Fig 4 - uploaded by Mohamed Yossef
Content may be subject to copyright.
Velocity distribution in the transverse direction of ( u ) along sections A to D for cases A1: emerged groynes, A2 and A3: submerged groynes (measured using EMF — time-averaged) 

Velocity distribution in the transverse direction of ( u ) along sections A to D for cases A1: emerged groynes, A2 and A3: submerged groynes (measured using EMF — time-averaged) 

Source publication
Article
Full-text available
Experiments have been carried out in a fixed-bed flume for a schematised straight river reach with groynes on one side to study the dynamics of the flow near groynes. The flume had a geometrical scale of 1∶40, based on typical dimensions of the Dutch River Waal. Both emergent and submerged groynes were studied. The measurements demonstrate the diff...

Contexts in source publication

Context 1
... considered representative of groyne fields in a straight reach with similar aspect ratio. The bed of the model is fixed and flat in the main channel; and slopes toward the bank in the area between the groynes [Fig. 2(b)]. The water level is controlled downstream by a manually controlled tailgate. In all tested cases, the Froude number ( F ) was around 0.2 to obtain subcritical flow as observed in the prototype, and the Reynolds number ( R ) was high enough to ensure turbulent flow in both the main channel region ( R 6 × 10 ) and the groyne fields region ( R ≅ 10 4 ). This paper considered only one of the tested groyne types, which represent a typical groyne as found in the Dutch Rhine branches. This groyne type is a straight, impermeable structure perpendicular to the riverbank with a crest length of 1.5 m and with slopes of 1 ∶ 3 on all sides, and the beach slope of the groyne fields is 1 ∶ 30 [see Fig. 2(b)]. The aspect ratio of the groyne fields (spacing between groynes/length of groyne) is 3.3. Uijttewaal (2005) rovides an overview of the effect of other tested geometries on the flow in groyne fields. Velocity measurements carried out during each test comprised two measurement techniques: electromagnetic flowmeters (EMF), and particle-tracking velocimetry (PTV). Four EMFs were used, meas- uring simultaneously the u -and v -components of the velocity in the horizontal plane (EMF sensor: electromagnetic, biaxial, 4-quadrant ellipsoid 11 × 33 mm; For this type of probe, the influence area has a cylindrical shape with approximately a thickness of 5 mm and 50 mm diameter, situated just below the electrodes). The measurements covered four sections, A to D as indicated in Fig. 2(a). The sections extended over 2.6 m in the transverse direction, and they were spaced at an equal distance of 0.75 m. Each section consisted of 12 points. At each point, velocities at two vertical elevations, located at 0.2 h and 0.8 h from the surface were recorded to dis- tinguish the difference between the near bed and the near surface velocities (little difference was observed in the velocity fluctuation pattern). PTV was used to measure the instantaneous surface velocity without causing disturbance and characterized the global structures of a complicated unsteady flow field. PTV was applied in four areas of 1 : 5 × 1 : 5 m 2 ; these areas are indicated as P1 to P4 in Fig. 2(a). Each time series lasted at least 5 min with a recording frequency of 30 Hz. Each frame contains about 2,000 particles of 2 mm diameter, floating on the free surface. The velocities were determined using a PTV-algorithm according to Bastiaans et al. (2002). The unstructured data were then interpolated onto a uniform rectangular grid with a characteristic spatial resolution of 5 cm. This resolution is, however, not sufficient to resolve the full spectrum of turbulent fluctuations, and so conclusions to be drawn from these measurements only apply to the large-scale turbulence structures. The absolute accuracy of the used EMF is less than 1 : 0 cm = s. When evaluating time-averaged velocities the uncertainty reduces because of averaging large number of samples. The accuracy of the PTV system is much less than the EMF. Accordingly, only quali- tative information is deduced from the PTV results. The test series discussed in this paper, Series-A, investigates the dynamics of the flow in the vicinity of a series of groynes at two different flow stages; namely, the emergent-and submerged groynes stages. The effect of flow stage is studied by changing the flow depth and keeping the main channel mean velocity constant at about 0 3 m s. Three water levels were tested; see hydraulic conditions in Table 1. A water level of 0.25 m represented the emergent condition (case A1), while two high water levels of 0.30 m and 0.35 m (cases A2 and A3), represented two submerged cases. The measurement locations covered the entire width of the mixing layer from section A to section D. The sampling time was guaranteed to be long enough to ensure a full coverage of the largest turbulence structure; the total recording period using EMF for a single point was 600 s, with a sampling rate of 10 Hz. The inflow conditions were adapted to the cross-section velocity profile far downstream such that beyond the first two groyne fields the flow can be considered as fully developed; this has been checked in the model using dye visualization. EMF velocity samples were measured throughout sections A to D. The time series (Fig. 3) taken at the center of the mixing