Next Article in Journal
Water and Sediment Quantity and Quality Generated in Check Dams as a Nature-Based Solutions (NbS)
Previous Article in Journal
New Underwater Image Enhancement Algorithm Based on Improved U-Net
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Hydraulic Characteristics of a New Vertical Slot Fishway with Staggered Baffles Configuration

1
The First Construction Company of CCCC Second Harbor Engineering Co., Ltd., Wuhan 430050, China
2
CCCC Second Harbor Engineering Company Ltd., Wuhan 430050, China
3
Research and Development Center of Transport Industry of Intelligent Manufacturing Technologies of Transport Infrastructure, Wuhan 430050, China
4
Engineering Research Institute, China Construction Eighth Engineering Division Corp., Ltd., Shanghai 200122, China
*
Author to whom correspondence should be addressed.
Water 2025, 17(6), 809; https://doi.org/10.3390/w17060809
Submission received: 11 January 2025 / Revised: 26 February 2025 / Accepted: 8 March 2025 / Published: 12 March 2025
(This article belongs to the Section Hydraulics and Hydrodynamics)

Abstract

:
The vertical slot fishway (VSF) has proven effective in mitigating the severe fragmentation of rivers caused by artificial hydraulic structures. While fishways with steeper slopes exhibit better economic performance, increased slope can raise the flow velocity and turbulence, which may hinder fish migration. To address this issue, this study investigated the application of a VSF with a staggered baffle configuration. Through numerical modeling, the hydraulic characteristics of the VSF under various slope ratios and chamber length-to-width (L/B) ratios were investigated, with data validated by physical models. An increase in the slope gradient resulted in higher flow velocities, greater maximum attenuation rates of mainstream velocity, and elevated turbulent kinetic energy (TKE) at the corners of the rectifier baffles and the ends of the divider baffles. Additionally, the overall maximum volumetric energy dissipation ( D ε ) increased, although its distribution pattern remained unaffected. Conversely, increasing the chamber L/B ratio significantly altered the distribution patterns of the flow velocity, TKE, and D ε , influencing their generation mechanisms. For instance, a higher chamber L/B ratio caused the maximum flow velocity (Vm) to deviate from the vertical slot and raised the maximum attenuation rate of the mainstream velocity. The L/B ratio also caused changes in the TKE distribution; as the ratio increased, the proportion of the chamber’s internal region with D ε 150   W / m 3 initially decreased and then increased. Overall, considering the flow velocity, TKE, and D ε , it is recommended that the chamber L/B ratio be maintained between 0.9 and 1.1 for slope ratios ranging from 1:20 to 1:50. The research results will offer practical insights for engineering applications, in engineering construction, contribute theoretical guidance for the optimized design of fish passages, promote sustainable hydraulic engineering practices, and aid in the protection of aquatic biodiversity.

1. Introduction

The establishment of hydraulic structures, including sluices and dams, is crucial for flood control, disaster mitigation, agricultural irrigation, alleviating water scarcity, and developing clean energy [1,2]. While these structures provide significant benefits, they also fragment river ecosystems by transforming continuous waterways into isolated sections that obstruct fish migration, adversely affecting fish populations and reproduction [3,4,5]. Consequently, the development of fish passage systems has become increasingly important. Fish ladders, the most commonly implemented solution, assist fish in overcoming barriers and re-establishing their migratory routes. These structures have demonstrated positive ecological effects by conserving fish resources and maintaining aquatic biodiversity. Among various fishway designs, the VSF has been widely utilized owing to its straightforward construction, efficient energy dissipation, operational flexibility, and excellent adaptability to fluctuations in upstream and downstream water levels [6].
In fishway design, the bottom slope gradient significantly impacts the overall length of the structure, thereby influencing the duration of fish ascension, layout, and construction costs [7]. Ahmadi et al. [8] highlighted that optimizing low-slope VSFs offers a more cost-effective solution, as steeper slopes reduce the overall length and construction costs. However, high flow velocities and turbulent flow fields induced by steep slopes pose significant challenges to fish species with limited swimming abilities during upstream migration [9]. To balance ecological restoration and engineering feasibility, researchers have performed numerous studies on the hydraulic optimization of VSFs [10,11,12,13]. Chorda et al. [14] tested three longitudinal slopes and validated the k ε closure model. They found that the energy dissipation rate substantially affects fish passage efficiency. Tarrade et al. [15] identified two distinct flow topologies with a swirling pattern based on fishway size and slope. Puertas et al. [16] examined the performance of two VSFs at different slopes and concluded that the velocity remains independent of discharge and constant with depth. Ahmadi et al. [17] numerically examined the impact of the slope and cylindrical structure within the fishway on fish swimming performance.
Rajaratnam et al. [18] highlighted the L/B ratio of fishways as a key factor influencing flow variations when examining fishways with different baffle configurations. Wang et al. [19] studied the turbulent characteristics in a scaled VSF model with varying width and slope and revealed that changes in the chamber width directly affected the flow structure and turbulence intensity. Therefore, the L/B ratio can be an important parameter for the optimal design of a fishway.
Currently, conventional vertical slot fishways (VSFs) position both the divider and rectifier baffles on the same side. This linear alignment of vertical slots generates a shorter mainstream flow path compared to staggered configurations. To enhance energy dissipation, this study proposes a staggered configuration for both baffle types, which elongates the mainstream flow path. This modification reduces the maximum velocity (Vm) in the pool chamber to below 1.2 m/s—a value lower than the burst swimming speed (1.2 m/s) for the four major Chinese carps. The staggered design also improves flow uniformity, while maintaining a recirculation zone with resting velocities of 0.3–0.5 m/s. These velocities align with the sustained swimming capacity of cyprinid species, thereby facilitating fish migration and resting. Numerical simulations carried out employed the RNG k-ε turbulence model, while physical modeling was performed to validate the results. Various structural configurations, including the bottom slope ratios and length-to-width ratios of the pool chamber, were examined for their impact on the hydraulic performance of this type of fishway. The findings of this research offer theoretical insights for the optimized design of fish passages and hold practical significance for engineering applications.

2. Materials and Methods

2.1. Research Object

This study involved a fishway located at a water conservancy project on the Han River in Hubei Province, China, which is a critical protection area for the Four major Chinese carps. The Vm within the fishway must not exceed the burst swimming speed of the fish [20], which was identified as less than 1.2 m/s for the four major Chinese carps [21,22,23]. The fishway, designed with a water depth of 2 m, a design discharge of 3.2 m3/s, and a flow velocity of 0.8 m/s, was developed considering the hydrological conditions and the specifications of the hydraulic structure.
The cross-section of the fishway pool chamber is rectangular, measuring 2.0 m in width (B) and featuring a divider baffle length of 1.26 m. The rectifier baffles extend 0.5 m. Both the divider and rectifier baffles have a thickness of 0.20 m. The vertical slot width (b) is 0.33 m. In this study, the L/B ratio of the fishway was controlled by varying the length of the fishway pool chamber (L) from 1.20 to 3.00 m in increments of 0.20 m, so that the L/B ratios varied between 0.6 and 1.5. The selected slope ratios (Ji) for this study are 1:20, 1:30, 1:40, and 1:50. The structural design of the fishway pool chamber is depicted in Figure 1.

