The effect of the end sill shape of steps for two-phase flow over stepped spillways was investigated. A quantitative assessment approach has been applied. Experimental and numerical tests were used for stepped spillways with different step configurations. The computational fluid dynamics (CFD) method with a two-dimensional (2D) flow model, including volume of fluid (VOF) and k-ε models was applied to conduct a critical analysis in different flow conditions. Step heights hs were 0.03 m and 0.05 m, and the number of steps Ns were 10 and 6, respectively. The step configurations assessed were Step models, Sill models, and Curve models with angle θ = 26.6°. The developed 2D model investigated the flow patterns, velocity distribution, and pressure for several step configurations. For validation, the correlation coefficient (R), mean absolute percentage error (MAPE), and root mean squared error (RMSE) were tested. The results showed negative pressure at the flat steps at Nappe flow, whilst negative pressure did not appear in the Curve model. MAPE and RMSE were 4.32 and 0.21 respectively, with R =0.996. In the CFD results, velocity and the pressure distributions were compared with the experiments and showed a correlation within a (±6%) tolerance. The Curve models were the best in terms of the flow characteristics.

  • Flow regimes were simulated for different flow conditions along the stepped spillway.

  • The velocity distribution was simulated at the first step, mid-step, and last step.

  • The pressure distribution was simulated for nappe and skimming flow at the first step, mid-step, and last step.

  • The water surface profile with flow streamlines were simulated along the stepped spillway.

Graphical Abstract

Graphical Abstract

Symbols used are as follows:

hs

step height (m).

ls

step length (m).

Lc

length of the crest (m).

Ns

Number of steps (unitless).

P

pressure (pa).

Q

discharge (m3/sec).

q

discharge per unit width ((m3/sec)/m).

Vc

velocity at critical depth (m/sec).

Vo

mean flow velocity at upstream (m/sec).

V1

the downstream velocity (m/sec).

W

the width of flume (m).

X

the distance (m).

yc

critical depth of the flow (m).

γ

water density (KN/m3).

θ

the angle of the slope (degree).

Stepped spillways are significant hydraulic structures. Flow characteristic, pressure, velocity distribution and step geometry are factors affecting stepped spillway design and safety (Boes & Hager 2003). Chinnarasri & Wongwises (2006) presented flow characteristics including flow regimes for the flow on a stepped spillway. Moreover, these factors were affecting the energy dissipation for different step shapes: a horizontal step, inclined step, and step with end sill. The outcomes showed that energy dissipation was lower at skimming flow than for a nappe flow regime. The step with end sills was the most efficient step geometry in terms of energy dissipation. The effect of the number of steps was evident since the energy dissipation increased with the number of steps. Also, it looks like stepped spillways with a modified step configuration might be a successful solution that can be applied (Jahad et al. 2018). Three different kinds of flow regimes can be seen in the flow over a stepped spillway depending on the effecting factors: nappe flow for low discharges, skimming flow for high discharges, and transition flow (Wan et al. 2019; Kaouachi et al. 2021). Furthermore, it is important to investigate other influences such as pressure distribution and velocity distribution across the step faces. Boes & Hager (2003) and Ghaderi et al. (2020) explained that hydrodynamic pressure and flow velocity are two of the significant factors that affect the design and safety of a stepped spillway. Takahashi & Ohtsu (2012) stated that as the impact region increases the relative critical depth (yc/hs) decreases whereas the location with the highest pressure on a sloping chute moves toward the inner step edge (x/ls increases). According to Frizell et al. (2013) two zones (vertical step face and horizontal step face) have drawn attention in all the previous investigations of pressure distribution over a stepped spillway.

At present, one of the most popular methods used to successfully apply hydrodynamics to hydraulic structures is the computational fluid dynamic (CFD) method. Chow (1979) noted that CFD is the best choice to solve the problem because of the number of assumptions that may be utilized numerically to obtain CFD outcomes. Using CFD modelling incorporated with free surface effects has been reported in the last decade. The models typically employ Reynolds-averaged Navier–Stokes (RANS) equations coupled with turbulence closure solved numerically on a computational grid. Biringen & Chow (2011), Dastgheib et al. (2012), Castillo et al. (2014), Shahheydari et al. (2015), Peng (2015), Hamedi & Ketabdar (2016), Daneshfaraz et al. (2020), Kaouachi et al. (2021) and Lebdiri et al. (2022) modelled flow characteristics over stepped spillways. Predictions of the velocity field over the spillways revealed the transition between different flow regimes (nappe, skimming and transition) for changing flow conditions, pressure distributions, and energy dissipation. Tabbara et al. (2005) predicted the water surface profile and the energy dissipation rate for the flow on a stepped spillway by using the CFD method. The turbulent flow model was k-ε with a skimming flow regime. The results agreed with and were close to the experimental work. Aydin & Ozturk (2009) verified and validated that the outcome of a CFD model predicted by FLUENT software is compared with the experimental data for the physical model. The verification was based on the American Society of Mechanical Engineers editorial policy statement (Roache et al. 1986) which provides a framework for CFD uncertainty analysis. Vosoughifar et al. (2013) developed a MATLAB code called V-Flow for CFD with a 2D flow model on a stepped spillway. The assumed flow model was laminar without turbulence. GAMBIT software coupled with V-Flow was used to solve the problem with different configurations of the spillway. FLUENT software and V-Flow were used in validation according to Gonzalez's experimental work. The results showed good agreement between the experimental and CFD data for velocity vectors, flow streamlines, static pressure, dynamic pressure, and total pressure on the spillway. Rajaa & Kamela (2020) used CFD ANSYS Fluent to simulate 2D flow over a spillway. Volume of fluid (VOF) was applied with renormalisation group (RNG) k-ε of RANS equations for turbulent flow. The results showed good agreement and that supports use of the program in future work. Lebdiri et al. (2020) investigated the impacts of step number on the energy dissipation rate, the actual shear stresses on the walls for a flow stepped spillway. Ansys-Fluent was used in the modelling to solve Navier-Stokes equations with both VOF and turbulent k-ε. The results encouraged the researchers to use the CFD approach. A CFD model can show the inside hydraulic variations which can sometimes not be shown physically.