layer at section A (at 0.2 h) gives an impression of the amplitude and the timescale of the velocity fluctuations. In all three cases, the velocity fluctuations in the transverse y -direction are slightly more pronounced than those in the streamwise direction. Comparison of the time signals of the different cases reveals that the velocity fluctuations in the submerged condition have a longer timescale than those in the emerged condition. Fig. 4 presents the time-averaged velocity distribution in the transverse direction of the streamwise velocity ( u ), measured at sections A through D for the three test cases A1, A2, and A3. A comparison between the two submerged cases shows little differences between the velocity profiles, whereas, larger differences are observed if they are compared with the emerged case. Moreover, the transverse velocity distribution in the emerged case [Fig. 4(a)] converges from section A to B, to reach a rather stable time- averaged distribution at sections C and D. In the submerged cases [Figs. 4(b) and 4(c)] the time-averaged velocity distribution converges within a rather short distance, as it is nearly the same for sections B, C, and D. Fig. 6 gives the transverse shear stresses ( τ xy ), and is a representation of the time-varying component of the velocity. The variation between the profiles of τ xy is a manifestation of changes in turbulence intensity due to the variation in the mixing layer. Fig. 5 presents the total turbulence intensity extracted from the PTV measurements for the three test cases A1, A2, and A3. The figure indicates the extent of the mixing layer in the different cases. In the emerged stage (case A1), the mixing layer starts with flow separation at the tip of the groyne and grows in width in the downstream direction. The total turbulence intensity and the transverse shear stress ( τ xy ), have their maxima around 1.5 m downstream of the tip of the groyne [see Figs. 5(a) and 6(a)]. Fig. 6 is based on measurements at 0.2 h from the surface; similar behavior, yet slightly less pronounced, is observed from the analysis of the measurements near the bottom. In the submerged stage, the mixing layer develops because of the steep velocity gradient between the fast stream in the main channel and the slow stream in the groyne fields region. It has a constant width [see Figs. 5(b), 5(c), 6(b), and 6(c)]. The width of the mixing layer is maintained because of the existence of a series of groynes that maintain the velocity difference between the fast stream (main channel), and the slow stream (groyne fields region). The nearly constant velocity difference maintains a constant velocity gradient across the mixing layer, thus keeping it constant in width. We can think of it (with reservations), to be analogous to a compound channel case with the mixing layer forming on the interface between the main channel and the floodplain (for example, see Van Prooijen et al. 2005). A comparison between the two submerged cases, A2 and A3 shows that with increasing submergence, the magnitude of the total turbulence intensity and the transverse shear stress decrease, while the width of the mixing layer remains the same [see Figs. 5(b), 5(c), 6(b), and 6(c)]. When the groynes are not submerged, the flow inside the groyne fields, in the horizontal plane, shows the circulation pattern, as reported previously by Uijttewaal et al. (2001) for a similar aspect ratio (see as well Uijttewaal 2005; Yossef 2005). Fig. 7 gives an illustration of the flow pattern, which is characterized by: • A primary eddy that forms in the downstream part of the groyne field and covers nearly two thirds of its spacing; the magnitude of the circulation velocity is about 30% to 40% of the main channel mean velocity; • A secondary eddy driven by the primary one with an opposite sense of rotation and a much smaller flow velocity; and • A dynamic eddy that sheds regularly from the tip of the upstream groyne. This eddy migrates in downstream direction and merges with the primary one, which in return changes in size due to the interaction with the migrating eddy. The whole circulation pattern is driven by the main stream via exchange of momentum through the interfacial mixing layer. Fig. 8 presents the autocorrelation functions of the transverse velocity for test case A1 (emergent groynes) measured using EMF at different locations. The large turbulence structures can be recognized through a correlation that extends over large time lags, and the characteristic time of the modulation is related to the timescale and consequently the length-scale of the turbulent structure. Inside the groyne fields [Fig. 8(a)], the presence of positive and negative correlations at large time lags corresponding to large structures is clear. Near the detachment point of the dynamic eddy [Fig. 8(b); section A], strong modulation with a shorter time span is visible, corresponding to the dynamic eddy. At section D, the autocorrelation shows the same characteristics as that inside the groyne field, implying that the same structure exists in these two points; evidence that it is the primary eddy. In the main stream [see Fig. 8(d)], the autocorrelations at large time lags are negligible. This suggests that only bottom-induced turbulence is present in the main stream far from the groynes. Unlike a free mixing layer, which is characterized by a single dominant peak in the energy ...
