10.1061/(ASCE)0733-9429(1983)109:8(1161)
    [5] A. J. Odgaard, J. F. Kennedy, Analysis of Sacramento River Bend Flows, and Development of a New Method for Bank Protection, Iowa, 1982.
    [6] B. Ghorbani, J. A. Kells, Effect of submerged vanes on the scour occurring at a cylindrical pier, J. Hydraul. Res., 46 (2010), 610–619. https://doi.org/10.3826/JHR.2008.3003 doi: 10.3826/JHR.2008.3003
    [7] L. Davoodi, M. S. Bejestan, International control of sediment entry to intake on a trapezoidal channel by submerged vane, Ecol. Environ. Conserv., 18 (2012), 165–169.
    [8] E. Gemici, Flow management with submerged vanes in open channels, Erciyes University, Graduate School of Natural and Applied Sciences, Ph.D. Thesis, 2015.
    [9] S. Mohammadiun, S. A. A. Salehi Neyshabouri, G. Naser, H. Vahabi, Numerical investigation of submerged vane effects on flow pattern in a 90° junction of straight and bend open channels, Iran. J. Sci. Technol., Trans. Civil Eng., 40 (2016), 349–365. https://doi.org/10.1007/s40996-016-0039-7 doi: 10.1007/s40996-016-0039-7
    [10] A. Fathi, S. M. A. Zomorodian, Effect of submerged vanes on scour around a bridge abutment, KSCE J. Civil Eng., 22 (2018), 2281–2289. https://doi.org/10.1007/s12205-017-1453-5 doi: 10.1007/s12205-017-1453-5
    [11] S. T. Kalathil, A. Wuppukondur, R. K. Balakrishnan, V. Chandra, Control of sediment inflow into a trapezoidal intake canal using submerged vanes, J. Waterw. Port Coastal Ocean Eng., 144 (2018), 04018020. https://doi.org/10.1061/(ASCE)WW.1943-5460.0000474 doi: 10.1061/(ASCE)WW.1943-5460.0000474
    [12] E. Zarei, M. Vaghefi, S. S. Hashemi, Bed topography variations in bend by simultaneous installation of submerged vanes and single bridge pier, Arabian J. Geosci., 12 (2019), 1–10. https://doi.org/10.1007/S12517-019-4342-Z/TABLES/2 doi: 10.1007/s12517-018-4128-8
    [13] M. Karami Moghadam, A. Amini, A. Keshavarzi, Intake design attributes and submerged vanes effects on sedimentation and shear stress, Water Environ. J., 34 (2020), 374–380. https://doi.org/10.1111/WEJ.12471 doi: 10.1111/wej.12471
    [14] R. Azizi, S. Bajestan, Iranian hydraulic association journal of hydraulics performance evaluation of submerged vanes by flow-3D numerical model, J. Hydraul., 15 (2020), 1–11. https://doi.org/10.30482/JHYD.2020.105497 doi: 10.30482/JHYD.2020.105497
    [15] J. Baltazar, E. Alves, G. Bombar, A. H. Cardoso, Effect of a submerged vane-field on the flow pattern of a movable bed channel with a 90 lateral diversion, Water, 13 (2021), 828. https://doi.org/10.3390/w13060828 doi: 10.3390/w13060828
    [16] R. W. Lake, S. Shaeri, S. T. M. L. D. Senevirathna, Design of submerged vane matrices to accompany a river intake in Australia, J. Environ. Eng. Sci., 16 (2021), 58–65. https://doi.org/10.1680/jenes.19.00037 doi: 10.1680/jenes.19.00037
    [17] R. R. W. Affonso, R. S. F. Dam, W. L. Salgado, A. X. da Silva, C. M. Salgado, Flow regime and volume fraction identification using nuclear techniques, artificial neural networks and computational fluid dynamics, Appl. Radiat. Isot., 159 (2020), 109103. https://doi.org/10.1016/J.APRADISO.2020.109103 doi: 10.1016/j.apradiso.2020.109103
    [18] P. P. Gadge, V. Jothiprakash, V. V. Bhosekar, Hydraulic investigation and design of roof profile of an orifice spillway using experimental and numerical models, J. Appl. Water Eng. Res., 6 (2016), 85–94. https://doi.org/10.1080/23249676.2016.1214627 doi: 10.1080/23249676.2016.1214627
    [19] F. Unes, H. Varcin, 3-D real dam reservoir model for seasonal thermal density flow, Environ. Eng. Manage. J., 16 (2017), 2009–2024. https://doi.org/10.30638/eemj.2017.209 doi: 10.30638/eemj.2017.209
    [20] F. Üneş, M. Demirci, H. Varçin, 3-D numerical simulation of a real dam reservoir: Thermal stratified flow, in Advances in Hydroinformatics, Springer, Singapore, (2016), 377–394. https://doi.org/10.1007/978-981-287-615-7_26