Studying the velocity and pressure distributions on steeped spillways by using new designs for the end sill (curved end sill) is significant. The characteristics of the velocity and pressure distributions have not been comprehensively studied in transition and skimming flow regimes over stepped spillways. In this study, the pressure measurement studied on both the horizontal and vertical faces and the velocity distributions for the steps on a stepped spillway with various step configurations were considered by using physical and numerical models to decrease the negative pressure and the cavitation impacts. The study aims are: (i) simulating flow regimes for different flow conditions with different step shapes (Step models (flat step), Sill models (normal end sill), and Curve models (quarter circle shape end sill which is the new step configuration); (ii) simulating the velocity distribution at the first step, mid-step, and last step for different step shapes; (iii) simulating the pressure distribution for the flow regimes at the first step, mid-step, and the last step for different step shapes; and (iv) simulating the water surface profile with streamlines of the flow along a stepped spillway.

Laboratory experiments

The current work was conducted at Deakin University in Australia. Laboratory experiments were carried out in a 7.0 m long, 0.5 m wide and 0.6 m deep flume (Figure 1). Discharge Q (L/s) was measured by a flowmeter with an accuracy of ±1 L/s. Flow was supplied from a 3.0 m3 storage tank via a 0.2 m diameter pipe to the inlet tank and was straightened before entrance into the flume before exiting via a free overfall. Three different step shapes, each with 2 different geometries, resulting in a total of 6 experimental setups, were tested. All six setups had a total length L = 0.7 m and spillway height Hd = 0.3 m, giving a chute angle θ = 26.6°. The three shapes comprised: (i) rectangular shape or Step Model (Figures 2(a) and 2(b)), where Step_1 has ten flat steps with step height (hs = 0.03 m) and Step_2 has six flat steps with step height (hs = 0.05 m), (ii) rectangular shape with an end sill or Sill Model (Figures 2(c) and 2(d)), where Sill_1 has ten steps (height of the end sill equals step height hs = 0.03 m) and Sill_2 has six steps (height of the end sill equals step height hs = 0.05 m, and (iii) a quarter circle curve step or Curve Model (Figures 2(e) and 2(f)), where Curve_1 has ten steps (height of the quarter circle equals hs = 0.03 m) and Curve_2 has six steps (height of the quarter circle equals hs = 0.05 m). All the test setups were made from plywood and finished with a coat of marine paint to a smooth finish. The Step Model was chosen as the basic shape, for comparison with previous studies conducted with the same chute angle, and ultimately with the Sill and Curve Models. The ranges of experimental parameters tested are provided in Table 1, where hs and ls are the step height and length, respectively, Ns is the steps number, and yc is the critical flow depth. All tests were carried out twice, to test for repeatability. A total of 144 experimental runs were therefore conducted for unit discharges, q ranging from 9.0 L/s.m. to a maximum of 111 L/s.m. Water depths were measured by point gauges, with an accuracy of ±0.2%.

Table 1

Details of the test models and flow conditions

Modelhs (m)ls (m)End sill typeEnd sill height (m)Nsyc(m)yc/hs
Step_1 0.03 0.067 – – 10 0.0198–0.0315 0.66–1.05 
Sill_1 0.03 0.067 normal 0.03 10 0.0198–0.0315 0.66–1.05 
Curve_1 0.03 0.067 quarter circle 0.03 10 0.0198–0.0315 0.66–1.05 
Step_2 0.05 0.120 – – 0.0198–0.0315 0.40–0.63 
Sill_2 0.05 0.120 normal 0.05 0.0198–0.0315 0.40–0.63 
Curve_2 0.05 0.120 quarter circle 0.05 0.0198–0.0315 0.40–0.63 
Modelhs (m)ls (m)End sill typeEnd sill height (m)Nsyc(m)yc/hs
Step_1 0.03 0.067 – – 10 0.0198–0.0315 0.66–1.05 
Sill_1 0.03 0.067 normal 0.03 10 0.0198–0.0315 0.66–1.05 
Curve_1 0.03 0.067 quarter circle 0.03 10 0.0198–0.0315 0.66–1.05 
Step_2 0.05 0.120 – – 0.0198–0.0315 0.40–0.63 
Sill_2 0.05 0.120 normal 0.05 0.0198–0.0315 0.40–0.63 
Curve_2 0.05 0.120 quarter circle 0.05 0.0198–0.0315 0.40–0.63 
Figure 1

The laboratory flume in the civil lab at Deakin University.

Figure 1

The laboratory flume in the civil lab at Deakin University.

Close modal
Figure 2

Stepped spillways tested with different shapes and dimensions. (a) Step_1, (b) Step_2, (c) Sill_1, (d) Sill_2, (e) Curve_1 and (f) Curve_2.