2.2. Physical Model

The validation data in this study were obtained from local model tests of the fishway associated with the aforementioned hydraulic engineering project. The experimental setup includes an inlet, an outlet, and ten conventional pools, each measuring L = 2.4 m, B = 2 m, and J = 1:20. The experimental water depth was maintained at 2 m. The model was constructed based on the Froude number similarity. With reference to existing research [14,24], the experimental ratio was set at 1:5. Tests were conducted in a multifunctional test tank, which features plexiglass panels for the sides, bottom, and baffles, with installation errors maintained blow 0.2 mm. The test tank consists of a water reservoir, an energy dissipation mesh, an electromagnetic flowmeter, a manual valve, a downstream sluice gate, a submersible pump, a reflux corridor, and other components. Its inlet section is 4 m long and equipped with an energy dissipation mesh to ensure uniform and smooth flow upstream to the first level of the fishway, while the 4 m long tailgate exit eliminates water turbulence. The experimental setup is illustrated in Figure 2. During testing, the inlet flow was regulated using the manual valve, and the flow rates were measured by the electromagnetic flowmeter, which has an accuracy of 0.4% and a sampling frequency of 10 Hz. To ensure a consistent water depth across all pools, the water depth was regulated by downstream sluice gates, with a water level difference of 0.12 m between adjacent pools.
Due to the vertical slot design and varying side configurations of the fishway, the flow structure can differ between adjacent pools. For validation purposes, the pools from slots “5” to “7” were chosen. In each pool, a cross-section labeled Si was selected at 0.4 m intervals, where i = 1 to 10. This operation resulted in a total of 10 validation cross-sections, each of which contained 10 measurement points; the spacing between measurement points is 0.18 m. To validate the vertical flow velocity, two measurement lines were established at the center of each slot, each of which had 9 measurement points that were vertically spaced at 0.2 m. To assess grid convergence, the velocities at the midpoints (VS2 and VS9) of cross-sections S2 and S9 were utilized to compute the GCI value. The layout of the measurement points for the flow velocities within the pool chamber is illustrated in Figure 3.

2.3. Numerical Simulation

2.3.1. Model Establishment

The numerical model was constructed using FLOW-3D® v11.2, a widely utilized computational fluid dynamics (CFD) software in fishway studies [25,26]. This general purpose program is able to provide transient 3D solutions for complex multiphysics flow problems by solving fluid motion equations [27]. FLOW-3D® has found extensive applications in studies related to hydraulic systems and engineering (e.g., [28,29,30,31,32,33]). A detailed description of FLOW-3D® is available in the Flow Science documentation [27].

2.3.2. Governing Equations

The Navier–Stokes equations were discretized using the finite difference method, while the free surface was modeled using the SOLution Algorithm-based Volume of Fluid method [34]. This method focuses solely on the fluid phase, disregards the air phase, and treats gas elements as voids. Previous studies have demonstrated that this method reduces computational time and effectively visualizes the free surface shape [34,35,36,37]. A Cartesian staggered grid was employed to solve the Reynolds-Averaged Navier–Stokes equations, which consist of continuity and momentum equations. The Cartesian coordinates (x, y, z) employed are defined as follows:
V f ρ t + ( ρ u A x ) x + ( ρ v A y ) y + ( ρ w A z ) z = R SOR
u t + 1 V f ( u A x u x + v A y u y + w A z u z ) = 1 ρ P x + G x + f x
v t + 1 V f u A x v x + v A y v y + w A z v z = 1 ρ P y + G y + f y
w t + 1 V f u A x w x + v A y w y + w A z w z = 1 ρ P w + G z + f z
where u, v, and w represent the velocity components in the x-, y-, and z-directions; Vf is the volume fraction of fluid in each element; RSOR  R SOR is the spring term; Ax, Ay, and Az are the flow area components; ρ is the fluid density; P is the hydrostatic pressure; Gx, Gy, and Gz are the gravitational accelerations in the respective directions; and f represents the Reynolds stress. The viscous accelerations in the x-, y-, and z-directions are denoted by fx, fy, and fz, respectively. The VOF transport equations are expressed as follows:
F t + 1 V F ( F A x u ) x + ( F A y v ) y + ( F A z w ) z = 0
where F denotes the fluid fraction, with F = 1 in fluid-filled elements and F = 0 in fluid-free elements (empty regions) [38,39]. The free surface region corresponds to intermediate values of F (typically F = 0.5, with customizable intermediate values).

2.3.3. Turbulence Model

Studies have demonstrated that the RNG k-ε model offers enhanced capabilities for handling high strain rates and large curvature streamlines [40,41,42,43] and provides more accurate reproductions of the velocity fields [44]. Additionally, it has also demonstrated its effectiveness in addressing complex issues in the simulation of hydraulic structures [45,46,47,48]. Consequently, this model was selected for simulating the water flow within the fishway pool chamber.
The model comprises two essential equations: Equation (6) defines the TKE, denoted k, while Equation (7) describes the turbulent dissipation rate (ε).
( ρ k ) t + ( ρ k u i ) x i = x j [ α k μ e f f k x j ] + G k G B ρ ε Y M + S k
( ρ ε ) t + ( ρ ε u i ) x i = x j [ α ε μ e f f ε x j ] + C 1 ε ε k ( G k + C 3 ε G b ) C 2 ε ρ ε 2 k R ε + S ε
where G k denotes the TKE arising from the velocity gradient; G B indicates the buoyancy-induced TKE; S k and S ε represent the source terms; α k and α ε are the inverse Prandtl numbers for k and ε, respectively; μ e f f is the effective viscosity, defined as μ e f f = μ + μ t , where μ t is the eddy viscosity.
R ε = C μ ρ η 3 ( 1 η / η 0 ) ε 2 k ( 1 + β η 3 )
μ t = ρ C μ k 2 ε
The constants of the model are as follows [43]:
C μ = 0.0845, C 1 ε = 1.42, C 2 ε = 1.68, C 3 ε = 1.0, σ k = 0.7194, σ ε = 0.7194, η 0 = 4.38, and β = 0.012.

2.3.4. Boundary Conditions

The boundary conditions were defined based on site-specific hydraulic calculations and biological requirements. The upstream velocity (0.8 m/s) was derived from the fishway’s design discharge (3.2 m3/s). This value aligns with the burst swimming speed limit of the four major Chinese carps (≤ 1.2   m / s ) and ensures hydrodynamic compatibility with the fishway’s slope (1:20) and roughness (n = 0.015). The downstream boundary was set as a pressure outlet with a fixed water depth of 2 m, consistent with engineering practice. The top boundary was specified as a pressure boundary with zero relative pressure. No-slip boundaries were applied to the side walls and bottom surface of the fishway, with a surface roughness of 0.015 m. An initial water depth of 2 m was established in the fishway model to facilitate convergence and reduce computational time. The model was solved using the Generalized Minimal Residual (GMRes) implicit solver, with an initial time step of 0.001 s and a minimum time step of 10−7 s.

2.4. Grid Independence Verification

A full-scale physical model was utilized to construct a numerical fishway model. A structured hexahedral mesh that demonstrates good convergence properties was utilized. A containing block was applied to the entire model, while a nested block with refined elements was employed for the target study area. The mesh was automatically aligned with the geometry, and grid processing was performed using the FAVOR function. To enhance numerical accuracy and solution convergence, the model’s geometry was refined by selecting an appropriate mesh size for grid division. Based on related research [49], a recommended grid refinement ratio (r = Gcoarse/Gfine) of 1.3 was adopted. The mesh size and total number of elements are presented in Table 1, and Figure 3 shows the mesh configuration.
The GCI values were used to evaluate the convergence of the mesh. This analysis was conducted based on the velocities at the midpoints of cross-sections S2 and S9 (VS2 and VS9), with the results presented in Table 2.
The approximate relative error, denoted by “ε”, was calculated using the following expressions: ε 32 = ϕ 3 ϕ 2 , ε 21 = ϕ 2 ϕ 1 , where ε32 and ε21 represent the approximate relative error between the coarse and medium meshes, as well as between the medium and fine meshes, respectively; ϕ 3 , ϕ 2 , and ϕ 1 denote the velocity values derived from the FLOW-3D® simulations for the fine, medium, and coarse meshes, respectively.
The apparent order of the computational method was determined by the following equations.
p = 1 ln ( r 21 ) | ln | ε 32 / ε 21 | + q ( p ) |
q ( p ) = ln r 21 p s r 32 p s
s = 1 sgn ( ε 32 / ε 21 )
Given a constant value of r = 1.3, q(P) is determined to be 0. The extrapolated values of the method were computed using Equation (13).
ϕ ext 21 = ( r 21 p ϕ 1 ϕ 2 ) / ( r 21 p 1 )
The calculation method for the fine-grid convergence index GCI 32   was consistent with that employed for GCI fine 21 .
GCI fine 21 = 1.25 e a 21 r 21 p 1
e a 21 = ϕ 1 ϕ 2 ϕ 1
e ext 21 = ϕ ext 12 ϕ 1 ϕ ext 12
where e a 21 and e ext 21 denote the approximate and extrapolated relative errors, respectively.
The results indicated that the value of GCI fine 21 was smaller than that of GCI32. It can be concluded that the lattice is independent. The computed value of the GCI fine 21 /GCI32 ratio approaching 1 indicated that the numerical solution fell within the asymptotic convergence region. Figure 4 illustrates the mesh configuration, where the dimensions of the containing and nested blocks are 5.0 cm and 3.8 cm, respectively.

3. Results and Discussion

3.1. Model Validation

As shown in Figure 5, the velocity vectors were extracted at different depths of z = 0.75 h, 0.50 h, and 0.25 h (where z is the depth of the cross-section and h is the water depth), referred to as the surface, middle, and bottom layers, respectively. The flow patterns at these layers for the VSF were fundamentally similar, and the flow field distribution patterns at different water depths were also similar. Due to the resemblance in the flow field distribution patterns at different water depths, a representative mid-level flow field was selected for analysis.
To validate the numerical model, the absolute velocities were calculated using the formula V = u 2 + v 2 and compared with the experimental results, as depicted in Figure 6. The numerical results derived from the medium and fine meshes showed good agreement with the experimental results. Table 3 presents a comparison between the numerical and experimental results at different water layers.
Table 3 demonstrated that the largest discrepancies occurred at Vl1 and Vl2, which correspond to the vertical slot region where the flow structure is highly complex. The maximum error was 5.63%, which indicated that the simulation effectively captured the flow characteristics of the VSF.

3.2. Flow Field

Figure 7 illustrates the velocity distribution in adjacent pools under varying slope ratios, revealing two distinct flow patterns. The first pattern features a mainstream zone at the center of the pool chamber, primarily formed by the jet flow through the slots. The second pattern consists of recirculation zones on either side of the mainstream zone, with one side exhibiting a larger recirculation zone. These recirculation zones were symmetrically distributed along the central axis of the pool chamber. The flow velocities in the recirculation zones were opposite to those in the mainstream zone, generating vertical vortices around the Z-axis perpendicular to the riverbed, which dissipated energy.
In the staggered-side slot-type fishway, the planform of the mainstream zone in adjacent pools exhibited an S-shape, contrasting significantly with the same-side slot-type fishway, where the mainstream zone within a single pool chamber consistently exhibited an S-shape pattern [50,51]. In both configurations, the location of maximum velocity at different cross-sections consistently aligned with the mainstream centerline, and the planform remained unchanged under varying slope ratios. Additionally, the flow velocity increased as the slope ratio increased. The maximum velocity ( U i ) values for each cross-section of the fishway were extracted from the numerical results. These values, along with their corresponding coordinates, were processed using a dimensionless method. By connecting the U i values from each cross-section, the U i curve of the main flow path was constructed [51], where U m a x represents the maximum U i and U m i n denotes the minimum U i . The dimensionless maximum mainstream velocity was determined using the formula U i / U m a x , and the maximum attenuation rate of the mainstream velocity was calculated using the formula 1 U m i n / U m a x . The resulting variation patterns are illustrated in Figure 8 and Table 4.
The results indicated that for a slope ratio of J1 = 1:20, the maximum attenuation rate of the mainstream velocity reached 0.51, indicating significant energy dissipation. When the slope ratio further decreased to J3 = 1:40, the maximum attenuation rate of the mainstream velocity dropped to 0.42 before increasing slightly. This behavior can be attributed to the higher flow velocity at steeper slope ratios, which creates a larger difference between the slot jet velocity and the pool chamber water velocity, thereby enhancing diffusion and energy dissipation. As the slope ratio decreased, the slot jet velocity decreased, weakening both the diffusion and energy dissipation effects. With further reduction in the slope ratio, the slot jet velocity continued to decrease, which increased the time required for the mainstream to traverse the pool chamber. This extended interaction allowed the slot jet to diffuse more completely and merge with the pool chamber water, leading to a slight recovery in energy dissipation efficiency.
The mainstream region serves as the primary passage for fish during upstream migration, and the L/B ratio of the pool chamber significantly influences the curvature of the mainstream flow. Consequently, the degree of curvature affects fish migration behavior. Figure 9 illustrates the velocity distribution within pools at different L/B ratios. When the L/B ratio was 0.6, the curvature of the mainstream flow between adjacent pools was pronounced, characterized by a small radius of curvature and significant distortion. At the vertical slot positions, the mainstream direction underwent abrupt changes, resulting in a linear flow distribution within each pool chamber, while the flow between adjacent chambers formed a complex Z-shaped pattern. At this stage, the smaller recirculation zones on either side of the mainstream were not fully developed, suggesting that when the pool chamber length was too short, the jet flow could not dissipate sufficient energy before entering the next chamber. With increase in the L/B ratio, the curvature of the mainstream flow gradually decreased, and the flow transitioned from a Z-shaped to an S-shaped pattern. During this transition, the recirculation zones in the smaller areas began to develop, although they remained relatively small, with limited effects on energy dissipation. These flow conditions remained suboptimal for fish upstream migration. When the L/B ratio exceeded 0.8, the flow pattern in the mainstream region of a single pool chamber exhibited a semi-parabolic shape, while adjacent chambers displayed an S-shaped flow pattern. As the L/B ratio increased further, the S-shaped flow pattern became more pronounced, and the recirculation zones beneath the rectifier and divider baffles expanded. When the L/B ratio was between 1.0 and 1.2, the mainstream region clearly exhibited an S-shaped flow pattern, with particularly prominent recirculation zones on both sides. This flow structure enhanced energy dissipation within the pool chamber and provided optimal flow conditions and suitable resting areas for fish. However, when the L/B ratio exceeded 1.2, a wall-attached flow phenomenon may occur in the mainstream region, potentially hindering fish migration.
The longitudinal maximum velocity curves for different pool chamber L/B ratios under varying slope ratios were derived from numerical simulations. As an example, the longitudinal maximum velocity curves for pools with different L/B ratios at a bottom slope ratio of J1 = 1:20 were analyzed, as shown in Figure 10. The distributions of these curves under varying L/B ratio conditions were similar. After the first vertical slot jet, the flow decelerated, reached a trough before the second vertical slot, and then accelerated as it passed through the slot. As the L/B ratio increased, the location of the peak velocity shifted towards the vertical slot. This was due to the fact that when the L/B ratio was between 0.6 and 0.8, the mainstream water did not dissipate sufficient energy during its forward movement in the pool chamber. The flow was further dissipated by the disturbance effects of the rectifier baffle, causing the water to decelerate and then accelerate upon entering the vertical slot. In contrast, when the L/B ratio was greater than or equal to 0.9, the mainstream water dissipated sufficient energy before reaching the next vertical slot, and the flow velocity increased immediately as it passed through the slot. To ensure appropriate curvature at the junctions of adjacent pools and to prevent the occurrence of wall-attached flow, it is recommended that the L/B ratio be maintained within the range of 0.9 to 1.1.
The maximum attenuation rate of the mainstream velocity for different pool chamber L/B ratios at various bottom slope ratios is shown in Figure 11. Generally, a larger change in the maximum mainstream velocity along the flow path corresponded to a more significant decay rate, indicating higher head loss and enhanced energy dissipation, which is beneficial for fish migration [52,53]. The results showed that, for a fixed L/B ratio, the distribution patterns of the mainstream maximum velocity decay rate along the flow path were similar across different slope ratios. Specifically, as the L/B ratio increased, the decay rate along the flow path also increased. When the L/B ratio exceeded 1, the velocity decay rate stabilized along the flow path.