Context 2
... groynes. Unlike a free mixing layer, which is characterized by a single dominant peak in the energy density spectrum (at a given location, see Van Prooijen 2004), the mixing layer from a series of groynes contains more peaks caused by the interaction between the shedding eddies from the groyne tips and those migrating from upstream groynes, see Figs. 9(a) – 9(c). Inside the groyne fields [Fig. 9(a)], two peaks can be observed in the energy density spectrum at section A. The peak at the higher frequency has faded away in section D, where the lower frequency dominates. This can be related to the difference in length-scale between the eddy at section A and that at section D. The existence of the two peaks is caused by the interaction between the primary and the secondary eddies. Near the center of the mixing layer (point #2), Fig. 9(b) shows that near the groyne tip at section A, in addition to the two peaks that are present at point #1, another sharp peak at a higher- frequency is found. The presence of this peak is caused by the dynamic eddy that sheds from the tip of the groyne at a high- frequency; it appears only near the tip of the groyne and not at section D. The same pattern can also be inferred from the autocorrelations presented in Fig. 8(b). The effect of the groynes further upstream is visible in the presence of the low-frequency peaks at section A. Point #3 at section A is located just outside the mixing layer near the groyne, but still within the mixing layer of the upstream groyne. The energy density spectra at that point [Fig. 9(c)] show the existence of only a low-frequency peak with relatively low energy content at section A, caused by the mixing layer of the upstream groyne. At section D, two peaks can be observed; the higher- frequency peak is due to the mixing layer that forms because of the nearby groyne and the low-frequency peak is attributed to the interaction of the migrating eddy from the upstream groyne. In the main stream (point #4), the large structures nearly disappear and only small-scale turbulence can be observed [see Figs. 8(d) and 9(d)]. When the groynes are submerged, the flow in the groyne fields region does not show the horizontal circulation pattern that is observed in the emerged case. The visual observations of no circulation are supported by the cross-section velocity profiles where reverse flow was not observed in the submerged groyne cases; Figs. 4(b) and 4(c) show no -ve velocity in the groyne field region. The groyne fields can be characterized as a low-velocity region. The momentum transfer by the water flowing over the groynes is sufficient to balance the momentum transfer through the mixing layer, that otherwise would have caused a recirculation inside the groyne field. As the groynes are submerged, the flow over the groynes hinders the horizontal recirculation, causing it to disappear after relatively low submergence. Generally speaking, the flow pattern in the submerged stage shows an alternating accelerating and decelerating pattern between flow over and around the groynes. From Fig. 10, we see that the turbulence characteristics in the submerged condition are different from that of emergent conditions. There is a slow periodical behavior with a timescale of around 25 s, which is present in all locations, illustrated in Figs. 10(a) – 10(d). This suggests a phenomenon that occupies the whole length of the groyne field. Yet this phenomenon does not seem to have a free-surface signature, in the sense that no significant free-surface waves were observed. The energy density spectra presented in Fig. 11 shows a distinct single peak at a similar frequency for both sections A and D, meaning the same fluctuation is present throughout the length groyne field. From Fig. 11(d), one can observe the existence of large structures at point #4 far into the main stream. Thus, the submerged case the mixing region extends further into the main channel than in the emerged case. In this section we parameterize the complex flow field near groynes in terms of depth-averaged components. The parameterization provides some building blocks that may serve further analysis of, for example, the sediment exchange process between the main channel and the groyne fields of a river (see Yossef 2005; Yossef and De Vriend 2010). In the previous section, large-scale velocity fluctuations in both horizontal directions were observed and the velocity field indicated a periodical behavior. The case of submerged groynes has an anal- ogy with the free mixing layer between a fast stream and a slow stream (see, for example, Uijttewaal and Booij 2000), where the fast stream is the main channel flow and the slow stream is the flow far inside the groyne field, away from the mixing layer. The case of emerged groynes includes more dynamics; viz, a dynamic eddy detaching from the tip of the groyne and the large circulation cell inside the groyne field. However, the time-averaged velocity profiles given in Fig. 4 and the transverse shear stresses given in Fig. 6 both indicate a level of similarity with free mixing layer. Accordingly, we may apply the same treatment as in the case of free mixing layer. Representing the fluctuating velocity component of the coherent structures in the mixing layer with a periodical function with prescribed amplitude, period and phase difference seems a natural choice (cf., Lyn and Rodi 1994). Assuming the full velocity signal takes the form u 1⁄4 u þ u 0 þ u 00 ð 1 Þ where u , v = time-averaged components; u , v = large-scale turbulence components; and u 00 , v 00 = small-scale (bottom) turbulence components. The large-scale turbulent part is represented as u 0 1⁄4 û · sin ð ω · t Þ ð 2 ...

Similar publications

Article
Full-text available
To assess the spatial and temporal fluctuations of zooplankton biomass, as dry organic weight per cubic meter (B), in the northernmost Gulf of California, twelve oceanographic cruises were carried out during October 1972-October 1973. There was no clear seasonal cycle. Highest B values (as high as 154 mg m-3) were always found in the channels aroun...
Article
We implemented a species distribution modelling approach to establish the potential distribution of Podocnemis lewyana, to explore the climatic factors that may influence the species’ distribution and to evaluate possible changes in distribution under future climate scenarios. The distribution models predicted a continuous distribution from south t...
Article
Full-text available
There are face slab cracks on the Gongboxia concrete face rockfill dam (CFRD), which is located on the Yellow River in China. They generally occur as vertical cracks at the top of the panel near the water level. The region where Gongboxia CFRD is located is subjected to low temperatures during the winter months and high temperatures during the summ...
Chapter
Full-text available
A detailed survey has been carried out in order to assess the pollution status of the Haora River at different critical points and in different time periods (2004–2013). All the 15 parameters of the ten selected water samples are categorised into three heads for analysing the pollution status in details. Among physical components of the water quali...

Citations

... The numerical representation of the hydraulic resistance of groynes at varying water levels remains subject to debate. In 1D (one-dimensional) and 2DH (two-dimensional horizontal) numerical models, groynes are commonly represented as subgrid features with a local energy loss according to a submergedweir discharge formula (Ali, 2013;Bloemberg, 2001;Jongeling et al., 2010;Mosselman & Struiksma, 1992;van Broekhoven, 2007;Yossef, 2005Yossef, , 2017Yossef & de Vriend, 2011;Zagonjolli et al., 2017). The problem is that flow over groynes differs from flow over weirs. ...
... These shear zones form mixing layers in the transition area between the main channel and the groyne field. Yossef and de Vriend (2011) show that the mixing layer extends in the flow direction with almost the same intensity when there is a series of groynes. This transfers momentum from the main channel to the groynes region. ...
Article
Full-text available
The hydraulic resistance of groynes is an important factor in the determination of design flood water levels on rivers and the assessment of how much these levels are lowered by modifying the groynes. In standard one‐ or two‐dimensional numerical hydrodynamic models for flood risk management, groynes are commonly represented as subgrid features with a local energy loss according to a weir formula. We tested this representation by using a two‐dimensional horizontal mesh at various groyne submergence degrees by comparing the results with those of flume experiments. We also compared the results with simulations using different 2D and 3D approaches on finer grids that incorporate groynes in the bed topography. In one of the two tested 3D models, complete Reynolds‐averaged Navier‐Stokes equations were solved. The second tested 3D model was constructed simpler by assuming hydrostatic pressure distribution in the vertical direction. We employed Delft3D software in construction and execution of all models. One of the 3D models did predict the hydraulic resistance at low submergence better than the standard model, but it slightly underestimated the resistance at higher submergences. Despite differences in flow characteristics, weirs and groynes were found to produce similar flow resistances for the same height and boundary conditions. Simulations of groyne modifications showed that hydraulic resistance decreased nonlinearly with groyne lowering and streamlining.
... Qasim et al. [5] did some experiments to study the influence of the bed flume contraction on the combined structure discharge coefficient, also this work also dealt with Reynolds number, Froude number, actual discharge, flow velocity and downstream water depth. Yossef and de Vriend [6] performed experimental investigation in order to study the flow pattern for both submerged and emergent groynes. The groynes are installed in one side of straight fixed bed flume. ...
... Mat. 3, Fig. S1) and the accelerated flow causes increased mixing between bulk flow and recirculation zone (Fig. 6). The mixing layer width increases with the distance from the upstream groyne which is consistent with the study of Yossef and de Vriend (2011). All those circumstances resulted in a situation when increasing discharge increased the number of particles entering the groyne field past the extension but at the same time, those particles were floating away from this area faster (Fig. 5, scenario c). ...