Figure 2

Stepped spillways tested with different shapes and dimensions. (a) Step_1, (b) Step_2, (c) Sill_1, (d) Sill_2, (e) Curve_1 and (f) Curve_2.

Close modal

Velocity was determined from head measurements using a Prandtl-Pitot tube, with 0.01 mm accuracy. The tube has an L-shape with diameter Ø = 3 mm. Tapping of the static pressure occurred 2.5 cm after the tip of the tube at the total head measurement opening. The Prandtl‒Pitot tube was attached at an inclination angle of 30° to reduce the tolerance of the tube reading to ±0.2 mm. In addition, the tube was installed on a digital gauge system, enabling its vertical placement with a 0.01 mm error. The Prandtl‒Pitot tube was supported by a horizontal beam to prevent vibration and movement during high flow. The Prandtl‒Pitot tube data were used to obtain the velocity distributions of non-aerated flows only.

Pressure measurements were recorded by piezoresistive transducers (Measure X, MRB20) with a measurement range of –2 m to +2 m and error of ±0.5%. Four pressure transducers were mounted on the horizontal face to measure the pressure (P1 at 2 cm, and P2 at 1 cm from the corner between the horizontal and vertical faces, Pmid at the mid point of the horizontal face, and Plast at 1 cm before the step edge). After that, for the same conditions, two pressure transducers were mounted on the vertical face (P3 at 1 cm and P4 at 2 cm from the corner between the horizontal and vertical faces). The pressure data was acquired using a data logger (HIOKI, MR 8875) set at a rate of 100 Hz with data capture over 60 secs period. Locations where the transducers were placed are shown in Figure 3. The transducers were used for measuring the pressure on the horizontal face and vertical face for step number one. The transducers were shifted to the mid-step (step number three for models with hs = 0.03 m, and step number 5 for models with hs = 0.05 m) and the last step (step number six for models with hs = 0.03 m, and step number ten for models with hs = 0.05 m) with the same conditions. Figure 3 shows step_2 model as an example. The process was repeated with changes to the discharge value.

Figure 3

The locations of the pressure ports over the stepped spillway (Step_2 model).

Figure 3

The locations of the pressure ports over the stepped spillway (Step_2 model).

Close modal

Numerical model

Governing equations

The ANSYS Fluent v19.0 CFD package was utilized with the VOF method in the current study. The VOF method is a multiphase numerical modelling procedure proposed by Hirt & Nichols (1981) and it is used in free-surface flow problems. In the VOF method, the momentum equation is solved for two phases of non-miscible fluids, and the fraction occupied by each fluid phase in a computational cell is tracked. The model solves the incompressible continuity and momentum equations for turbulent flow (Torrano et al. 2015):
(1)
(2)
where, ρ is the density of the fluid, μ is molecular viscosity, ui is the velocity component, xi is the coordinate component, t is time, and p is pressure. The deviatoric stress component in Equation (2) can be expressed as the following:
(3)
where, μt is the turbulent viscosity, and the stress tensors are when i = j and when i≠j.

Turbulence modelling

In the CFD model, it is important to treat the turbulence of the flow. Solving RANS equations is the method that has been widely utilized in engineering applications. The models of this type depend on averaging flow leading to the RANS equations. The transport equations were added to the Navier–Stokes equations to achieve a closer mathematical flow model. In particular, the transport equations are used to represent the turbulence behavior of the flow to relate that with the turbulent viscosity, which is related to Reynolds stresses (Pope & Pope 2000). Among the available turbulent models, a widely utilized model is the standard k-ε model (Lopez et al. 2015).

Turbulence closure was solved using the k-ε model (Torrano et al. 2015):
(4)
(5)
(6)
where, Cu = 0.09, σk = 1.0, σε = 1.3, C = 1.44 and C2ε = 1.92 (Torrano et al. 2015).

A numerical model (2D flow) was run to steady state for all experimental conditions listed in Table 1. For the inlet, the boundary conditions were separated into two parts to consider the air and water fluid components in all tested models. Velocity equals to average velocity. Velocity was specified for the water flow component, and atmospheric pressure was specified for the air layer at the inlet boundary. Then, the pressure outlet was specified as the outlet boundary condition for the water layer, and atmospheric pressure was specified for the air layer at the inlet, outlet, and top boundary. The flume bed and the surfaces of the steps were defined as the wall boundary conditions.

Geometry and mesh

To represent the domain of the geometry taken by the flow of the air‒water, two meshing approaches are used and both were considered: unstructured mesh and structured mesh. The most appropriate meshing approach depends on the specific case. According to Kim & Boysan (1999), unstructured mesh is suitable for selective refinement to prevent over-refinement in the zones that have expected small gradients. Also, unstructured mesh is best in multi-component geometry. This mesh has fewer closure problems, and its arbitrary topology makes automating the meshing process easier (Biswas & Strawn 1998). In some models, the mesh was slightly refined in the proximity of solid boundaries to accurately resolve the flow features in boundary layers, where larger gradients occur. This may cause the generation of very deformed elements, and it is not an actual problem if the mesh axes remain perpendicular to the solid boundaries (Hirsch 2007). For all these reasons, a static structured rectangular hexahedral mesh is considered the best choice for the cases in the current study. Moreover, inflation layers were used in the mesh because they have the critical components of a good CFD.