3.3. TKE

TKE is a critical hydraulic parameter that characterizes the turbulence within the pool chamber and plays a vital role in assessing the hydraulic performance of fishways. Generally, lower TKE levels are more conducive to fish migration, as higher TKE can hinder fish movement by increasing swimming resistance and reducing their ability to navigate the migration paths. TKE levels are typically classified into two categories: “low” (TKE < 0.05 J/kg) and “high” (TKE > 0.05 J/kg) [54,55]. In fishways, fish tend to prefer areas with “low” TKE levels [56,57]. Figure 12 illustrates the distribution of TKE in pools under varying slope ratios. It was evident that higher TKE values were concentrated around the rectifier baffle angles and the divider baffle heads. This observation indicated that these baffle angles and ends induce flow infusion, which promoted energy dissipation and enhanced turbulence near these structures. The TKE in the recirculation zones flanking the mainstream varied between 0.01 and 0.05 J/kg, suggesting that these zones are suitable for fish to rest during their upstream migration. When the slope ratio was altered while maintaining the configuration and dimensions of the rectifier and divider baffles constant, the TKE distribution pattern remained unchanged, although the TKE decreased as the slope ratio decreased. Notably, the reduction in TKE was most pronounced near the rectifier baffle angles and the ends of the divider baffles. This indicated that a larger slope ratio led to higher energy dissipation efficiency within the fishway.
Figure 13 illustrates the distribution of TKE in pools with varying L/B ratios under a slope ratio of J1 = 1:20. As the L/B ratio increased, the area of the turbulent flow region expanded. As the L/B ratio increased from 0.6 to 1.2, the increase in TKE was primarily due to the combined effects of higher flow velocity, flow guidance, and energy dissipation from the rectifier and divider baffles. When the L/B ratio is greater than or equal to 1.2, this increase was attributed to mainstream flow impacting the chamber walls in the form of a jet, which led to the onset of wall-attached flow. This phenomenon intensified turbulence, thereby increasing TKE and affecting the upstream migration of fish.

3.4. Volumetric Energy Dissipation

The design of VSFs needs to account for the dissipation rate D ε , which represents the volumetric energy dissipation within the pool. Insufficient dissipation can result in excessive energy expenditure by the fish during their upstream migration, potentially hindering successful passage. To simplify fishway design, the global volumetric dissipated power DV has commonly been employed as a key discriminant parameter. It is calculated using the formula DV = P/VP, where P = ρ g Q δ h ( ρ is the water density, g is the gravity, Q is the flow rate, and δ h is the head difference). The DV encapsulates power and friction losses in addition to turbulence losses, making it a global pool parameter [14]. In contrast, D ε reflects the dissipation distribution within the pool. The determination of D ε within the pool chamber is theoretically significant for refining the design and optimizing the internal structure of the pool chamber. Existing research indicated that D ε values between 150 and 200 W/m3 are suitable for fish species with strong swimming abilities. For smaller fishways or species with weaker swimming abilities, such as many cyprinids, D ε should be maintained below 150 W/m3 [58]. Accurately determining the spatial distribution of D ε is essential for evaluating migratory opportunities for fish passing through VSFs [7].
The formula for calculating the D ε   is given by (17):
D ε = 1 V p V p ρ ε d V P
where ρ is the fluid density (kg/m3); V P is the unit volume of the pool chamber; d V P is the infinitesimally small volume of the pool chamber; and ε represents the turbulent dissipation rate.
Figure 14 illustrates the distribution of D ε under varying slope ratios. The highest values of D ε   were located at the heads of the divider baffles and at the angles of the rectifier baffles. Although these regions exhibited higher D ε values, they were relatively small and had minimal impact on fish upstream migration. In contrast, D ε was lower in the recirculation zones, which occupied a larger area. As the slope ratio decreased, D ε also decreased; however, the overall distribution pattern remained unchanged.
Figure 15 and Figure 16 illustrate the distribution of D ε   under different L/B ratios for slope ratios of J1 = 1:20 and J2 = 1:30, and the area proportion of the pool chamber with D ε 150   w / m 3 (Ap) for different L/B ratios, respectively. Figure 15 indicates that D ε 150   w / m 3 was predominantly distributed in the recirculation zones. According to Figure 16, for a slope ratio of J1 = 1:20, the minimum Ap occurred when the L/B ratio equaled 1, whereas for J2 = 1:30, the minimum Ap appeared when the L/B ratio equaled 1.1. This phenomenon was attributed to the increase in flow velocity with rising slope ratios, which led to the earlier onset of wall-attached flow and a quicker reduction in Ap. For a given slope ratio, as the L/B ratio increased, Ap initially decreased and then increased. For instance, when J1 = 1:20 and the L/B ratio was less than 1, an increase in the pool chamber length resulted in higher D ε . Furthermore, the increase in maximum flow velocity was relatively high, thereby leading to better energy dissipation and a reduced Ap. However, when the L/B ratio exceeded 1, as the pool chamber length continued to increase, some of the flow from the vertical-slot jet began to interact with the pool chamber walls, further enhancing energy dissipation. At this point, the chamber length became sufficiently long, causing a slight increase in Ap.