    [21] N. Penna, M. de Marchis, O. B. Canelas, E. Napoli, A. H. Cardoso, R. Gaudio, Effect of the junction angle on turbulent flow at a hydraulic confluence, Water, 10 (2018), 469. https://doi.org/10.3390/W10040469 doi: 10.3390/w10040469
    [22] F. Salmasi, A. Samadi, Experimental and numerical simulation of flow over stepped spillways, Appl. Water Sci., 8 (2018), 1–11. https://doi.org/10.1007/S13201-018-0877-5/TABLES/3 doi: 10.1007/s13201-017-0639-9
    [23] E. Gabreil, S. J. Tait, S. Shao, A. Nichols, SPHysics simulation of laboratory shallow free surface turbulent flows over a rough bed, J. Hydraul. Res., 56 (2018), 727–747. https://doi.org/10.1080/00221686.2017.1410732 doi: 10.1080/00221686.2017.1410732
    [24] E. Gabreil, H. Wu, C. Chen, J. Li, M. Rubinato, X. Zheng, et al., Three-dimensional smoothed particle hydrodynamics modeling of near-shore current flows over rough topographic surface, Front. Mar. Sci., 9 (2022), 935098. https://doi.org/10.3389/fmars.2022.935098 doi: 10.3389/fmars.2022.935098
    [25] A. J. Odgaard, River channel stabilization with submerged vanes, in Advances in Water Resources Engineering, Springer International Publishing, Cham, (2015), 107–136. https://doi.org/10.1007/978-3-319-11023-3_3
    [26] E. Turhan, H. Ozmen-Cagatay, S. Kocaman, Experimental and numerical investigation of shock wave propagation due to dam-break over a wet channel, Pol. J. Environ. Stud., 28 (2019), 2877–2898. https://doi.org/10.15244/pjoes/92824 doi: 10.15244/pjoes/92824
    [27] H. Ozmen-Cagatay, E. Turhan, S. Kocaman, An experimental investigation of dam-break induced flood waves for different density fluids, Ocean Eng., 243 (2022), 110227. https://doi.org/10.1016/J.OCEANENG.2021.110227 doi: 10.1016/j.oceaneng.2021.110227
    [28] B. E. Launder, D. B. Spalding, Mathematical Models of Turbulence, Academic Press, New York, 1972.
    [29] H. Kim, P. Nanjundan, Y. W. Lee, Numerical study on the sloshing flows in a prismatic tank using natural frequency of the prismatic shapes, Prog. Comput. Fluid Dyn., 21 (2021), 152–160. https://doi.org/10.1504/PCFD.2021.115129 doi: 10.1504/PCFD.2021.115129
    [30] O. Simsek, M. S. Akoz, N. G. Soydan, Numerical validation of open channel flow over a curvilinear broad-crested weir, Prog. Comput. Fluid Dyn., 16 (2016), 364–378. https://doi.org/10.1504/PCFD.2016.080055 doi: 10.1504/PCFD.2016.080055
    [31] C. W. Hirt, B. D. Nichols, Volume of fluid (VOF) method for the dynamics of free boundaries, J. Comput. Phys., 39 (1981), 201–225. https://doi.org/10.1016/0021-9991(81)90145-5 doi: 10.1016/0021-9991(81)90145-5
    [32] S. M. Salim, M. Ariff, S. C. Cheah, Wall y+ approach for dealing with turbulent flows over a wall mounted cube, Prog. Comput. Fluid Dyn., 10 (2010), 341–351. https://doi.org/10.1504/PCFD.2010.035368 doi: 10.1504/PCFD.2010.035368
    [33] A. Gerasimov, Modeling Turbulent Flows with Fluent, presentation, Seminar on Turbulence Modelling for FLUENT users, 2006. Avalible from: http://www.ae.metu.edu.tr/seminar/Turbulence_Seminar/Modelling_turbulent_flows_with_FLUENT.pdf.
    [34] A. J. Odgaard, River training and sediment management with submerged vanes, Am. Soc. Civil Eng., 2009 (2009). https://doi.org/10.1061/9780784409817 doi: 10.1061/9780784409817
    [35] P. Biswas, A. K. Barbhuiya, Effect of submerged vane on three dimensional flow dynamics and bed morphology in river bend, River Res. Appl., 35 (2019), 301–312. https://doi.org/10.1002/rra.3402 doi: 10.1002/rra.3402
  • This article has been cited by:

    1. Neda Safaripour, Mohammad Vaghefi, Amin Mahmoudi, An experimental comparison of 3D velocity components around single and twin piers installed in a sharp bend under the influence of upstream implemented vanes, 2024, 14, 2190-5487, 10.1007/s13201-024-02177-4
  • Reader Comments
  • © 2023 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(2064) PDF downloads(96) Cited by(1)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog