Next Article in Journal
Combining Statistical Clustering with Hydraulic Modeling for Resilient Reduction of Water Losses in Water Distribution Networks: Large Scale Application Study in the City of Patras in Western Greece
Next Article in Special Issue
Discharge Coefficients of a Specific Vertical Slot Fishway Geometry—New Fitting Parameters
Previous Article in Journal
Fe3O4 Nanoparticles Loaded Bentonite/Sawdust Interface for the Removal of Methylene Blue: Insights into Adsorption Performance and Mechanism via Experiments and Theoretical Calculations
Previous Article in Special Issue
Brown Trout Upstream Passage Performance for a Fishway with Water Drops between Pools beyond Fish Passage Design Recommendations
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Uphill Flow Rock Ramps. How the Design Impacts Their Functionality

by
Leticia Carrero-Díez
*,
Carolina Martínez Santa-María
and
J. Anastasio Fernández-Yuste
Department of Forest and Environmental Engineering and Management, School of Forest Engineering and Natural Resources, Universidad Politécnica de Madrid, C/José Antonio Novais 10, 28040 Madrid, Spain
*
Author to whom correspondence should be addressed.
Water 2022, 14(21), 3492; https://doi.org/10.3390/w14213492
Submission received: 15 September 2022 / Revised: 26 October 2022 / Accepted: 27 October 2022 / Published: 1 November 2022
(This article belongs to the Special Issue Advances and Experiences in Fishway Design and Assessment)

Abstract

:
Enhancing river passability is considered a central part of the efforts to maintain fish population and achieving good ecological status, according to the EU Water Framework Directive (WFD). One commonly proposed approach to achieving this aim involves the creation of fishways. However, recent studies have shown that many of these fish passes are often lacking an optimal design with far reaching consequences for fish migration. Several promising new designs such as nature-like rock ramps, with uphill flow, have been recently developed. Such studies attempt to address these drawbacks by adapting several structure-related features (i.e., boulder size and shape and friction walls). In this study, we used a 2D computational fluid dynamic model to assess how the key hydraulic variables (water depth, velocity and turbulent kinetic energy) were impacted by different design elements of uphill flow fishways with different configurations. With regard to the standard boulder shapes and sizes, our results reveal that: (1) doubling the boulder sizes results in a decrease of turbulent kinetic energy in resting corridors of up to 33%; (2) the inclusion of small friction-walls in the ramp design increases uphill velocity in the intermediate corridors by up to 49%; (3) the trapezoidal shape of the boulder leads to the largest decreases in maximum velocity in the gaps (16%) and the largest increases in the uphill velocity in the resting corridors (180%). These results may allow us to optimise the uphill flow rock-ramp design to improve the passability of this type of fishways.

Graphical Abstract

1. Introduction

Many of the aquatic organisms in rivers migrate in order to complete their biological cycle. These migrations allow them to access habitats with suitable environmental conditions (i.e., water temperature, dissolved oxygen) and suitable hydraulic biotopes (depth, velocity and roughness) to take advantage of both high-quality breeding sites and bountiful feeding areas [1]. The fragmentation of rivers through barriers on lotic systems has had negative consequences on riverine ecosystems. Therefore, the viability of many fluvial species depends on the functional connectivity of their populations through dispersal. Transversal barriers induce localized changes that alter the continuum of stream temperature, water chemistry, energy, sediment and aquatic organism movements [2,3]. One commonly proposed approach to conserve connectivity or to mitigate the fragmentation is the construction of fish passages. There are a number of diverse topologies of fish passages [4,5,6]. Despite the clear importance of their use, their optimal design remains an object of intense debate and research [7]. Therefore, there are a myriad of studies in the literature focused on assessing aspects such as fishway attraction and entrance, and passage efficiency [8]. Nature-like fishways are currently very popular across the world as an alternative passage system with comparable efficiencies to technical fishways [9].
Rock ramps, or more commonly, ramps, are one of the most highly recommended passage devices for low-height obstacles due to their naturalized nature, their versatility, and their high degree of passability by fish fauna [10]. They consist of a long gentle sloping channel with a naturalized substrate and interspersed boulders, designed to maintain biologically suitable depth and velocity conditions over a limited range of water level changes. Their hydraulic conditions are dependent on the slope of the channel and by the arrangement and size of the boulders. Regarding the arrangements of the boulders, two approaches are widely used: longitudinally aligned in rows perpendicular to the main flow; or randomly distributed [11]. In these ramps, the water flows through the spillways, or the existing notches/gaps, between two consecutive boulders. The critical conditions for fish passability are given by the velocity in the flow gap, the minimum depth achieved in the channel, the jump in the cases of no submerged passage, and the power generated in the water drop and its dissipation to fulfil suitable thresholds. The design of these ramps must ensure certain hydraulic conditions of the structure for the operating flows. The evaluation of the suitability of these hydraulic conditions must be conducted through the variables mentioned above, ensuring that their ranges of variation are compatible with the swimming capacities of the target species [4,11,12,13,14,15]. Moreover, recent studies including both fish behaviour and hydraulics in detail, and have pointed out the importance of other, more specific, hydraulic variables, such as velocity distribution, turbulent kinetic energy, turbulent intensity and vorticity to improve the design, evaluation and efficiency of these type of ramps [16,17,18,19,20,21,22,23,24].
A new design for rock-ramps has recently been developed: uphill flow rock-ramps [25]. The uphill flow ramps, while keeping many of the elements of the traditional rock-ramps, are able to generate a very particular hydraulic performance, as they induce uphill flows paths in opposition to the main flow in the ramp. This is achieved through their particular zig-zag distribution of boulders within the row (Figure 1).
The biological importance of these new uphill flow paths shows great potential as they favour the progression of upstream fishes to such an extent that they could advance up the ramp, pushed by these uphill flows, and would only have to swim to pass the gap between two boulders.
An uphill flow ramp design has already been built in the Manzanares river (Madrid, Spain), but neither hydraulic variables data nor fish passage records are currently available. However, new uphill flow ramps are currently under planning in diverse rivers belonging to the Duero, Guadalquivir, Guadiana and Miño-Sil river basins (Spain). Updated one-dimensional equations [26] considering non-uniform flow analysis [27] have been used for their design [25].
The main objective of this research was to identify alternative designs that improve the functionality of uphill flow rock ramps by applying IBER, an bidimensional model [28]. In this paper, four different uphill flow ramp configurations were performed. We aimed to answer three questions: Do changes in boulder dimensions improve uphill flow paths conditions and reduce the turbulent kinetic energy? How can the incorporation of friction-increasing walls improve the uphill velocity conditions and decrease the turbulent kinetic energy? Does the shape of the boulder reduce the maximum water velocity in the flow gaps? The results of this study may be useful to fish pass planners for enhancing the design of uphill flow rock-ramps, concerning their functionality and, hence, improving their passage efficiency.

2. Materials and Methods

2.1. Description of the Geometrical Designs

The present paper studied an uphill flow ramp, designed by upflow ramps dimensioning assistant [29]. We defined four different ramp layouts in order to numerically investigate the previously formulated hypothesis. The first layout (R3) reproduced the hydraulic behaviour of a standard three module uphill flow ramp (Figure 2 and Figure 3). The other three layouts replicated the standard ramp with some geometric modifications, in order to aim at exploring the effect of: (1) boulder dimensions (Layout DR3); (2) friction (Layout DRb3); (3) shape of boulders (Layout DT3). These layouts (Figure 4, Figure 5 and Figure 6) allowed direct comparison between the standard ramp (R3) and the other designs, and allowed us to verify, using a numerical model performance, their effects over some specific hydraulic variables.
The generation of the uphill flow is guaranteed, as long as the boulders in the rock-row are distributed with a certain disposition, as exposed in Figure 1. This distribution causes the flow through the pools to be compartmentalized. Then areas with uphill flow are interspersed with downhill flow areas, which are referred to as corridors. The uphill flow corridors are always located in the vicinity of the walls due to the influence of the roughness of the contour, and are under the protection of the down face of the lowest positioned boulder in each row.
With this boulder distribution in mind, the four different layouts were designed.
-
Rectangular boulders (R3)
This study used a standard ramp (R3) with rectangular boulders and three modules. A module is the minimum unit required to generate uphill flow and is formed by three consecutive boulders, belonging to one particular rock-row and placed in the correct distribution. The uphill flow ramp design comprises three phases: hydraulic dimensioning, geometric dimensioning, and power dissipation dimensioning. The most relevant data for the hydraulic dimensioning is the minimum flow for which the ramp must be functional. In this case, 2 m3 s−1 was fixed by considering the construction of three modules for the design and ensuring a minimum depth. The dimensioning was conducted in a uniform regime, assuming that the depths upstream and downstream of the boulders are the same in all rock-rows, being 0.6 and 0.3 m, respectively, and with a geometric gap width (WGg) of 0.35 m. Using the aforementioned values, the geometric data were generated by the application, fixing a height difference on the ramp bed between the inlet and outfall of 1.9 m in seven rows and six pools. Considering a compatible boulder diameter (Db) of 0.7 m, a boulder width (Wb) of 0.5 m, and the angle (α) defining the alignment between two consecutive boulders of 45 degrees and fixing 0.06 as the slope, the values of pool length and width (LP = 5 m WP = 6.4 m) were generated as well as the pool volume (5 m3) and the dissipated power (400 W m−3). Finally, the model domain was 55 m long with six target pools, one outfall pool (with double length) and eight rock-rows (the last one is the control row). The control row and the outfall pool are used to guarantee a steady flow. Due to the influences that the inlet and outlet conditions could have in the target pools, extra upstream and downstream pools were added in the ramp layout. A representation of the R3 layout domain can be seen in Figure 2 and the details of a pool and a rock-row dimensions are shown in Figure 3.
Figure 2. Rectangular boulder layout (R3) for 2D model domain from a top view.
Figure 2. Rectangular boulder layout (R3) for 2D model domain from a top view.
Water 14 03492 g002
Figure 3. Details of a pool and a rock-row geometry of rectangular boulder layout (R3).
Figure 3. Details of a pool and a rock-row geometry of rectangular boulder layout (R3).
Water 14 03492 g003
-
Double Rectangular boulders (DR3)
The wider the boulder, the more uphill flow surface area it will generate. To this end, a R3 case was reproduced, but the boulder diameter of the downiest positioned boulder in each row was duplicated (Db = 1.4 m). As result, from the seven corridors, the odd ones were wider (Wc), and these were the corridors where the uphill flows occurred (Figure 4).
Figure 4. Schematic diagram of boulders and corridors geometry of double rectangular boulder layout (DR3).
Figure 4. Schematic diagram of boulders and corridors geometry of double rectangular boulder layout (DR3).
Water 14 03492 g004
-
Double Rectangular boulders with a small barrier (DRb3)
The decrease in velocity caused by friction is widely known. This layout was designed to assess the effect of friction by incorporating small friction walls in the mid-boulders (corridors 3 and 5) of the DR3 arrangement, as show Figure 5. These barriers were 0.2 m width and 2 m long.
Figure 5. Details of friction walls position of double rectangular boulder with a small barrier layout (DRb3).
Figure 5. Details of friction walls position of double rectangular boulder with a small barrier layout (DRb3).
Water 14 03492 g005
-
Double Trapezoidal boulders (DT3)
This layout was configured in order to evaluate the effect of the boulder shape on the maximum velocity in the gaps. Among the works found in the literature, spherical and square-shaped obstacles were used [16,17,30,31]. The aim here was to investigate how the trapezoidal shape impacted the uphill flow areas as it maximizes the diameter downhill of the boulder, and then of the uphill flow areas. For this purpose, new boulders were designed with 1.4 m of diameter (Db) and 1.2 m of the shorter diameter (Db*) for the wider boulders, while the first row of each rock-row were 0.7 and 0.5 m, respectively (Figure 6).
Figure 6. Detail of trapezoidal boulder dimension of double trapezoidal boulder layout (DT3). * indicates the shorter base of the trapezoid.
Figure 6. Detail of trapezoidal boulder dimension of double trapezoidal boulder layout (DT3). * indicates the shorter base of the trapezoid.
Water 14 03492 g006
The different design parameters are listed in Table 1. A total of four configurations were conducted.

2.2. Numerical Model

The flow pattern in the pools is of great importance for guiding the fish through the fishway. Such flow generates complications because they create separation flow effects around obstacles with wake interactions. Therefore, the 3D flows appear to be dominant and they have been investigated by many authors [32,33,34]. However, an efficient 3D model necessitates using large adaptive meshes, making CPU time consuming. The flow field in several designs of vertical slot fishways (VSF) and rock-weir fishways were also studied experimentally and numerically with 2D models by Puertas et al. [35], Marriner [36] and Tran [22], among others. By comparing the numerical and experimental results, Cea [37] exposed that the depth-averaged shallow water equations, with a suitable turbulence model, could be used in order to calculate these variables in vertical slot fishways (VSF). Cea included in its analysis RANS 3D depth-averaged with k-ε roughness model. Fuentes-Perez [37] demonstrated that RANS 3D with k-ε slightly underestimated TKE with respect to LES (large eddy simulation method using the Smagorinsky turbulence model), but advocated the use of RANS to simulate larger spatial scales corresponding to the time-averaged flow, and LES in regions where a more detailed analysis was required. In our study, we considered that the flow behaviour in VSF and in uphill flow rock ramps presented similar hydraulic characteristics: (i) the inflow into the pool went through flow gaps without bottom orifices and without weirs; (ii) the gap was narrow in comparison to the width of the pool; (iii) and the energy dissipation in the pool was linked to the turbulent kinetic energy. These similarities allowed us to assume that the results of Cea and Fuentes are also valid for the type of fishway studied in this work. Tran [22] also opened up the possibility of using a depth-averaged shallow water model in nature-like fishways in order to investigate design options. This study could show that a 2D depth integrated model can help the designer when the rock ramp has less than 7% slope and a boulder concentration between 6–20%. Consequently, the 2D model, which is quite easy to use, appears to be an appropriate tool for considering special geometrical configuration and for adapting designs, and could lead to a more detailed examination (3D modelling) of a final design for nature-like fishways.
In this study, we used IBER 2.5.2 [28] (depth-averaged model with k-ε model turbulence) to obtain depths, velocity and TKE fields. It has been successfully used for a wide range of applications, including the modelling of river flow and quality [38], instream wood transport [39], geomorphic impacts of dam failure [40] or performance vertical slot fishways [41]. The model solves the depth-averaged shallow water equations coupled with a turbulence model. These equations were obtained after vertical integration of the three-dimensional Reynolds averaged Navier-Stokes (3D-RANS) equations over the water depth (see further details in Blade et al. [28]). For turbulence modelling, this work employed k-ε model from Rastogi and Rodi [42]. These equations are given by
h t + hU x x + hU y x = Ms hU x t + hU x 2 x + hU x U y y = gh Z x x + τ s , x ρ τ b , x ρ g ρ h 2 2 p x + 2 Ω sin λ U y + h τ xx e x + h τ xy e y + M x hU y t + hU x U y x + hU y 2 y = gh Z s y + τ s , y ρ τ b , y ρ g ρ h 2 2 p y 2 Ω sin λ U x + h τ xy e x + h τ yy e y + M y
where h is depth; Ux and Uy are the depth averaged horizontal velocities; g is the gravity acceleration; Z is the free layer elevation; τs is the free surface friction due to wind induced friction; τb is the bed friction; ρ is the water density; Ω is the earth’s rotation angular velocity; λ is the latitude of the studied point; τexx, τexy, and τeyy are the effective horizontal tangential stresses; and Ms, Mx, and My, respectively, are the terms of mass source/drain and momentum, which are used to model precipitation, infiltration and drainage.
Boundary conditions were applied to the inlet (discharge Q = 2 m3 s−1) and the outlet in critical/subcritical conditions, while the wall condition was employed for the bottom boulders and side walls (Manning coefficient equal to 0.02 s m−1/3). This coefficient is recommended by Zheng [10] for roughness particle sizes of 2 cm, which is the worst case scenario (lower roughness values implies higher velocities). The four ramp layouts were transformed into a regular calculation mesh, with a resolution equal to 0.05 m using the Geospatial Data Abstraction Library (GDAL) approach. The simulations were stopped as soon as the hydrodynamics reached the steady condition (the difference between the inlet and outlet flows is negligible).

2.3. Hydraulic Variables Monitoring

Changes in hydraulic conditions, as reflected in water velocity, turbulence characteristics, and momentum, cause the mayor cue fish to seek a migration pathway in rivers [43]. However, these variables need to be considered in order to design fishways that are easily passable.
The design and the dimensioning of these rock-ramp models were planned to ensure uphill flow areas. However, the different layouts provided diverse values of the hydraulic variables. The three variables involved in this study were: uphill component of velocity vector (vy); maximum velocity throughout the flow gap (vmax); and the turbulent kinetic energy (TKE).
The uphill component of velocity vector represents the “y” component of the velocity vector, where the “y” axe follows the longitudinal direction of the ramp. Positive values indicate uphill flow and negative values show downhill flow.
Iber estimates the turbulent kinetic energy as TKE = 1 2 ( ( u ¯ ) 2 + ( v ¯ ) 2 .
In addition to the control variables listed above, the depth (h) was also collected to ensure that the different model configurations did not result in drastic changes, which would potentially impact on the passability of the ramp. Although the simulations offered the values of these variables in each cell of the calculating mesh, their values were only relevant in certain considered zones for evaluation. Table 2 presents the zones of the ramp where their values were taken for each variable and the justification for this zoning.
Once the data were obtained, the differences were analysed. We assessed the difference between variable data found in the standard ramp (R3) and in the other layouts. Initially, for the R3 ramp analyse, we considered all pools (1–6). Then, we observed that some variables had significant differences, depending on the pool, because they were under the influence of boundary conditions at the end of the simulation. Therefore, we assumed only the stable pools for the analysis.
Statistical analyses were realized with the software IBM SPSS Statistics Base 25. The significance of differences among the sets (i.e., between pools or layouts) was contrasted through the nonparametric Kruskal-Wallis test, for two or more groups, respectively. When testing differences between more than two groups, we identified which pairs of groups were different by applying a post hoc pairwise comparison test with the Dunn-Bonferroni correction for multiple testing. Significance was set to p-value < 0.05.

3. Results

3.1. The Influence of Pool, Corridor and Rock-Row Position (R3)

We observed a relationship between the location of pools, the depth (h) and the TKE (Figure 7a,b). The Bonferroni test, with a p-value lower than 0.001, showed significant differences between the values found in pool 1 and the other pools. Lower depth and TKE values were reached in pool 1 in comparison the others. The differences in median were 13% and 42% for depth and TKE. The pattern of maximum velocity (Figure 7c) concluded that the values of this variable were independent of their row location.
The test allowed us to rule out pool 1 because the control variables between that pool and the others showed significant differences. The records of this pool were disregarded because the control variables may have been affected by instabilities related to the inflow boundary condition, variabilities that could mask the effect of the design modifications.
The analysis of the corridors (Figure 8) showed a clear difference in the values, according to their location. The uphill velocity revealed differences (p-value < 0.001) between corridors 1–7, 2-4-6, and 3–5. According to the values of the variables, it was possible to identify resting corridors (corridors 1 and 7), where velocities were close to 0; fast paths with high downhill velocities (median value of −0.86 m s−1); and intermediate corridors, where the median was −0.66 m s−1 (corridors 3 and 5). The following table (Table 3) shows the p-values when the Kruskal-Wallis test did not allow to reject the null hypothesis.

3.2. Influence of Designs in Water Depth

The test found significant differences (p-value < 0.001) in depth for each simulated design. The median values were: 0.546 m in R3 layout (interquartile range: 0.49–0.59 m); 0.539 m in DR3 (interquartile range: 0.49–0.59 m); 0.532 m and 0.526 m in DRb3 (interquartile range: 0.48–0.58 m); and DT3 (interquartile range: 0.48–0.58 m). Although statistical differences were found, the design effect on depth was not relevant for fish passability.

3.3. Influence of Designs in Uphill Velocity

We observed a robust relationship between the uphill velocities and the ramp designs. Figure 9 gives the contours of uphill velocity for three designs: R3 (a), DRb3 (b), and DT3 (c). In this case, three different regions could be distinguished: zones of high downhill velocities colored in red (vy< −0.5 m s−1); a secondary region (yellow) of low downhill velocities (−0.5 < vy < 0 m s−1); and a third area (green), with clear uphill velocities (vy > 0 m s1). Across the designs, the highest values of uphill velocity are found in corridors 1 and 7, near the contour of the pools. Under the protection of the double boulders, larger areas with uphill velocities are also found (Figure 9b,c) where corridors three and five are identified because of the recirculation pattern effect. In this sense, these areas are even larger when the wall-friction barrier are included in the design (DRb3). In order to compare the size occupied by each type of region, Table 4 was created. The data show twice the uphill velocity flow area in ramps with double boulder than with single boulder.
The numerical results show similar trends, although the pattern was slightly different for fast corridors (Figure 10). The double boulder (DR3) layout presented higher velocities in the y-axe than R3. In the intermediate corridors, vy increased from −0.65 to −0.42 m s−1, where these values are −0.05 and 0.02 in the resting corridors. Overall, the influence of the friction wall improved the functionality of the ramp (DRb3 vs R3). This is particularly relevant in the intermediate corridors, where the uphill velocity module decreased (vy has increased from −0.65 to −0.33 m s−1). In the DT3 design, the median vy reached its highest value, 0.04 m s−1, in the resting corridors. In the intermediate corridors, the shape of the boulders had the same effect as the boulder dimension, decreasing the downhill velocity modules (vy increased from −0.65 to −0.43 m s−1), while in fast corridors, the downhill velocity module was increased (vy decreased from −0.85 m s−1 in R3 to −1.08 in DT3).

3.4. Influence of Designs in Maximum Velocity

The DR3 simulation generated lower values of maximum velocity in the gaps: 1.65 m s−1, compared to 1.81 in R3 (Figure 11). The maximum velocity differences were also noticeable in the R3 layout, although the barrier had a negative impact because this difference is slightly lower (0.12 m s−1) than with DR3 design. The reduction in maximum velocity in the gaps is a great contribution of the DT3 design. The simulation using a trapezoidal shape for the boulders showed a median value of 1.52 m s−1, decreasing this value 0.29 m s−1 in relation to R3.

3.5. Influence of Designs in Turbulent Kinetic Energy

In relation to TKE, in intermediate corridors the median values in the R3 and DR3 layouts were 0.39 and 0.31 m2 s−2, respectively, and 0.18 and 0.12 m2 s−2 in resting zones. The influence of the friction walls also enhanced the turbulent values in the intermediate corridors, with a reduction of 0.11 m2 s−2 in the R3 design (Figure 12).
Similar to the uphill velocity, the distribution of TKE values also shows a well-defined distribution pattern (Figure 13). It is noteworthy that the double rectangular boulder with a barrier design was able to open lower-energy corridors in the center of the pool. In these intermediate corridors, the TKE decreased under 0.3 m2 s−2 and even areas appeared, under the boulder protection, with TKE values lower than 0.2 m2 s−2. It should be noted (Table 5) that the area with the lowest values of TKE in the resting corridors was increased by 51% in DRb3 compared to R3.

4. Discussion

This section discusses the key question: Is there any alternative design which can improve the passage efficiency of uphill flow rock ramp?
As previously mentioned, the flow pattern in the pools is of great importance for guiding the fishes through the fishway. Both water depth and velocity primarily affect the swimming costs of fishes, although the turbulence level has also been shown to increase them considerably. Clay [44] has suggested that, when designing a fishway, it is necessary to locate the areas of low velocities (and also low turbulences) and assess how they can influence fish behaviour, since these are the hydraulic variables involved in passage efficiency.
In this paper, four different uphill flow ramp configurations were tested to show which design elements can positively impact the functionality of this type of ramp.
It is a prerequisite for all types of fishways that the depth must be sufficient for the fish to swim without difficulty, fully submerged, including the dorsal fin. The recommendation is for a depth of flow greater than, or equal to, two times the largest fish’s body depth [4]. In the designs considered in this study, the median depth of the pools took the lowest value for DT3 (0.526 m) and the highest value for R3 (0.546 m). With this range of depths between models, transit was guaranteed for body depths from 0.26 m to 0.27 m. This small variation clearly showed that the design effect on depth was not relevant for fish passability.
Clay [44] has suggested that, when designing a fishway, it is necessary to locate the areas of low velocities and also low turbulences and assess the influence on the fish behaviour, since these hydraulic variables were critically involved in the passage efficiency. Regarding fishway design, the velocity in the gap (vmax) plays an important role, because it must present values lower than the sprint speed of the fish and with a passage time lower than the time the fish can maintain this velocity [45]. In all of the designs within this study, vmax was reduced with respect to the original R3 layout, getting the highest reduction (16%) in DT3. The absolute value of this variable should be interpreted in terms of the swimming capabilities of the target species. DT3 presented a median value of 1.52 m s−1 (interquartile range: 1.15–1.81 m s−1). Sanz-Ronda [45] studied two cyprinid species endemic to Spain and found total passability for lengths of 2 m and 4 m, with velocities of 2 m s−1 and 1.5 m s−1, respectively. Bearing in mind that the maximum length of vmax in this study was 1.3 m (length between diagonally opposite ends in the gap), this variable would not suppose any limitation for the passage of these species. These contributions allowed us to conclude that the shape of the boulders contributes to reducing the vmax in the gap.
The three different uphill flow ramp configurations, with regards to R3, including increased boulder size (DR3) and the incorporation of friction walls (DRb3), showed the influence of these geometrical aspects in the magnitude of turbulent kinetic energy. Turbulence and associated air entrainment is linked to the energy dissipation that occurs in the pools and is considered advantageous in the fish passage [23,46]. High turbulence may decrease swimming performance [47] and increase the energetic cost of swimming performance [46,48]. Fish have also exhibited preferences for low turbulence regions within fishways [49,50,51] and, in general, high turbulence levels seem to negatively affect fishway passage [52].
To evaluate this flow characteristic in one-dimensional hydraulic models, variables that measure it indirectly have been used: volumetric power dissipation (VPD), energy dissipation rate, or energy dissipation factor. These variables measure the power of the flow entering the pool, divided by the pool volume, W m−3. The maximum recommended values vary according to the species or group of species: for cyprinids, a classical reference is 150 W m−3 [53] and for salmonids, 200 W m−3 [11], although other authors [54] raised this threshold to 240 W m−3. However, the VPD assumes an average value of dissipated turbulence for an entire pool, omitting the distribution of turbulence [37], which can vary significantly across the pool.
TKE is a direct measure of the characteristics of the turbulence of the flow [55], quantifying the kinetic energy of the fluctuations of the components of the velocity vector. This energy is extracted from the mean flow due to shear between the mean and fluctuating velocities and gradients in the mean velocity field [56]. TKE has shown negative correlation with the critical swimming speed of multiple fish species [47] and with the time needed for the fish to ascend a fishway [51].
Koziol [57], in a laboratory channel with a slope of 0.5 per thousand and relative depth 0.43 m, measured values up to 0.2 m2 s−2. Stone [58], in a gravel river (mean diameter 10.5 cm; flow 3.1 m3 s−1; slope 0.0012), measured values up to 0.05 m2 s−2 in the pool, 0.12 m2 s−2 in the run, and 0.33 m2 s−2 in the riffle. Puzdrowska [34], in laboratory model of ramps with aligned boulders, obtained the field of TKE values, with maximum values of 0.6 m2 s−2 and mean stream´s flow area TKE does not exceed the value of 0.1 m2 s−2. Guiny [48] studied the passage of atlantic salmon with TKE 0.4–1.2 m2 s−2 in a pool with an orifice fishway. Silva [51] evaluated the passage of a cyprinid species, the Iberian barbel (Luciobarbus bocagei Steindachner 1864), in an experimental channel and found that the highest number of passages occurred at low TKE values (<0.05 m2 s−2, with high transit times variability). At 0.05 < TKE < 0.3 m2 s−2 fewer fish passed, but transit time and variability were considerably reduced. The study recorded a passage with TKE = 0.36 m2 s−2 and one of the lowest transit times. Marrinier [59] has used TKE=0.05 m2 s−2 as a “low” threshold, referring to the results of Silva [51]. This value should not be interpreted as a threshold limit value but as preference value for passage.
Among the three designs, and in comparison to R3, DRb3 showed the greatest reduction in TKE when resting (median value TKE 0.13 m2 s−2) and intermediate corridors (median value TKE 0.28 m2 s−2) were considered together (Figure 12). Consequently, the flow area under this condition was 75%. This design also had the largest increase in surface area, with TKE < 0.2 m2 s−2 (36.09%, Table 5), and the largest surface area, with TKE ≤ 0.05 m2 s−2 (9.37%). These values were cited by Silva and Marrinier [51,59] as the preferred corridor. DRb3 was also the model with the lowest surface areas, with TKE above 0.6 (13%).
To summarise, this paper aimed to gain insights on potential alternative designs, able to outperform the passage efficiency of uphill flow rock ramp. Our results showed that:
(a)
in terms of uphill velocity and maximum velocity, the design with the greatest impact on the functionality of this type of ramp is the trapezoidal shaped double boulder design (DT3), as it maximized the uphill velocity in the resting corridors and, at the same time, has the maximum speed reduction in the row gaps. That is in accordance with Miranda et al.’s [17] statement: the obstacle shape is an important modifiable parameter for fishways designing.
(b)
in terms of the TKE, the most evident improvements were achieved in the resting corridor, with the double boulder layout (DR3), and with double boulder and friction wall in the intermediate corridors (DRb3).
Finally, these results provide integrated and tangible results that should be valuable to designing improved fish passes, aiming to maximize passability. Our results demonstrated that further investigation into the combined effect of the trapezoidal shape of the boulder, together with the friction barrier, is clearly needed due to their importance for increasing uphill velocities and reducing the TKE value, both in resting and intermediate corridors.
The comparison of the results obtained with the modelling of different designs provides useful information to be considered in the design of uphill flow ramps. However, it must be complemented with data that allow the evaluation of the fish passage performance (number of fish passing; passage time; number of attempts). Additional research, studying the biological response to changes in hydraulic variables, is needed to guarantee the improvement of future designs.