4. Conclusions

Based on three-dimensional numerical simulations, this study investigated the hydraulic characteristics of a novel VSF with staggered baffles configuration, using the slope and the length-to-width (L/B) ratio of the pool chamber as variables. The main findings are as follows:
(1)
Steeper slopes enhance the vertical-slot velocities and the chamber velocity differentials, thereby boosting the energy dissipation efficiency that is critical for fish migration. L/B ratios of 0.9–1.1 optimize the flow patterns, though increased slope steepness necessitates proportional adjustments to the L/B ratios.
(2)
Slope elevation amplifies the turbulent kinetic energy (TKE) at baffle junctions, whereas recirculation zones maintain stable turbulence levels, which are crucial for fish resting. When L/B exceeds or equals 1.2, TKE variations become predominantly governed by wall-attached flow.
(3)
Peak D ε zones at baffle corners exhibit slope and L/B-dependent sensitivity; however, these regions are positioned outside the primary fish migration pathways, thus exerting minimal impact on upstream fish passage. The threshold of D ε ≤ 150 W/m3—derived from fish endurance studies [58]—predominantly occurs in recirculation zones. Non-linear L/B relationships suggest optimal configurations that balance hydraulic performance with ecological requirements.
There is potential for further improvement in fishway configurations for steep slopes. Future studies should consider different flow rates and velocities for a comprehensive analysis. To optimize the VSF design, experimental efforts should emphasize hydraulic characteristics while simultaneously accounting for the swimming behaviors and performance of fish across different fish species and life stages. Future research can concentrate on these two aspects.

Author Contributions

Conceptualization, Y.L.; methodology, Y.L.; investigation, Z.Z.; resources, Z.W.; data curation, D.Z.; writing—original draft preparation, Z.Z.; writing—review and editing Z.W.; visualization, Y.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Data are contained within the article.

Acknowledgments

The authors are very grateful to the Research and Development Center of Transport Industry of Intelligent Manufacturing Technologies of Transport Infrastructure, for supporting this study.