Article
Macroplastic, a precursor of microplastic pollution, has become a new scope of research interest. However, the physical processes of macroplastic transport and deposition in rivers are poorly understood, which makes the decisions of where to locate macroplastic trapping infrastructure difficult. In this research, we conducted a series of experiments in a laboratory channel, exploring the impact of groynes and flexible artificial vegetation on the floating macroplastic litter. The goal was to investigate the litter paths with different obstruction arrangements, which was done by implementing a particle tracking technique on video recordings from each experimental run. We found that increasing discharge correlated with the number of plastic litter floating into the recirculation zone within the groyne fields, especially if the upstream groyne had an extended length. This produced a strong mixing interface between the main flow and the groyne field, while a vegetation patch added in the same groyne field changed the paths of plastic litter by deflecting the flow. We noticed that during a moderate discharge rate, the litter pieces flowing into the groyne field with the vegetation circulated there for the longest period, and some of them got entangled between floating stems when discharge was at its lowest. This phenomenon points to the conclusion that low flow velocity paired with the presence of vegetation can be a primer for plastic deposition and consequently, its degradation. The insights from the experiment allowed us to recommend a place downstream of an extended groyne as the desirable (efficient) area for installing a plastic trapping infrastructure or conducting plastic cleaning actions.
... Spurs or groynes are considered prime hydraulic transverse structures to enhance navigational features and mitigate riverbank erosion. The structures constructed with various inclinations concerning the bankline to deflect water flow from the erosion-prone zone are spurs or groynes (Yossef & de Vriend 2011). Spurs are built using materials such as concrete, rock, timber, earth, or geosynthetics (Barkdoll et al. 2007;Dehghani et al. 2013). ...
Article
Full-text available
In this paper, the significance of the shape of spurs on flow and bed morphology is understood, accompanied by an appeal to investigate alternative shapes like L-head, T-head spurs, and hockey-shaped spurs, which could offer considerable benefits for river management. The spurs with T-head shape are found to be more efficient in reducing the bed scouring of the channel relatively, although it depends on the orientation of the spur, location of installation, and shape of the channel. Furthermore, development of horseshoe vortex is found insignificant near the base of the T-head spur dike. T-head spurs could more effectively redirect flow and prevent erosion while, L-head spurs may enhance riverbank stability. This highlights the need for more investigation into how different spur shapes collectively affect river morphology, flow velocity, and sediment transport patterns. Temporal variations in bed morphology, especially the scour depth and sediment transport dynamics near spur dikes should also be investigated to understand their changing impact better. Finally, this study provides a complete summary of existing knowledge gaps and future research initiatives relevant to various shapes of spur dikes and their role in riverbank erosion management.
... Studies by [21,22] and others have investigated that groin fields having slow flow velocities and shallow water encourage vegetation growth, silt deposition, and the best conditions for fish to survive and feed. [23] studied the effects of flow velocity and water depth on fish habitats both with and without groins. ...
... Salmon habitats downstream of the restored reach were improved due to the bed grain size fining. [21] used a scaled physical model of 1:40 to study the flow near the Dutch River Waal, the results showed variations in turbulence between the submerged groins and the emergent ones. The flow in the groin fields region in the submerged stage exhibits an alternating accelerating and decelerating pattern between flow over and around the groins but does not exhibit the circulation pattern as seen in the emerged state. ...
... Dikes and groynes have been widely used around the world to regulate and modify hydrodynamics, sediment dynamics, and morphodynamics in rivers, estuaries, and coasts (Yossef and de Vriend 2011). They are hard coastal protection structures that aim to promote navigation and to protect the shoreline from coastal erosion. ...
... In coastal zones, they are usually constructed perpendicular to shorelines, functioning as a protection for coasts and beaches from being eroded by alongshore wave-driven currents. Up to now, the effects of groynes were primarily investigated in the context of (1) two-dimensional (2D) hydrodynamic exchange mechanisms, e.g., horizontal eddy (Sukhodolov et al. 2002), turbulence and shear (Uijttewaal 2005), water exchange (Uijttewaal et al. 2001), (2) sediment related issues, e.g., sediment exchange (McCoy et al. 2007;Yossef and de Vriend 2011), sedimentation and morphodynamics (Glas et al. 2018;Sukhodolov et al. 2002), and (3) threedimensional (3D) flow patterns within the groyne fields (Biron et al. 2005; Ouillon and Dartus 1997) influencing the residence time within the groyne field (Tritthart et al. 2009). But although groynes are more and more constructed in estuaries, one aspect of groynes that has not been systematically investigated is the role of salinity-driven exchange flows. ...