In the current study, two geometry types were considered: low complex geometry (the rectangular channel) and very complex geometry (the flow over the stepped spillway). Based on the geometry type, the mesh grid was created of quadrilaterals with two different mesh sizes. The mesh shape is very significant in the analysis, and it affects the total number of elements, run time, and accuracy. For the low complex geometry, a mesh size of 2 mm and element size of 3 mm without inflation layers were used. The very complex geometry has a mesh size of 1 mm, element size of 2 mm, and 10 inflation layers for more accuracy. Figure 4 presents the section of the geometry which has the same dimensions as the physical model and the mesh. The number of elements was 368,490, and the number of nodes was 369,709. The geometry details are listed in Table 1.

Figure 4

The geometry details, (a) geometry, and (b) mesh.

Figure 4

The geometry details, (a) geometry, and (b) mesh.

Close modal

Boundary conditions

The boundary conditions do not change from one case to another except for the inlet velocity. The velocity inlet was calculated using continuity equations for five discharges (Q = 4.35 L/s, 10.10 L/s, 25 L/s, 38.46 L/s, and 55.55 L/s) from experimental work. The velocity was defined as velocity magnitude and the x-direction was used as a velocity direction in the ANSYS-Fluent software. The velocity used was vo = 0.0275, 0.0605, 0.1326, 0.1895, and 0.2566 (m/s) as the data input in ANSYS-Fluent software for all the models, where vo is the inlet velocity. A VOF model was used to simulate the two-phase flow (air and water) with a k-ε model to identify the turbulent flow turbulence in ANSYS-Fluent software. A numerical model was run to steady state for all experimental conditions listed in Table 1. The inlet boundary conditions were separated into two parts to account for air and water fluid components. The average velocity was set for the water flow component. The atmospheric pressure was specified for the air layer. Pressure outlet was taken as the outlet boundary condition for the water layer. In addition, atmospheric pressure was specified for the outlet air layer and top boundary. The flume bed and the surfaces of the steps were defined by the wall boundary (nonslip) condition.

Figure 5 illustrates the initial boundary conditions applied in the CFD model. The upstream boundary was separated into water and air inlets to distinguish between the two-fluid media. The velocity inlet was set as the boundary condition for the part occupied by water flow (BC). The atmospheric pressure was defined for AB and AD boundaries. The nonslip wall boundary condition was used for the steps and flume bed. Pressure outlet was specified in the outlet section (DE). All the dimensions in the model are similar to the physical model (spillway height = 0.3 m, section length (AD line) = 3.0 m, and depth (DE line) = 0.6 m).

Figure 5

The initial boundary conditions.

Figure 5

The initial boundary conditions.

Close modal

CFD model validation

An experimental approach was used define the velocity downstream of the stepped spillway to obtain the CFD results. The stepped spillway researchers were at the start of their assessment. Therefore, it was necessary to check the results. Furthermore, in this study, the CFD model was investigated parallel to other experiments. The root mean square error (RMSE) and the mean absolute percentage error (MAPE) in (Equations (7) and (8)) respectively are the criteria used to validate the CFD models.
(7)
(8)
Also, the correlation coefficient (R) was used as a tool to determine the quality of fit between the experimental and CFD results. Equation (9) was used to compute R.
(9)
where: xi and yj are defined as the CFD and experimental output results, respectively; Cyjdj is the covariance between the CFD (xj) and the physical model results (yj); σxj and σyj are the standard deviations of the CFD and the physical model results, respectively.

Table 2 presents a comparison between the experimental and CFD results. The statistical values such as R, MAPE and RMSE were computed for verification. The benefit of MAPE and RMSE were used as an indicator for the error which should be zero or close to it. In the present study, the model has an index value nearest to zero which is considered more precise. As shown in Table 2, MAPE and RMSE are 4.32 and 0.21 respectively with R = 0.996 for five values of discharges. RMSE was close to zero, signifying the best agreement. In additional, Figure 6 presents critical flow depth (yc) on the x-axis and downstream velocity on the y-axis, where V1 was the flow velocity downstream of the stepped spillway with hs being the step height and vc the flow velocity at the critical depth. When the discharge changed, V1 and the flow depth downstream had significant effects on the hydraulic characteristics because they were the main variables in this process. The results present the minimum and maximum errors for validation data, being 4.32% and 8.21% respectively. Figure 6 shows the downstream velocity V1 that was obtained by the developed 2D flow model. It shows a close correlation within a (±6.27%) tolerance when compare with the experimental work. There was a good agreement between the experimental data and the simulated data.

Table 2

The statistical calculations of experimental and CFD results step_2 model with five different discharges

ModelQ m3/sExperimental ResultsCFD Modelling ResultsThe relative Error %RMSEMAPE %Correlation (R)
V1 m/sV1 m/s
Step_2 0.004 0.147 0.155 5.03 0.021 4.32 0.996 
0.010 0.297 0.312 5.05 
0.025 0.505 0.492 2.57 
0.039 0.615 0.638 3.74 
0.055 0.671 0.706 5.22 
ModelQ m3/sExperimental ResultsCFD Modelling ResultsThe relative Error %RMSEMAPE %Correlation (R)
V1 m/sV1 m/s
Step_2 0.004 0.147 0.155 5.03 0.021 4.32 0.996 
0.010 0.297 0.312 5.05 
0.025 0.505 0.492 2.57 
0.039 0.615 0.638 3.74 
0.055 0.671 0.706 5.22 
Figure 6

Downstream velocity for experiments and CFD models.