Conflicts of Interest

Authors Yong Lu, Zhimin Wang and Zichen Zhao were employed by the company The First Construction Company of CCCC Second Harbor Engineering Co., Ltd. Author Dongliang Zhao was employed by the company CCCC Second Harbor Engineering Company Ltd. Author Yonggang Zhang was employed by the company Engineering Research Institute, China Construction Eighth Engineering Division Corp., 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. Kuriqi, A.; Pinheiro, A.N.; Sordo-Ward, A.; Garrote, L. Influence of hydrologically based environmental flow methods on flow alteration and energy production in a run-of-river hydropower plant. J. Clean. Prod. 2019, 232, 1028–1042. [Google Scholar] [CrossRef]
  2. Chen, S.; Chen, B.; Fath, B.D. Assessing the cumulative environmental impact of hydropower construction on river systems based on energy network model. Renew. Sust. Energ. Rev. 2015, 42, 78–92. [Google Scholar] [CrossRef]
  3. Zarfl, C.; Berlekamp, J.; He, F.; Jähnig, S.C.; Darwall, W.; Tockner, K. Future large hydropower dams impact global freshwater megafauna. Sci. Rep. 2019, 9, 18531. [Google Scholar] [CrossRef] [PubMed]
  4. Couto, T.B.; Messager, M.L.; Olden, J.D. Safeguarding migratory fish via strategic planning of future small hydropower in Brazil. Nat. Sustain. 2021, 4, 409–416. [Google Scholar] [CrossRef]
  5. Verhelst, P.; Reubens, J.; Buysse, D.; Goethals, P.; Van Wichelen, J.; Moens, T. Toward a roadmap for diadromous fish conservation: The Big Five considerations. Front. Ecol. Environ. 2021, 19, 396–403. [Google Scholar] [CrossRef]
  6. Katopodis, C.; Cai, L.; Johnson, D. Sturgeon survival: The role of swimming performance and fish passage research. Fish. Res. 2019, 212, 162–171. [Google Scholar] [CrossRef]
  7. Quaranta, E.; Katopodis, C.; Comoglio, C. Effects of bed slope on the flow field of vertical slot fishways. River Res. Appl. 2019, 35, 656–668. [Google Scholar] [CrossRef]
  8. Ahmadi, M.; Ghaderi, A.; MohammadNezhad, H.; Kuriqi, A.; Di Francesco, S. Numerical investigation of hydraulics in a vertical slot fishway with upgraded configurations. Water 2021, 13, 2711. [Google Scholar] [CrossRef]
  9. An, R.; Li, J.; Liang, R.; Tuo, Y. Three-dimensional simulation and experimental study for optimising a vertical slot fishway. J. Hydro-Environ. Res. 2016, 12, 119–129. [Google Scholar] [CrossRef]
  10. Tarrade, L.; Texier, A.; David, L.; Larinier, M. Topologies and measurements of turbulent flow in vertical slot fishways. Hydrobiologia 2008, 609, 177–188. [Google Scholar] [CrossRef]
  11. Calluaud, D.; Pineau, G.; Texier, A.; David, L. Modification of vertical slot fishway flow with a supplementary cylinder. J. Hdraul. Res. 2014, 52, 614–629. [Google Scholar] [CrossRef]
  12. Sanagiotto, D.G.; Rossi, J.B.; Bravo, J.M. Applications of computational fluid dynamics in the design and rehabilitation of nonstandard vertical slot fishways. Water 2019, 11, 199. [Google Scholar] [CrossRef]
  13. Zhang, D.; Xu, Y.; Deng, J.; Shi, X.; Liu, Y. Relationships among the fish passage efficiency, fish swimming behavior, and hydraulic properties in a vertical-slot fishway. Fish. Manag. Ecol. 2024, 31, e12681. [Google Scholar] [CrossRef]
  14. Chorda, J.; Maubourguet, M.M.; Roux, H.; Larinier, M.; Tarrade, L.; David, L. Two-dimensional free surface flow numerical model for vertical slot fishways. J. Hydraul. Res. 2010, 48, 141–151. [Google Scholar] [CrossRef]
  15. Tarrade, L.; Pineau, G.; Calluaud, D.; Texier, A.; David, L.; Larinier, M. Detailed experimental study of hydrodynamic turbulent flows generated in vertical slot fishways. Environ. Fluid Mech. 2011, 11, 1–21. [Google Scholar] [CrossRef]
  16. Puertas, J.; Pena, L.; Teijeiro, T. Experimental approach to the hydraulics of vertical slot fishways. J. Hydraul. Eng. 2004, 130, 10–23. [Google Scholar] [CrossRef]
  17. Ahmadi, M.; Kuriqi, A.; Nezhad, H.M.; Ghaderi, A.; Mohammadi, M. Innovative configuration of vertical slot fishway to enhance fish swimming conditions. J. Hydrodyn. 2022, 34, 917–933. [Google Scholar] [CrossRef]
  18. Rajaratnam, N.; Katopodis, C.; Solanki, S. New designs for vertical slot fishways. Can. J. Civ. Eng. 1992, 19, 402–414. [Google Scholar] [CrossRef]
  19. Wang, R.W.; David, L.; Larinier, M. Contribution of experimental fluid mechanics to the design of vertical slot fish passes. Knowl. Manag. Aquat. Ecosyst. 2010, 396, 2. [Google Scholar] [CrossRef]
  20. Peake, S.J.; Beamish, F.W.; McKinley, R.S.; Scruton, D.A.; Katopodis, C. Relating swimming performance of lake sturgeon, Acipenser fulvescens, to fishway desing. Can. J. Fish Aquat. Sci. 1997, 54, 1361–1366. [Google Scholar] [CrossRef]
  21. Tan, J.; Li, H.; Guo, W.; Tan, H.; Ke, S.; Wang, J.; Shi, X. Swimming performance of four carps on the Yangtze River for fish passage design. Sustainability 2021, 13, 1575. [Google Scholar] [CrossRef]
  22. Li, H.; Cai, D.; Yang, P. Swimming ability and behavior of different sized silver carp. J. Hydroecol. 2016, 37, 88–92. [Google Scholar]
  23. Hou, Y.; Newbold, L.; Cai, L.; Wang, X.; Hu, W.; Qiao, Y. Swimming performance of juvenile Aristichthys nobilis under fixed velocity swimming tests. Chin. J. Ecol. 2016, 35, 1583–1588. [Google Scholar]
  24. Bombač, M.; Novak, G.; Rodič, P.; Četina, M. Numerical and physical model study of a vertical slot fishway. J. Hydrol. Hydromech. 2014, 62, 150–159. [Google Scholar] [CrossRef]
  25. Tarena, F.; Nyqvist, D.; Katopodis, C.; Comoglio, C. Computational fluid dynamics in fishway research-a systematic review on upstream fish passage solutions. J. Ecohydraul. 2024, 10, 107–126. [Google Scholar] [CrossRef]
  26. Fuentes-Pérez, J.F.; Quaresma, A.L.; Pinheiro, A.; Sanz-Ronda, F.J. OpenFOAM vs. FLOW-3D: A comparative study of vertical slot fishway modelling. Ecol. Eng. 2022, 174, 106446. [Google Scholar] [CrossRef]
  27. Flow Science, Inc. Flow-3D Version 11.2 User Manual; Flow Science, Inc.: Los Alamos, NM, USA, 2016. [Google Scholar]
  28. Dargahi, B. Flow characteristics of bottom outlets with moving gates. J. Hydraul. Res. 2010, 48, 476–482. [Google Scholar] [CrossRef]
  29. Savage, B.M.; Johnson, M.C. Flow over ogee spillway: Physical and numerical model case study. J. Hydraul. Eng. 2001, 127, 640–649. [Google Scholar] [CrossRef]
  30. Bombardelli, F.A.; Meireles, I.; Matos, J. Laboratory measurements and multi-block numerical simulations of the mean flow and turbulence in the non-aerated skimming flow region of steep stepped spillways. Environ. Fluid Mech. 2011, 11, 263–288. [Google Scholar] [CrossRef]
  31. Abad, J.D.; Rhoads, B.L.; Güneralp, İ.; Garcia, M.H. Flow structure at different stages in a meander-bend with bendway weirs. J. Hydraul. Eng. 2008, 134, 1052–1063. [Google Scholar] [CrossRef]
  32. Duguay, J.M.; Lacey, R.W.J.; Gaucher, J. A case study of a pool and weir fishway modeled with OpenFOAM and FLOW-3D. Ecol. Eng. 2017, 103, 31–42. [Google Scholar] [CrossRef]