Author Contributions

Conceptualization, J.A.F.-Y. and C.M.S.-M.; methodology, J.A.F.-Y. and L.C.-D.; software, L.C.-D.; validation, J.A.F.-Y., C.M.S.-M. and L.C.-D.; formal analysis, L.C.-D.; investigation, J.A.F.-Y. and L.C.-D.; resources, L.C.-D.; data curation, L.C.-D.; writing—original draft preparation, L.C.-D.; writing—review and editing, J.A.F.-Y., C.M.S.-M. and L.C.-D.; supervision, J.A.F.-Y. and C.M.S.-M.; project administration, J.A.F.-Y. 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 available upon reasonable request to the corresponding author.

Acknowledgments

We warmly thank Pepa Aroca for her technical support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. McIntyre, P.B.; Liermann, C.R.; Childress, E.; Hamann, E.J.; Hogan, J.D.; Januchowski-Hartley, S.R.; Koning, A.A.; Neeson, T.M.; Oele, D.L.; Pracheil, B.M. Conservation of Migratory Fishes in Freshwater Ecosystems. In Conservation of Freshwater Fishes; Cambridge University Press: Cambridge, UK, 2015. [Google Scholar]
  2. Cooper, A.R.; Infante, D.M.; Daniel, W.M.; Wehrly, K.E.; Wang, L.; Brenden, T.O. Assessment of Dam Effects on Streams and Fish Assemblages of the Conterminous USA. Sci. Total Environ. 2017, 586, 879–889. [Google Scholar] [CrossRef] [PubMed]
  3. Fuller, M.R.; Doyle, M.W.; Strayer, D.L. Causes and Consequences of Habitat Fragmentation in River Networks. Ann. N. Y. Acad. Sci. 2015, 1355, 31–51. [Google Scholar] [CrossRef] [PubMed]
  4. USFWS (U.S. Fish and Wildlife Service). Fish Passage Engineering Design Criteria; Report; USFWS: Hadley, MA, USA, 2017; p. 248.
  5. Sanz-Ronda, F.J.; Bravo-Córdoba, F.J.; Fuentes-Pérez, J.F.; Ruiz-Legazpi, J.; García-Vega, A.; Ramos-González, N.; Salgado-González, V.M.; Martínez de Azagra Paredes, A. Pasos Para Peces: Escalas y Otros Dispositivos de Paso. Notas técnicas del CIREF 2013, 7, 17. [Google Scholar]
  6. Franklin, P.; Gee, E.; Baker, C.; Bowie, S. New Zealand Fish Passage Guidelines for Structures up to 4 Meters; NIWA: Hamilton, New Zeland, 2018; p. 211. [Google Scholar]
  7. Silva, A.T.; Lucas, M.C.; Castro-Santos, T.; Katopodis, C.; Baumgartner, L.J.; Thiem, J.D.; Aarestrup, K.; Pompeu, P.S.; O’Brien, G.C.; Braun, D.C.; et al. The Future of Fish Passage Science, Engineering, and Practice. Fish Fish. 2018, 19, 340–362. [Google Scholar] [CrossRef] [Green Version]
  8. Cooke, S.J.; Hinch, S.G. Improving the Reliability of Fishway Attraction and Passage Efficiency Estimates to Inform Fishway Engineering, Science, and Practice. Ecol. Eng. 2013, 58, 123–132. [Google Scholar] [CrossRef]
  9. Katopodis, C.; Williams, J.G. The Development of Fish Passage Research in a Historical Context. Ecol. Eng. 2012, 48, 8–18. [Google Scholar] [CrossRef]
  10. Zheng, T.; Niu, Z.; Sun, S.; Shi, J.; Liu, H.; Li, G. Comparative Study on the Hydraulic Characteristics of Nature-like Fishways. Water 2020, 12, 955. [Google Scholar] [CrossRef] [Green Version]
  11. DVWK. Fish Passes: Design, Dimensions and Monitoring; FAO/DVWK: Rome, Italy, 2002; p. 136. [Google Scholar]
  12. Baudoin, J.-M.; Burgun, V.; Chanseau, M.; Larinier, M.; Ovidio, M.; Sremski, W.; Steinbach, P.; Voegtle, B. Assessing the Passage of Obstacles by Fish. Concepts, Design and Application; French National Agency for Water and Aquatic Environments: Paris, France, 2015; p. 200.
  13. Bretón, F.; Baki, A.B.M.; Link, O.; Zhu, D.Z.; Rajaratnam, N. Flow in Nature-like Fishway and Its Relation to Fish Behaviour. Can. J. Civ. Eng. 2013, 40, 567–573. [Google Scholar] [CrossRef]
  14. Turek, J.; Haro, A.; Towler, B. Federal Interagency Nature-like Fishway Passage Design Guidelines for Atlantic Coast Diadromous Fishes. Interag. Tech. Memo. 2016, 52, 47. [Google Scholar]
  15. Bravo-Córdoba, F.J.; Fuentes-Pérez, J.F.; Valbuena-Castro, J.; de Azagra-Paredes, A.M.; Sanz-Ronda, F.J. Turning Pools in Stepped Fishways: Biological Assessment via Fish Response and CFD Models. Water 2021, 13, 1186. [Google Scholar] [CrossRef]
  16. Golpira, A.; Baki, A.B.M.; Zhu, D.Z. Higher-Order Velocity Moments, Turbulence Scales and Energy Dissipation Rate around a Boulder in a Rock-Ramp Fish Passage. Sustainability 2020, 12, 5385. [Google Scholar] [CrossRef]
  17. Miranda, F.C.; Cassan, L.; Laurens, P.; Tran, T.D. Study of a Rock-Ramp Fish Pass with Staggered Emergent Square Obstacles. Water 2021, 13, 1175. [Google Scholar] [CrossRef]
  18. Mao, X.; Huang, S.; Xu, Y.; Cheng, Z.; Zhang, J. The Hydraulics of Nature-like Fishways. Stavební Obz.-Civ. Eng. J. 2021, 30. [Google Scholar] [CrossRef]
  19. Abe, K.; Hayashida, K.; Kayaba, Y. Functional Evaluation of Full-Width Rock-Ramp Fish Passage Using Flow Tracers and Drones. J. Japan Soc. Civ. Eng. Ser. G (Environ. Res.) 2021, 77, II_53–II_58. [Google Scholar] [CrossRef]
  20. Cassan, L.; Laurens, P. Design of Emergent and Submerged Rock-Ramp Fish Passes. Knowl. Manag. Aquat. Ecosyst. 2016, 417, 45. [Google Scholar] [CrossRef] [Green Version]
  21. Mooney, D.D.M.; Holmquist-Johnson, C.L.; Broderick, S. Rock Ramp Design Guidelines. US Dep. Inter. 2007, 110, 102. [Google Scholar]
  22. Tran, T.D.; Chorda, J.; Laurens, P.; Cassan, L. Modelling Nature-like Fishway Flow around Unsubmerged Obstacles Using a 2D Shallow Water Model. Environ. Fluid Mech. 2016, 16, 413–428. [Google Scholar] [CrossRef] [Green Version]