Figure 6

Downstream velocity for experiments and CFD models.

Close modal

Experimental data compared with previous study

Figure 7 compares the experimental data of the pressure for the models (Step_2, Sill_2, Curve_2) as a function for x/ls with a previous study where x is the horizontal distance from point O and ls is the step length in Figure 2. The present research obtained low pressure at low discharges in all the models and the relationship between the pressure and the discharge was a positive relationship. In general, pressure increases when discharge increases, consistent with the study by Sánchez-Juny et al. (2007). The differences between the results can be attributed to the influence of the flow conditions and geometry on the flow characteristics.

Figure 7

Pressure comparison with the previous study for Skimming flow.

Figure 7

Pressure comparison with the previous study for Skimming flow.

Close modal

Flow regimes

The CFD model results are presented by model Step_2 as an example. Figure 8 shows the experimental observations and the predicted flow regimes under different flow conditions. A nappe flow regime (NA) occurred in the experiments for q≈20.0 L/s.m and was reproduced by the numerical model (Figure 8(a)). The modelled free surface was noticed to be wavy and disturbed because of the plunging flow. In the experiments, plunging flows were accompanied by substantial amounts of air entrainment, especially for the upper steps, in the flat-step mode. This manifests as regions of white water in the experiments and low (<1) water to air (w/a) ratios. The white water was observed on every step, intensifying as the flow progresses down the steps. The NA CFD results show good similarity with the experimental observations.

Figure 8

Modelled free surface and volume fraction of water at Step_2, (a) nappe flow (b) transition flow and (c) skimming flow.

Figure 8

Modelled free surface and volume fraction of water at Step_2, (a) nappe flow (b) transition flow and (c) skimming flow.

Close modal

The progress to a transition flow regime (TR) was also well predicted by the numerical model (Figure 8(b)). The longitudinal water surface profile was predicted to change from a skimming flow regime (SK) (first three steps) to NA (lower steps) when 20.0<q <39.0 L/s.m. It was consistent with the experimental observations. In the experiments (Figures 8(c)), the results show that the flow started out as SK at the uppermost steps and showed that as the available energy was unable to support SK further downstream, the flow changed to NA, accompanied by increased air entrainment. These observations were also well supported by the numerical model results. SK was predicted by the model for larger flow rates (q≥39.0 L/s.m) where the flow surface appears to be relatively undisturbed, especially for the flat-step model. For the normal sill and quarter- circle sill configurations, some waviness was seen because of the obstruction to the flow by the forward end sills. The observations showed flow regimes delayed in the flow over the new step shape (step with quarter-circle end sill) because of the effects of the quarter circle sill on the recirculated flow. Also, in the pooled stepped models, NA and TR regimes were noted at higher discharges than in the flat-step models.

Moreover, the differences in flow regimes on the pooled step models for the corresponding unit discharge and yc/hs in the flat step came from the impacts of the end sill, especially for the steps with quarter-circle end sill models because of circulation. In all models, the flow of water can be divided into two parts: (1) the first part goes straight to the downstream section, similar to the method in flat-step models; (2) the second part circulates back to the step surface or to the end sills that were oriented perpendicular to the main discharge. Moreover, in the second part of the flow, a small hydraulic jump is produced on the surface of the same step, while the main flow generates a hydraulic jump on the subsequent step. The two parts of the flow then collide in the next step. The hydraulic jump generated by the main part of the discharge disappears, the discharge is disturbed, and the hydraulic jumps become drowned, with more air (bubbles) becoming entrained. Because of these effects, the step height, step length, and end sills affect the flow pattern and its characteristics (hydraulic jump, air entrainment, inception point, and turbulence). The quarter circle end sill shape affected the flow circulation positively, producing a high flow circulation.

Velocity distribution

The flow over the stepped spillway was characterised by high velocity with a big aeration process after the inception point in TR and SK flow regimes. It is important to investigate the velocity distribution for the flow to show its effect on the aeration mechanism. The vertical velocity distribution on the stepped spillway was measured at the midpoint of the horizontal face of the first, middle, and last steps in the experiments (Figure 3). The measurements were taken across the range of the discharge (4.35–55.55 l/s) which included TR (Q = 10.10 l/s) and SK (Q = 55.55 l/s) flow regimes. The CFD model results were validated from comparisons of the modelled velocity distribution over the spillways with laboratory measurements.

Figure 9(a) compares the laboratory measured and modelled velocity profiles for transition and skimming flow conditions. The water depth (y) was normalized with respect to the tread of the step. The water depth y90 is equal to 0.9 of y, whilst v is the velocity at y and v90 is the velocity at 0.9 the depth. In general, laboratory measured profiles agree with power law distributions observed in open channel flow for step number one. In subsequent steps, the measured velocity profiles are relatively large for approximately 90% of the flow depth. The influence of the recirculating pool on the step had a small influence on the velocity profile at the first step. However, the profiles of the velocities at step 5 and the last step of the stepped spillway showed the interaction effects of the recirculating pool. The minimum velocities were measured close to the surface of the steps because of the high impact roughness and shear boundary. The greatest velocities were measured at the top because the roughness effect was very low there.

Figure 9

Velocity distributions, (a) transition flow, step 1, (b) skimming flow, step 1.

Figure 9

Velocity distributions, (a) transition flow, step 1, (b) skimming flow, step 1.