  33. Bayon, A.; Valero, D.; García-Bartual, R.; López-Jiménez, P.A. Performance assessment of OpenFOAM and FLOW-3D in the numerical modeling of a low Reynolds number hydraulic jump. Environ. Model. Softw. 2016, 80, 322–335. [Google Scholar] [CrossRef]
  34. 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]
  35. Ghaderi, A.; Abbasi, S. Experimental and numerical study of the effects of geometric appendance elements on energy dissipation over stepped spillway. Water 2021, 13, 957. [Google Scholar] [CrossRef]
  36. Daneshfaraz, R.; Ghaderi, A.; Sattariyan, M.; Alinejad, B.; Asl, M.M.; Di Francesco, S. Investigation of local scouring around hydrodynamic and circular pile groups under the influence of river material harvesting pits. Water 2021, 13, 2192. [Google Scholar] [CrossRef]
  37. Di Francesco, S.; Biscarini, C.; Manciola, P. Numerical simulation of water free-surface flows through a front-tracking lattice Boltzmann approach. J. Hydroinf. 2015, 17, 1–6. [Google Scholar] [CrossRef]
  38. Silva, A.T.; Bermúdez, M.; Santos, J.M.; Rabuñal, J.R.; Puertas, J. Pool-type fishway design for a potamodromous cyprinid in the Iberian Peninsula: The Iberian barbel—Synthesis and future directions. Sustainability 2020, 12, 3387. [Google Scholar] [CrossRef]
  39. Roache, P.J. Perspective: A method for uniform reporting of grid refinement studies. J. Fluids Eng. 1994, 116, 405–413. [Google Scholar] [CrossRef]
  40. Quaresma, A.L.; Romao, F.; Branco, P.; Ferreira, M.T.; Pinheiro, A.N. Multi slot versus single slot pool-type fishways: A modelling approach to compare hydrodynamics. Ecol. Eng. 2018, 122, 197–206. [Google Scholar] [CrossRef]
  41. Ghaderi, A.; Dasineh, M.; Aristodemo, F.; Aricò, C. Numerical simulations of the flow field of a submerged hydraulic jump over triangular macroroughnesses. Water 2021, 13, 674. [Google Scholar] [CrossRef]
  42. Pourshahbaz, H.; Abbasi, S.; Pandey, M.; Pu, J.H.; Taghvaei, P.; Tofangdar, N. Morphology and hydrodynamics numerical simulation around groynes. ISH J. Hydraul. Eng. 2022, 28, 53–61. [Google Scholar] [CrossRef]
  43. Yakhot, V.; Orszag, S.A. Renormalization group analysis of turbulence. I. Basic theory. J. Sci. Comput. 1986, 1, 3–51. [Google Scholar] [CrossRef]
  44. Santos, H.A.; Pinheiro, A.P.; Mendes, L.M.M.; Junho, R.A.C. Turbulent flow in a central vertical slot Fishway: Numerical assessment with RANS and LES schemes. J. Irrig. Drain. Eng. 2022, 148, 04022025. [Google Scholar] [CrossRef]
  45. Ghaderi, A.; Dasineh, M.; Aristodemo, F.; Ghahramanzadeh, A. Characteristics of free and submerged hydraulic jumps over different macroroughnesses. J. Hydroinf. 2020, 22, 1554–1572. [Google Scholar] [CrossRef]
  46. Daneshfaraz, R.; Ghaderi, A.; Akhtari, A.; Di Francesco, S. On the effect of block roughness in ogee spillways with flip buckets. Fluids 2020, 5, 182. [Google Scholar] [CrossRef]
  47. Ghaderi, A.; Abbasi, S.; Di Francesco, S. Numerical study on the hydraulic properties of flow over different pooled stepped spillways. Water 2021, 13, 710. [Google Scholar] [CrossRef]
  48. Ran, D.; Wang, W.; Hu, X. Three-dimensional numerical simulation of flow in trapezoidal cutthroat flumes based on FLOW-3D. Front. Agric. Sci. Eng. 2018, 5, 168–176. [Google Scholar] [CrossRef]
  49. Celik, I.B.; Ghia, U.; Roache, P.J.; Freitas, C.J. Procedure for estimation and reporting of uncertainty due to discretization in CFD applications. J. Fluids Eng.-Trans. ASME 2008, 130, 078001. [Google Scholar]
  50. Liu, M.; Rajaratnam, N.; Zhu, D.Z. Mean flow and turbulence structure in vertical slot fishways. J. Hydraul. Eng. 2006, 132, 765–777. [Google Scholar] [CrossRef]
  51. Guo, W.D.; Sun, L.; Gao, Y.; Liang, Y.; Xu, W.; Hu, Z.H. Study of Hydraulic Characteristics in One Side Vertical Slot Fish Way. Water Resour. Power 2012, 30, 81–83. [Google Scholar]
  52. Zhang, C.; Sun, S.K.; Li, G.N. Study on the improvement of detailed structure in the vertical slot fishways. J. China Inst. Water Resour. Hydropower Res. 2017, 15, 389–396. [Google Scholar]
  53. Lü, Q.; Sun, S.K.; Bian, Y.H. Hydraulic Characteristics and Flow Structure in A Double Vertical Slot Fishway. J. Hydroecol. 2016, 37, 55–62. [Google Scholar]
  54. Santos, J.M.; Silva, A.; Katopodis, C.; Pinheiro, P.; Pinheiro, A.; Bochechas, J.; Ferreira, M.T. Ecohydraulics of pool-type fishways: Getting past the barriers. Ecol. Eng. 2012, 48, 38–50. [Google Scholar] [CrossRef]
  55. Marriner, B.A.; Baki, A.B.; Zhu, D.Z.; Cooke, S.J.; Katopodis, C. The hydraulics of a vertical slot fishway: A case study on the multi-species Vianney-Legendre fishway in Quebec, Canada. Ecol. Eng. 2016, 90, 190–202. [Google Scholar] [CrossRef]
  56. Silva, A.T.; Katopodis, C.; Santos, J.M.; Ferreira, M.T.; Pinheiro, A.N. Cyprinid swimming behaviour in response to turbulent flow. Ecol. Eng. 2012, 44, 314–328. [Google Scholar] [CrossRef]
  57. Li, Y.; Wang, X.; Xuan, G.; Liang, D. Effect of parameters of pool geometry on flow characteristics in low slope vertical slot fishways. J. Hydraul. Res. 2019, 58, 395–407. [Google Scholar] [CrossRef]
  58. Tan, J.; Gao, Z.; Dai, H.; Yang, Z.; Shi, X. Effects of turbulence and velocity on the movement behaviour of bighead carp (Hypophthalmichthys nobilis) in an experimental vertical slot fishway. Ecol. Eng. 2019, 127, 363–374. [Google Scholar] [CrossRef]
Figure 1. Layout of fishways (unit: m): B and L represent the width and length of a conventional pool, respectively; cy and cx denote the length and width of the divider baffle, respectively; cd is the diameter of the pre-positioned head of the divider baffle; β is the angle of the rectifier baffle; dy and dx represent the length and width of the rectifier baffle, respectively; l is the distance from the center of the vertical slot to the chamber wall; h represents the height of both the rectifier and divider baffles; J is the fishway slope ratio. (a) Plan view of the fishway; (b) Elevation view of the fishway.
Figure 1. Layout of fishways (unit: m): B and L represent the width and length of a conventional pool, respectively; cy and cx denote the length and width of the divider baffle, respectively; cd is the diameter of the pre-positioned head of the divider baffle; β is the angle of the rectifier baffle; dy and dx represent the length and width of the rectifier baffle, respectively; l is the distance from the center of the vertical slot to the chamber wall; h represents the height of both the rectifier and divider baffles; J is the fishway slope ratio. (a) Plan view of the fishway; (b) Elevation view of the fishway.
Water 17 00809 g001
Figure 2. Conceptual diagram of the experimental model.
Figure 2. Conceptual diagram of the experimental model.
Water 17 00809 g002
Figure 3. Layout of flow velocity measurement points in the chamber at L/B = 1.2 of the J1 = 1:20: (a) Within the chamber; (b) At the vertical slots.