  23. Towler, B.; Mulligan, K.; Haro, A. Derivation and Application of the Energy Dissipation Factor in the Design of Fishways. Ecol. Eng. 2015, 83, 208–217. [Google Scholar] [CrossRef]
  24. Baki, A.B.M.; Zhu, D.Z.; Harwood, A.; Lewis, A.; Healey, K. Rock-Weir Fishway I: Flow Regimes and Hydraulic Characteristics. J. Ecohydraulics 2017, 2, 122–141. [Google Scholar] [CrossRef]
  25. Fernández-Yuste, J.A.; Martínez, C. Uphill Flow Rock Ramp Design; Confederación hidrográfica del Duero: Valladolid, Spain, 2021; p. 130. [Google Scholar]
  26. Fuentes-Pérez, J.F.; Sanz-Ronda, F.J.; Paredes, A.M.D.A.; García-Vega, A. Modeling Water-Depth Distribution in Vertical-Slot Fishways under Uniform and Nonuniform Scenarios. J. Hydraul. Eng. 2014, 140, 1–5. [Google Scholar] [CrossRef] [Green Version]
  27. Fuentes-Pérez, J.F.; Sanz-Ronda, F.J.; de Azagra, A.M.; García-Vega, A. Non-Uniform Hydraulic Behavior of Pool-Weir Fishways: A Tool to Optimize Its Design and Performance. Ecol. Eng. 2016, 86, 5–12. [Google Scholar] [CrossRef]
  28. Bladé, E.; Cea, L.; Corestein, G.; Escolano, E.; Puertas, J.; Vázquez-Cendón, E.; Dolz, J.; Coll, A. Iber: Herramienta de Simulación Numérica Del Flujo En Ríos. Rev. Int. Métodos Numéricos para Cálculo y Diseño en Ing. 2014, 30, 1–10. [Google Scholar] [CrossRef] [Green Version]
  29. Fernández-Yuste, J.A.; Martínez, C. Upflow Ramps. Asistente Para El Cálculo de Rampas de Flujo Ascendente. Available online: https://rampas.chduero.es/ (accessed on 7 October 2022).
  30. Baki, A.B.M.; Zhu, D.Z.; Rajaratnam, N. Flow Simulation in a Rock-Ramp Fish Pass. J. Hydraul. Eng. 2016, 142, 1–12. [Google Scholar] [CrossRef]
  31. Baki, A.B.M.; Zhu, D.Z.; Rajaratnam, N. Mean Flow Characteristics in a Rock-Ramp-Type Fish Pass. J. Hydraul. Eng. 2014, 140, 156–168. [Google Scholar] [CrossRef]
  32. Hinterberger, C.; Fröhlich, J.; Rodi, W. Three-Dimensional and Depth-Averaged Large-Eddy Simulations of Some Shallow Water Flows. J. Hydraul. Eng. 2007, 133, 857–872. [Google Scholar] [CrossRef]
  33. Chang, K.; Constantinescu, G. Numerical Investigation of Flow and Turbulence Structure through and around a Circular Array of Rigid Cylinders. J. Fluid Mech. 2015, 776, 161–199. [Google Scholar] [CrossRef]
  34. Puzdrowska, M.; Heese, T. Turbulent Kinetic Energy in Fish Passes of Various Types of Construction. WIT Trans. Ecol. Environ. 2019, 234, 81–90. [Google Scholar] [CrossRef] [Green Version]
  35. 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]
  36. Marriner, B.A.; Baki, A.B.M.; Zhu, D.Z.; Thiem, J.D.; Cooke, S.J.; Katopodis, C. Field and Numerical Assessment of Turning Pool Hydraulics in a Vertical Slot Fishway. Ecol. Eng. 2014, 63, 88–101. [Google Scholar] [CrossRef]
  37. Fuentes-Pérez, J.F.; Tuhtan, J.A.; Branco, P.; Eckert, M.; Ferreira, M.T.; Kruusmaa, M. A 3D Data-Driven Approach to Study the Hydrodynamic Preferences of Fish in Fishways. Int. Symp. Ecohydraulics 2018, 4, 1–4. [Google Scholar]
  38. Sañudo, E.; Cea, L.; Puertas, J. Modelling Pluvial Flooding in Urban Areas Coupling the Models Iber and SWMM. Water 2020, 12, 2647. [Google Scholar] [CrossRef]
  39. Ruiz-Villanueva, V.; Gamberini, C.; Bladé, E.; Stoffel, M.; Bertoldi, W. Numerical Modeling of Instream Wood Transport, Deposition, and Accumulation in Braided Morphologies Under Unsteady Conditions: Sensitivity and High-Resolution Quantitative Model Validation. Water Resour. Res. 2020, 56, e2019WR026221. [Google Scholar] [CrossRef]
  40. Álvarez, M.; Puertas, J.; Peña, E.; Bermúdez, M. Two-Dimensional Dam-Break Flood Analysis in Data-Scarce Regions: The Case Study of Chipembe Dam, Mozambique. Water 2017, 9, 432. [Google Scholar] [CrossRef] [Green Version]
  41. Cea, L.; Pena, L.; Puertas, J.; Asce, A.; Vázquez-Cendón, M.E.; Peña, E. Application of Several Depth-Averaged Turbulence Models to Simulate Flow in Vertical Slot Fishways. J. Hydraul. Eng. 2007, 133, 160–172. [Google Scholar] [CrossRef]
  42. Rastogi, A.K.; Rodi, W. Predictions of Heat and Mass Transfer in Open Channels. ASCE J Hydraul Div 1978, 104, 397–420. [Google Scholar] [CrossRef]
  43. Williams, J.G.; Armstrong, G.; Katopodis, C.; Larinier, M.; Travade, F. Thinking like a Fish: A Key Ingredient for Development of Effective Fish Passage Facilities at River Obstructions. River Res. Appl. 2012, 28, 407–417. [Google Scholar] [CrossRef] [Green Version]
  44. Clay, C.H. Design of Fishways and Other Fish Facilities, 2nd ed.; CRS Press: Boca Raton, FL, USA, 2019; p. 256. [Google Scholar]
  45. Sanz-Ronda, F.J.; Ruiz-Legazpi, J.; Bravo-Córdoba, F.J.; Makrakis, S.; Castro-Santos, T. Sprinting Performance of Two Iberian Fish: Luciobarbus Bocagei and Pseudochondrostoma Duriense in an Open Channel Flume. Ecol. Eng. 2015, 83, 61–70. [Google Scholar] [CrossRef]
  46. Enders, E.C.; Boisclair, D.; Roy, A.G. A Model of Total Swimming Costs in Turbulent Flow for Juvenile Atlantic Salmon (Salmo Salar). Can. J. Fish. Aquat. Sci. 2005, 62, 1079–1089. [Google Scholar] [CrossRef] [Green Version]
  47. Lupandin, A.I. Effect of Flow Turbulence on Swimming Speed of Fish. Biol. Bull. 2005, 32, 461–466. [Google Scholar] [CrossRef]
  48. Guiny, E.; Ervine, D.A.; Armstrong, J.D. Hydraulic and Biological Aspects of Fish Passes for Atlantic Salmon. J. Hydraul. Eng. 2005, 131, 542–553. [Google Scholar] [CrossRef] [Green Version]
  49. Duarte, B.A.D.F.; Ramos, I.C.R.; Santos, H.D.A. Reynolds Shear-Stress and Velocity: Positive Biological Response of Neotropical Fishes to Hydraulic Parameters in a Vertical Slot Fishway. Neotrop. Ichthyol. 2012, 10, 813–819. [Google Scholar] [CrossRef] [Green Version]