Close modal

The velocity profiles over the stepped spillway pointed out some differences which are related to inflow condition and step configurations. Figure 9(b) presents the CFD and experiment results at the three locations (first step, mid step, and last step) for all the step configurations. The measurements were taken in the SK regime. The CFD profiles of the velocities had slight differences with the flow of ideal fluid theory. However, the velocity distributions presented a similar behaviour to the velocity profile of free overfalls at SK especially at the last step (Henderson 1966). The minimum velocities were greater at SK than TR because the impact of the roughness and shear boundary became less with high discharge. Furthermore, the results showed the impact of gravitational force when the air boundary has its maximum effect on the flow. The velocity profile with the pooled step models (step with quarter circle end sill) was steady being low velocity at the beginning and then rapidly changing when the step pool fill with water. This is a good impact on the velocity because it leads to increase the residence time which was the contribution of the new model (step with quarter circle end sill).

Pressure distribution

The total pressure distribution included static and dynamic pressures. Several pressure measurements were taken over the stepped spillway for all step configurations. Six pressure ports (pressure transducers) were installed on the first, mid and last steps. The measurements were taken for aerated and non-aerated flows on horizontal and vertical step surfaces. For the horizontal step face, four points were used to carry out the pressure measurements on point P1, P2, P3 and P4 as shown in Figure 3. The dimensional less value of yc/hs verses p/γhs were used to see the effects of the flow regimes, step configurations, and chute angle.

The data for the first step is described and discussed below as an example.

Several pressure measurements were carried out on the first, mid, and last steps of all models to see the pressure changes regarding the flow, step shape and the port location. In addition to the four pressure ports installed along the horizontal step face of each step, pressure ports were also installed on the vertical step face for points P5 and P6 at ys = 1 and 2 cm from the corner of the step.

The total pressure measurements on the horizontal step face (P1, P2, P3, and P4) for all the discharge range are shown in Figures 10 and 11. Figure 10 presents the non-dimensional total pressure (P/γyc) at y-axis and non-dimensional value (yc/hs) at x-axis to see the pressure variation at each point. The results show that there was no similarity for the pressure distribution between the step configurations especially for low discharges in nappe flow. The pressure values showed slight convergence with increasing discharge and divergence with decreasing discharge. In Step_1 and Step_2 models, the differences between the measurements were clear because of the impacts of the step height. This happened earlier for the skimming regime compared to the other models. However, a slight difference between the pressure results for models with an end sill (Sill_1, Sill_2, Curve_1, and Curve _2) because of the pooled water effect over the step. The results showed negative pressure at point P1 in both models at low discharge (nappe flow regime) because of the air pocket effect. On the other hand, the total pressure distributions had slight similarity at intermediate discharge and had good similarity at large discharges (transition and skimming flow) because the flow became partial aerated flow or non-aerated flow. Moreover, the results of total pressure distributions in the pooled models (Sill_1, Sill_2, Curve_1, and Curve_2) were all positive values and they had good similarity because the flow was non-aerated flow (there was no chance for air-pocket impact (no air-pocket)).

Figure 10

Modelled total pressure distribution, Nappe flow: (a) Step_2, (d) Sill_2 and (g) Curve_2, where q = 20.0 L/s.m, yc/hs = 0.69. Transition flow: (b) Step_2, (e) Sill_2 and (f) Curve_2, where 20.0<q<39.0 L/s.m, 0.07≤yc/hs≤1.09. Skimming flow: (c) Step_2 (f) Sill_2 and (i) Curve_2, where q = 39.0 L/s.m, yc/hs = 1.08.

Figure 10

Modelled total pressure distribution, Nappe flow: (a) Step_2, (d) Sill_2 and (g) Curve_2, where q = 20.0 L/s.m, yc/hs = 0.69. Transition flow: (b) Step_2, (e) Sill_2 and (f) Curve_2, where 20.0<q<39.0 L/s.m, 0.07≤yc/hs≤1.09. Skimming flow: (c) Step_2 (f) Sill_2 and (i) Curve_2, where q = 39.0 L/s.m, yc/hs = 1.08.

Close modal
Figure 11

Pressure comparison at horizontal step face, (a) point P1, (b) point P2, (c) point P3, (d) point P4.

Figure 11

Pressure comparison at horizontal step face, (a) point P1, (b) point P2, (c) point P3, (d) point P4.

Close modal

For the vertical step face, the comparison between the step configurations for points P5 and P6 are presented in Figure 12(a) and 12(b) the same behaviours were indicated for pressure distribution similarity between the step configurations. In Step_1 and Step_2 models, the changes in the pressure values were higher than the other models (Sill_1, Sill_2, Curve_1, and Curve _2) because of the impacts of the step height, flow depth, and the pooled water over the step.

Figure 12

Pressure comparison on the vertical step face, (a) point P5, (b) point P6.

Figure 12

Pressure comparison on the vertical step face, (a) point P5, (b) point P6.

Close modal

Crest-free surface profiles

This section presents the results for five discharges (q = 20.0 L/s.m with yc/hs = 0.69 to q = 39.0 l/m.s with yc/hs = 1.08) by using the simulation. The free surface shape upstream and above the crest of the stepped spillway was smooth. Figure 13 shows the free-surface profiles obtained by the CFD for Step_1 model as an example. The data highlighted that the water surface profile had a curvature profile, as indicated by (Woodburn 1932), especially at low discharges. At high discharges, the water surface profile has a wavy shape due to the effect of the interactions of the developing boundary layer with the main flow, as explained by Isaacs (1981). The results agreed with the experimental data including the water surface profiles (CFD data) above the crest.