Figure 3. Layout of flow velocity measurement points in the chamber at L/B = 1.2 of the J1 = 1:20: (a) Within the chamber; (b) At the vertical slots.
Water 17 00809 g003
Figure 4. Sketch of the mesh configuration.
Figure 4. Sketch of the mesh configuration.
Water 17 00809 g004
Figure 5. The flow field distribution at L/B = 0.9 of the J1 = 1:20 of the pool: (a) The surface; (b) The middle; (c) The bottom.
Figure 5. The flow field distribution at L/B = 0.9 of the J1 = 1:20 of the pool: (a) The surface; (b) The middle; (c) The bottom.
Water 17 00809 g005
Figure 6. Comparison of numerical and experimental results: (a) Validation of planar flow velocity; (b) Validation of perpendicular flow velocity at the vertical slots.
Figure 6. Comparison of numerical and experimental results: (a) Validation of planar flow velocity; (b) Validation of perpendicular flow velocity at the vertical slots.
Water 17 00809 g006
Figure 7. Flow velocity distribution in the chamber at L/B = 0.7 under varying slope ratios: (a) J1 = 1:20; (b) J2 = 1:30; (c) J3 = 1:40; (d) J4 = 1:50.
Figure 7. Flow velocity distribution in the chamber at L/B = 0.7 under varying slope ratios: (a) J1 = 1:20; (b) J2 = 1:30; (c) J3 = 1:40; (d) J4 = 1:50.
Water 17 00809 g007
Figure 8. Distribution of the dimensionless maximum mainstream velocity at L/B = 0.7 (from slots “5” to “9”). Ui represents the maximum velocity at each cross-section within the pool chamber, and Umax denotes the peak value of the mainstream maximum velocity curve.
Figure 8. Distribution of the dimensionless maximum mainstream velocity at L/B = 0.7 (from slots “5” to “9”). Ui represents the maximum velocity at each cross-section within the pool chamber, and Umax denotes the peak value of the mainstream maximum velocity curve.
Water 17 00809 g008
Figure 9. Flow velocity distribution in the chamber at different L/B ratios: (a) J1 = 1:20, L/B = 0.6; (b) J2 = 1:30, L/B = 0.6; (c) J1 = 1:20, L/B = 0.8; (d) J2 = 1:30, L/B = 0.8; (e) J1 = 1:20, L/B = 1.0; (f) J2 = 1:30, L/B = 1.0; (g) J1 = 1:20, L/B = 1.2; (h) J2 = 1:30, L/B = 1.2.
Figure 9. Flow velocity distribution in the chamber at different L/B ratios: (a) J1 = 1:20, L/B = 0.6; (b) J2 = 1:30, L/B = 0.6; (c) J1 = 1:20, L/B = 0.8; (d) J2 = 1:30, L/B = 0.8; (e) J1 = 1:20, L/B = 1.0; (f) J2 = 1:30, L/B = 1.0; (g) J1 = 1:20, L/B = 1.2; (h) J2 = 1:30, L/B = 1.2.
Water 17 00809 g009
Figure 10. Longitudinal maximum flow velocity curves for different L/B ratios at J1 = 1:20.
Figure 10. Longitudinal maximum flow velocity curves for different L/B ratios at J1 = 1:20.
Water 17 00809 g010
Figure 11. Maximum attenuation rate of the mainstream velocity for different L/B ratios under varying slope ratios.
Figure 11. Maximum attenuation rate of the mainstream velocity for different L/B ratios under varying slope ratios.
Water 17 00809 g011
Figure 12. Distribution of turbulent kinetic energy in the chamber under varying slope ratios at L/B = 0.7: (a) J1 = 1:20; (b) J2 = 1:30; (c) J3 = 1:40; (d) J4 = 1:50.
Figure 12. Distribution of turbulent kinetic energy in the chamber under varying slope ratios at L/B = 0.7: (a) J1 = 1:20; (b) J2 = 1:30; (c) J3 = 1:40; (d) J4 = 1:50.
Water 17 00809 g012
Figure 13. Distribution of TKE in the chamber at different L/B ratios for J1 = 1:20: (a) L/B = 0.6; (b) L/B = 0.8; (c) L/B = 1.0; (d) L/B = 1.2.
Figure 13. Distribution of TKE in the chamber at different L/B ratios for J1 = 1:20: (a) L/B = 0.6; (b) L/B = 0.8; (c) L/B = 1.0; (d) L/B = 1.2.
Water 17 00809 g013
Figure 14. Distribution of D ε in the chamber at L/B = 0.7 under varying slope ratios: (a) J1 = 1:20; (b) J2 = 1:30; (c) J3 = 1:40; (d) J4 = 1:50.
Figure 14. Distribution of D ε in the chamber at L/B = 0.7 under varying slope ratios: (a) J1 = 1:20; (b) J2 = 1:30; (c) J3 = 1:40; (d) J4 = 1:50.
Water 17 00809 g014
Figure 15. Distribution of D ε in the chamber under different L/B ratios: (a) J1 = 1:20, L/B = 0.6; (b) J2 = 1:30, L/B = 0.6; (c) J1 = 1:20, L/B = 0.8; (d) J2 = 1:30, L/B = 0.8; (e) J1 = 1:20, L/B = 1.0; (f) J2 = 1:30, L/B = 1.0; (g) J1 = 1:20, L/B = 1.2; (h) J2 = 1:30, L/B = 1.2.
Figure 15. Distribution of D ε in the chamber under different L/B ratios: (a) J1 = 1:20, L/B = 0.6; (b) J2 = 1:30, L/B = 0.6; (c) J1 = 1:20, L/B = 0.8; (d) J2 = 1:30, L/B = 0.8; (e) J1 = 1:20, L/B = 1.0; (f) J2 = 1:30, L/B = 1.0; (g) J1 = 1:20, L/B = 1.2; (h) J2 = 1:30, L/B = 1.2.
Water 17 00809 g015
Figure 16. Area proportion of the pool chamber with D ε 150 W / m 3 under different L/B ratios.
Figure 16. Area proportion of the pool chamber with D ε 150 W / m 3 under different L/B ratios.
Water 17 00809 g016
Table 1. Details of the constructed meshes.
Table 1. Details of the constructed meshes.
MeshNested Block
Element Size
Containing Block Element SizeNumber of ElementMesh Type
13.8 cm5.0 cm2,689,588Fine
24.2 cm5.5 cm2,008,890Medium
34.6 cm6.0 cm1,533,248Coarse
Table 2. Summary of grid convergence index (GCI).
Table 2. Summary of grid convergence index (GCI).
Quantity ε 32 ε 21 p GC I fine 21 GCI32 GC I fine 21 /GCI32
V S 2 −0.15−0.120.770.490.550.90
V S 9 −0.12−0.090.990.380.420.92
Table 3. Comparison of velocities from numerical modeling and experimental measurements at different water layers.
Table 3. Comparison of velocities from numerical modeling and experimental measurements at different water layers.
LocationVl1Vl2VS2VS9
NumExpNumExpNumExpNumExp
z = 0.4 m0.0180.0640.0800.1000.0140.0800.0140.075
z = 0.8 m0.4320.4220.2320.2160.1860.1730.1740.162
z = 1.2 m0.4460.4950.4460.4550.3570.3640.3350.341
z = 1.6 m0.8210.8850.9490.7500.9600.6000.9000.563
z = 2 m0.0810.0360.0600.0980.0500.0780.0450.090
Mean Error (%)4.285.635.54.7
Table 4. The maximum attenuation rate of the mainstream velocity.
Table 4. The maximum attenuation rate of the mainstream velocity.
Slope Ratio 1 U min / U max
J1 = 1:200.51
J2 = 1:300.48
J3 = 1:400.42
J4 = 1:500.43
Note: Umin is the trough value of the mainstream maximum velocity curve.
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

Lu, Y.; Wang, Z.; Zhao, Z.; Zhao, D.; Zhang, Y. Hydraulic Characteristics of a New Vertical Slot Fishway with Staggered Baffles Configuration. Water 2025, 17, 809. https://doi.org/10.3390/w17060809

AMA Style

Lu Y, Wang Z, Zhao Z, Zhao D, Zhang Y. Hydraulic Characteristics of a New Vertical Slot Fishway with Staggered Baffles Configuration. Water. 2025; 17(6):809. https://doi.org/10.3390/w17060809

Chicago/Turabian Style

Lu, Yong, Zhimin Wang, Zichen Zhao, Dongliang Zhao, and Yonggang Zhang. 2025. "Hydraulic Characteristics of a New Vertical Slot Fishway with Staggered Baffles Configuration" Water 17, no. 6: 809. https://doi.org/10.3390/w17060809

APA Style

Lu, Y., Wang, Z., Zhao, Z., Zhao, D., & Zhang, Y. (2025). Hydraulic Characteristics of a New Vertical Slot Fishway with Staggered Baffles Configuration. Water, 17(6), 809. https://doi.org/10.3390/w17060809

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