  50. 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] [Green Version]
  51. Silva, A.T.; Santos, J.M.; Ferreira, M.T.; Pinheiro, A.N.; Katopodis, C. Effects of Water Velocity and Turbulence on the Behaviour of Iberian Barbel (Luciobarbus Bocagei, Steindachner 1864) in an Experimental Pool-Type Fishway. River Res. Appl. 2011, 27, 360–373. [Google Scholar] [CrossRef]
  52. Mallen-Cooper, M.; Zampatti, B.; Stuart, I.; Baumgartner, L. Innovative Fishways-Manipulating Turbulence in the Vertical-Slot Design to Improve Performance and Reduce Cost; Murray Darling Basin Commission, Fishway Consulting Services: Sidney, Australia, 2008; p. 19. [Google Scholar]
  53. Larinier, M. Pool fishways, pre-barrages and natural bypass channels. Bull. Français la Pêche la Piscic. 2002, 364, 54–82. [Google Scholar] [CrossRef]
  54. Maine Department of Transportation. Waterway and Wildlife Crossing Policy and Design Guide, 3rd ed.; Maine DOT: Augusta, ME, USA, 2008; p. 128.
  55. Odeh, M. Discharge Rating Equation and Hydraulic Characteristics of Standard Denil Fishways. J. Hydraul. Eng. 2003, 129, 341–348. [Google Scholar] [CrossRef]
  56. Reynolds, A.J. Turbulent Flows in Engineering; John Wiley & Sons: Hoboken, NJ, USA, 1974; p. 462. [Google Scholar]
  57. Kozioł, A. Turbulent Kinetic Energy of Water in a Compound Channel. Ann. Warsaw Univ. Life Sci.-SGGW. L. Reclam. 2012, 43, 193–205. [Google Scholar] [CrossRef]
  58. Stone, M.C.; Hotchkiss, R.H. Turbulence Descriptions in Two Cobble-Bed River Reaches. J. Hydraul. Eng. 2007, 133, 1367–1378. [Google Scholar] [CrossRef] [Green Version]
  59. Marriner, B.A.; Baki, A.B.M.; 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]
Figure 1. Illustration of resulting uphill flows due to zig-zag boulder distribution.
Figure 1. Illustration of resulting uphill flows due to zig-zag boulder distribution.
Water 14 03492 g001
Figure 7. Boxplot of ensemble results of numerical simulation for rectangular boulder layout (R3). (a) water depth; (b) turbulent kinetic energy; and (c) maximum velocity in flow gaps. The boxplot represents the minimum, maximum, median, first quartile (25th percentile), third quartile (75th percentile) and outliers.
Figure 7. Boxplot of ensemble results of numerical simulation for rectangular boulder layout (R3). (a) water depth; (b) turbulent kinetic energy; and (c) maximum velocity in flow gaps. The boxplot represents the minimum, maximum, median, first quartile (25th percentile), third quartile (75th percentile) and outliers.
Water 14 03492 g007
Figure 8. Boxplot of uphill velocities by corridors when rectangular boulder layout (R3) was simulated. Positive values indicate uphill flow and negative values show downhill flow.
Figure 8. Boxplot of uphill velocities by corridors when rectangular boulder layout (R3) was simulated. Positive values indicate uphill flow and negative values show downhill flow.
Water 14 03492 g008
Figure 9. Uphill component of velocity vector in pool 3 and 4 for rectangular boulder layout (a), double rectangular boulder with a small barrier layout (b) and double Trapezoidal boulders (c).
Figure 9. Uphill component of velocity vector in pool 3 and 4 for rectangular boulder layout (a), double rectangular boulder with a small barrier layout (b) and double Trapezoidal boulders (c).
Water 14 03492 g009
Figure 10. Median values of uphill velocity component by corridor type: fast (FAST), resting (REST) and intermediate (INT) and design layout: rectangular boulder (R3), double rectangular boulder (DR3), double rectangular boulder with a small barrier (DRb3) and double trapezoidal boulder (DT3). Positive values indicate uphill flow and negative values show downhill flow.
Figure 10. Median values of uphill velocity component by corridor type: fast (FAST), resting (REST) and intermediate (INT) and design layout: rectangular boulder (R3), double rectangular boulder (DR3), double rectangular boulder with a small barrier (DRb3) and double trapezoidal boulder (DT3). Positive values indicate uphill flow and negative values show downhill flow.
Water 14 03492 g010
Figure 11. Histogram of maximum velocity by design layout: rectangular boulder (R3), double rectangular boulder (DR3), double rectangular boulder with a small barrier (DRb3) and double trapezoidal boulder (DT3). The red line represents the median values.
Figure 11. Histogram of maximum velocity by design layout: rectangular boulder (R3), double rectangular boulder (DR3), double rectangular boulder with a small barrier (DRb3) and double trapezoidal boulder (DT3). The red line represents the median values.
Water 14 03492 g011
Figure 12. Median values of turbulent kinetic energy by corridor type: fast (FAST), resting (REST) and intermediate (INT) and design layout: rectangular boulder (R3), double rectangular boulder (DR3), double rectangular boulder with a small barrier (DRb3) and double trapezoidal boulder (DT3).
Figure 12. Median values of turbulent kinetic energy by corridor type: fast (FAST), resting (REST) and intermediate (INT) and design layout: rectangular boulder (R3), double rectangular boulder (DR3), double rectangular boulder with a small barrier (DRb3) and double trapezoidal boulder (DT3).
Water 14 03492 g012