Figure 13

Free surface flow profile over the broad crest for M1-1 model (CFD and Experimental data).

Figure 13

Free surface flow profile over the broad crest for M1-1 model (CFD and Experimental data).

Close modal

Based on the findings of this study, the following conclusions can be drawn regarding stepped spillways with modified step configurations.

Nappe flow, transition flow, and skimming flow regimes were observed for all step configurations and geometries investigated. These observations were consistent with those made in past studies on stepped spillways with various step heights, step lengths, step numbers, and discharges. However, the onset of the TR and SK regimes was dependent on step shape and their modification (the new step shape is referred to as a ‘step with a quarter-circle end sill’), and the flow regime change from NA to TR and to SK was delayed as the discharge increased.

The results showed that the velocity profiles were agreed with a 1/10 power law distribution in all the steps over the stepped spillway. Also, the recirculating pool effect was small on step 1 (first step for 0<x≤Lcrest) but high at the mid-step and last step (middle of the stepped spillway for x = L/2). The maximum velocities were recorded at 90% of flow depth on step 1 and close to the water surface at mid- and last steps.

The total pressure and the flow measurements illustrated the total pressure fluctuations in all the step shapes including the modified models. The total pressure fluctuations were induced by a combination of step-generated turbulence and density fluctuations. Two different regions on the step have been identified: the outer edge of the step (points P1 and P5), governed by impact with the upper jet, and the inner region (points P2, P3, P4 and P6) controlled by recirculating internal eddies. At the flat step shape with nappe flow regime, the pressure distribution was a non-uniform distribution due to the effects of the air pocket. The air pocket was affected by the step shape and the flow characteristics. In the air pocket zone at both the horizontal and vertical faces, with a nappe flow regime some of the pressure measurements were negative at points P2 and P6. However, the pressure distribution on the horizontal step face was a uniform distribution for the TR and SK regimes. The flow was partial aerated flow and fully aerated flow respectively. In the pooled step shape, there was not a negative pressure because the stored water prevents the generation of the air pocket. Also, when the discharge increases, the effect of friction trapped the circulation which leads to reducing the vortices and the circulations. This process affects pressure distribution. In general, the steps with an end sill (normal and curve end sill) have better aeration with no cavitation risk regarding negative pressure.

Comparisons with numerical predictions show that the CFD model can capture the main flow features based on the laboratory results. The water profiles predicted by the CFD model showed good similarity throughout the range of experimental conditions tested. The onset of nappe, transition, and skimming flow regimes predicted by the model for all step shapes corresponded well with laboratory observations. In addition, strong recirculation with a single vortex of regular shape was predicted for steps with quarter-circle end sill models. For steps with normal end sill models, the vortex shape was irregular and occasionally accompanied by a secondary, shear-induced counter-rotating vortex at the upstream corner of the sill. The traditional step model is accompanied by a single shear-induced vortex that is similar to the flow over a backward-facing step, albeit with a step length much smaller than the reattachment length. From the above, the curve models (quarter circle shape end sill which is the new step configuration) showed better performance than other models regarding the chosen flow characteristics.

The authors would like to thank all the staff of the Digital Manufacturing and Civil Laboratories at Deakin University for their assistance and support. The first author would like to acknowledge the full scholarship support provided by the Ministry of Higher Education and Scientific Research of Iraq.

All relevant data are included in the paper or its Supplementary Information.

The authors declare there is no conflict.

Biringen
S.
&
Chow
C. Y.
2011
An introduction to computational fluid mechanics by example
.
John Wiley & Sons
,
New Jersey, USA
.
Biswas
R.
&
Strawn
R. C.
1998
Tetrahedral and hexahedral mesh adaptation for CFD problems
.
Applied Numerical Mathematics
26
(
1
),
135
151
.
Boes
R. M.
&
Hager
W. H.
2003
Hydraulic design of stepped spillways
.
Journal of Hydraulic Engineering
129
(
9
),
671
679
.
Castillo
L. G.
,
Carrillo
J. M.
,
García
J. T.
&
Vigueras-Rodríguez
A.
2014
Numerical simulations and laboratory measurements in hydraulic jumps
. In:
11th International Conference on Hydroinformatics, HIC 2014
,
New York City, USA
,
Chinnarasri
C.
&
Wongwises
S.
2006
Flow patterns and energy dissipation over various stepped chutes
.
Journal of Irrigation and Drainage Engineering
132
(
1
),
70
76
.
Chow
C. Y.
1979
An Introduction to Computational Fluid Mechanics
.
Colorado University
,
Boulder, CO
.
Daneshfaraz
R.
,
Ghaderi
A.
,
Akhtari
A.
&
Di Francesco
S.
2020
On the effect of block roughness in ogee spillways with flip buckets
.
Fluids
5
(
4
),
182
.
http://dx.doi.org/10.3390/fluids5040182
.
Dastgheib
A.
,
Niksokhan
M. H.
&
Nowroozpour
A. R.
2012
Comparing of flow pattern and energy dissipation over different forms of stepped spillway
. In:
World Environmental and Water Resources Congress 2012: Crossing Boundaries
, pp.
1246
1259
.
Frizell
K. W.
,
Renna
F. M.
&
Matos
J.
2013
Cavitation potential of flow on stepped spillways
.
Journal of Hydraulic Engineering
139
(
6
),
630
636
.
Ghaderi
A.
,
Daneshfaraz
R.
,
Torabi
M.
,
Abraham
J.
&
Azamathulla
H. M.
2020
Experimental investigation on effective scouring parameters downstream from stepped spillways
.
Water Supply
20
(
5
),
1988
1998
.
Hamedi
A.
&
Ketabdar
M.
2016
Energy loss estimation and flow simulation in the skimming flow regime of stepped spillways with inclined steps and end sill: a numerical model
.
International Journal of Science and Engineering Applications
5
(
7
),
399
407
.
Henderson
F. M.
1966
Open Channel Flow (No. 532 H4)
.
Hirt
C. W.
&
Nichols
B. D.
1981
Volume of fluid (VOF) method for the dynamics of free boundaries
.
Journal of Computational Physics
39
(
1
),
201
225
.
Hirsch
C.
2007
Numerical computation of internal and external flows: The fundamentals of computational fluid dynamics
.
Elsevier, Burlington
,
USA
.
Isaacs
L. T.
1981
‘Effects of Laminar Boundary Layer on a Model Broad-Crested Weir.’ Research Report No. CE28
.
Department of Civil Engineering, The University of Queensland
,
Brisbane
,
Australia
, p.
20
.
(in chapter 5)
.
Jahad
U.
,
Al-Ameri
R.
,
Chua
L.
&
Das
S.
2018
Investigating the effects of geometry on the flow characteristics and energy dissipation of stepped spillway using two-dimensional flow modelling
. In:
Proceedings of the International Association for Hydro-Environment Engineering and Research. Asia Pacific Division Congress
,
2–5 September 2018
,
Yogyakarta, Indonesia
, pp.
289
296
.
Kaouachi
A.
,
Carvalho
R. F.
,
Lopes
P.
,
Benmamar
S.
&
Gafsi
M.
2021
Numerical investigation of alternating skimming flow over a stepped spillway
.
Water Supply
21
(
7
),
3837
3859
.
Kim
S. E.
&
Boysan
F.
1999
Application of CFD to environmental flows
.
Journal of Wind Engineering and Industrial Aerodynamics
81
(
1–3
),
145
158
.
Lebdiri
F.
,
Seghir
A.
&
Berreksi
A.
2020
Steps number effect on hydraulic parameters of flows in stepped spillways
.
Larhyss Journal P-ISSN 1112-3680/E-ISSN 2521-9782
.
3
(
42
),
41
51
.
Lebdiri
F.
,
Seghir
A.
&
Berreksi
A.
2022
Multi-objective optimization of stepped spillway and stilling basin dimensions
.
Water Supply
22
(
1
),
766
778
.
Lopez-Egea
M.
,
Nistor
I.
,
Towsend
R.
,
Paudel
B.
&
Sullivan
P.
2015
An experimental and numerical study of submerged hydraulic jumps forming at low-head dams
. In:
E-proceedings of the 36th IAHR World Congress
,
June
, Vol.
28
, pp.
24
35
.
Pope
S. B.
&
Pope
S. B.
2000
Turbulent flows
.
Cambridge University Press
,
Cambridge, UK
.
Rajaa
A. I.
&
Kamela
A. H.
2020
Performance study of fluent-2D and flow-3D platforms in the CFD modeling of a flow pattern over ogee spillway
.
Anbar Journal for Engineering Sciences
8
(
4
),
317
328
.
Roache
P. J.
,
Ghia
K. N.
&
White
F. M.
1986
Editorial Policy Statement on the Control of Numerical Accuracy
.
J. Fluids Eng.
108
(
1
),
2
https://doi.org/10.1115/1.3242537
.
Sánchez-Juny
M. J.
,
Blade
E.
&
Dolz
J.
2007
Pressures on a stepped spillway
.
Journal of Hydraulic Research
45
(
4
),
505
511
.
Shahheydari
H.
,
Nodoshan
E. J.
,
Barati
R.
&
Moghadam
M. A.
2015
Discharge coefficient and energy dissipation over stepped spillway under skimming flow regime
.
KSCE Journal of Civil Engineering
19
(
4
),
1174
1182
.
Tabbara
M.
,
Chatila
J.
&
Awwad
R.
2005
Computational simulation of flow over stepped spillways
.
Computers & Structures
83
(
27
),
2215
2224
.
Takahashi
M.
&
Ohtsu
I.
2012
Aerated flow characteristics of skimming flow over stepped chutes
.
Journal of Hydraulic Research
50
(
4
),
427
434
.
Torrano
I.
,
Tutar
M.
,
Martinez-Agirre
M.
,
Rouquier
A.
,
Mordant
N.
&
Bourgoin
M.
2015
Comparison of experimental and RANS-based numerical studies of the decay of grid-generated turbulence
.
Journal of Fluids Engineering
137
(
6
),
061203
.
Vosoughifar
H. R.
,
Dolatshah
A.
,
Shokouhi
S. K. S.
&
Nezhad
S. R. H.
2013
Evaluation of fluid flow over stepped spillways using the finite volume method as a novel approach
.
Strojniški vestnik - Journal of Mechanical Engineering
59
(
5
),
301
310
.
Woodburn
J. G.
1932
Tests of broad-crested weirs
.
Transactions of the American Society of Civil Engineers
96
(
1
),
387
416
.
This is an Open Access article distributed under the terms of the Creative Commons Attribution Licence (CC BY 4.0), which permits copying, adaptation and redistribution, provided the original work is properly cited (http://creativecommons.org/licenses/by/4.0/).