Figure 13. Turbulent kinetic energy in pool 3 and 4 for rectangular boulder layout (a), double rectangular boulder with a small barrier layout (b).
Figure 13. Turbulent kinetic energy in pool 3 and 4 for rectangular boulder layout (a), double rectangular boulder with a small barrier layout (b).
Water 14 03492 g013
Table 1. Layout parameters of every comparison scheme: rectangular boulder (R3); double rectangular boulder (DR3); double rectangular boulder with small barriers (DRb3) and double trapezoidal boulder (DT3). * The number indicate the corridor position.
Table 1. Layout parameters of every comparison scheme: rectangular boulder (R3); double rectangular boulder (DR3); double rectangular boulder with small barriers (DRb3) and double trapezoidal boulder (DT3). * The number indicate the corridor position.
LayoutBoulder Diameter
(m)
Shorter Boulder Diameter
(m)
Boulder Width
(m)
Geometric Gap Width
(m)
Ramp Width
(m)
Corridor Width
1, 7 *
(m)
Corridor Width 2, 4, 6 *
(m)
Corridor Width
3, 5 *
(m)
R30.70.70.50.356.40.8250.950.95
DR30.7/1.40.7/1.40.50.359.21.5250.951.65
DRb30.7/1.40.7/1.40.50.359.21.5250.951.65
DT30.7/1.40.5/1.20.50.358.5751.50.71.7
Table 2. Primary details of the zones where the hydraulic variables were taken.
Table 2. Primary details of the zones where the hydraulic variables were taken.
Hydraulic VariableZoneCell Numbers in R3Justification
Depth
h
(m)
Pool
Water 14 03492 i001
Pool = 1820
Totalramp = 10,920
Depth was measured in every target pool because the fish must be guaranteed a minimum depth that allows it to move easily and safely.
Uphill component of velocity vector
vy
(m s−1)
Corridor
Water 14 03492 i002
Pool
Corridor 1, 7 = 351
Corridor 2, 4, 6 = 351
Corridor 3, 5 = 390
Totalramp = 15,210
The velocity on the flow direction (y-axis) was measured in each of the seven corridors into which every pool was divided (positive values indicate uphill flow and negative values show downhill flow).
As a consequence of the zigzag distribution of the boulders in the row, uphill secondary flows were generated in the corridors 3 and 5 where slow, null or even uphill flow velocities may have been generated.
Corridors 1 and 7 were affected not only by the uphill behaviour but also by the roughness of the contour (higher uphill flows and lower TKE values).
The flow through the spillways generated corridors (2, 4 and 6) where the downhill velocity and the TKE were high (fast corridor).
Maximum velocity
vmax
(m s−1)
Flow gap
Water 14 03492 i003
Rock-row = 294
Totalramp = 2058
The critical conditions of fish passability were given by the velocity in
the flow gap.
Turbulent kinetic energy
TKE
(m2 s−2)
Corridor
Water 14 03492 i004
Pool
Corridor 1, 7 = 351
Corridor 2, 4, 6 = 351
Corridor 3, 5 = 390
Totalramp = 15,210
Same justification as uphill velocity vector.
Table 3. p-Values from Kruskal-Wallis test when uphill velocities from the different corridors are compared.
Table 3. p-Values from Kruskal-Wallis test when uphill velocities from the different corridors are compared.
Corridor 4Corridor 5Corridor 6Corridor 7
Corridor 1 0.582
Corridor 20.238 0.479
Corridor 3 0.064
Corridor 4 0.059
Table 4. Percentage of flow area occupied by each uphill velocity range and layout type: rectangular boulder layout (R3), double rectangular boulder (DR3), double rectangular boulder with a small barrier layout (DRb3) and double trapezoidal boulders (DT3). The percentage has been calculated respect of the pool area.
Table 4. Percentage of flow area occupied by each uphill velocity range and layout type: rectangular boulder layout (R3), double rectangular boulder (DR3), double rectangular boulder with a small barrier layout (DRb3) and double trapezoidal boulders (DT3). The percentage has been calculated respect of the pool area.
vy (m s−1)R3DR3DRb3DT3
≤−0.563.0538.6942.5142.94
−0.5–025.5040.9634.0333.14
≥011.4520.3523.4623.92
Table 5. Percentage of flow area occupied by each turbulent kinetic energy range and layout type: rectangular boulder layout (R3), double rectangular boulder (DR3), double rectangular boulder with a small barrier layout (DRb3) and double trapezoidal boulders (DT3). The percentage has been calculated respect of the pool area.
Table 5. Percentage of flow area occupied by each turbulent kinetic energy range and layout type: rectangular boulder layout (R3), double rectangular boulder (DR3), double rectangular boulder with a small barrier layout (DRb3) and double trapezoidal boulders (DT3). The percentage has been calculated respect of the pool area.
TKE (m2 s−2)R3DR3DRb3DT3
≤0.217.6529.8736.0926.96
0.2–0.442.5644.2838.9843.63
0.4–0.617.2011.9912.0916.00
≥0.622.5913.8612.8413.40
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Carrero-Díez, L.; Martínez Santa-María, C.; Fernández-Yuste, J.A. Uphill Flow Rock Ramps. How the Design Impacts Their Functionality. Water 2022, 14, 3492. https://doi.org/10.3390/w14213492

AMA Style

Carrero-Díez L, Martínez Santa-María C, Fernández-Yuste JA. Uphill Flow Rock Ramps. How the Design Impacts Their Functionality. Water. 2022; 14(21):3492. https://doi.org/10.3390/w14213492

Chicago/Turabian Style

Carrero-Díez, Leticia, Carolina Martínez Santa-María, and J. Anastasio Fernández-Yuste. 2022. "Uphill Flow Rock Ramps. How the Design Impacts Their Functionality" Water 14, no. 21: 3492. https://doi.org/10.3390/w14213492

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