Article
Full-text available
Existing knowledge about groyne-induced effects is primarily based on riverine or coastal environments where salinity gradients are absent or limited. However, in estuaries, salinity gradients drive physical processes such as longitudinal and lateral residual flows. The effect of groynes is much more complex because they can modulate channel hydrodynamics and directly affect lateral salinity gradients. In this study, an idealized model is applied to investigate the effects of groyne layouts in estuarine environments, including effects on (1) channel hydrodynamics, (2) lateral water exchange, (3) Coriolis effects, and (4) saltwater intrusion. Model results show that the aspect ratio (the width of groyne fields to the length of groynes) of groyne fields plays an important role. Groynes also induce asymmetry of lateral flows, for example, increasing near-bottom shoal-to-channel flows during low water slack. The aspect ratio has opposite effects on horizontal and vertical components of water exchange. A large aspect ratio strengthens horizontal exchange and weakens density-driven currents. For a large-scale groyne field (several kilometers), Coriolis effects introduce a substantial difference in exchange mechanisms along the north and south banks. A medium range of aspect ratio (2.0-3.0) leads to the strongest saltwater intrusion during both neap and spring tides. Practical Applications: Dikes and groynes are common engineering structures in waterways designed to increase along-channel flow velocities and prevent sediment deposition, thereby maintaining navigability. However, especially in estuarine environments, cross-channel flow velocities resulting from these structures may generate sediment transport toward the channel, in contrast to their original intention. Another important impact of these structures is on saltwater intrusion, which is crucial to freshwater resource management. Therefore, this research investigates the groyne-induced effects on cross-channel flows and the saltwater intrusion problem by comparing different layouts of groynes and their consequences. Two important findings may benefit practical applications. First, the interaction between saltwater and freshwater in estuaries enhances cross-channel flows during certain periods, thereby influencing sediment transport. Second, the maximal saltwater intrusion occurs for intermediate width-to-length ratios of the groynes, with lower saltwater intrusion for either very small or very wide groyne fields.
... The presence of lateral cavities in open channels leads to the formation of strong vortices in the separated shear zones, resulting in a range of water quality and ecological effects (Valentine andWood 1979, Hill 2014). The hydrodynamic instabilities may be highly coherent when some coupling occurs between the separation layer and the natural wave modes in the channel and lateral cavity (Tuna et al. 2013, Perrot-Minot et al. 2020.Despite a wealth of research on subcritical flows past lateral cavities (Ohmoto et al. 2005, Yossef andDe Vriend 2011), the study of flows past successive lateral cavities, including the interactions between cavities, has attracted a lesser interest, except a few studies (Meile et al. 2011, Tuna and Rockwell 2014, Leng and Chanson 2020b. Herein, new laboratory experiments were performed at the University of Queensland, in a relatively large and near-full-scale fish-friendly culvert model. ...
Article
Full-text available
A near-full-scale physical modelling was performed on the oscillation and instability of open-channel flow in a fish-friendly culvert equipped with full-height sidewall baffles. High-resolution measurements of the instantaneous flow velocity were conducted using a Vectrino+ acoustic Doppler velocimeter (ADV). The physical results were marked by the existence of some low-frequency seiche phenomenon. A well-established triple decomposition technique was applied to the time series of free-surface and velocity time-series. The low-pass components confirmed a unique flow structure, consisting of a high-velocity zone in the main channel and a low-velocity flow reversal within the lateral cavities. The band-pass components corresponded to the low-frequency flow oscillations, highlighting the complicated transverse interactions between the lateral cavity and the main channel. The high-pass velocity components were related to the 'true' turbulence characteristics. The current study provides some further insights into the sustainable design of culverts to assist with upstream fish migration in man-made and natural fast waterways.
... These studies mostly adopt impermeable groynes or similar structures with different layouts. Numerous studies have been conducted for analysing the scouring and related turbulent characteristics around the groynes in straight channels using impermeable groynes (Pandey et al. 2022(Pandey et al. , 2021aAtarodi et al. 2021;Sharma and Ahmad 2020;Yarahmadi et al. 2020;Kumar and Ojha 2019;Nayyer et al. 2019;Yarahmadi and Bejestan 2016;Zhang et al. 2009Zhang et al. , 2012Yossef and Vriend 2011;Yazdi et al. 2010) and permeable groynes (Iqbal et al. 2021;Mostafa et al. 2019;Ahmed et al. 2015;Kang et al. 2011;Yeo et al. 2005). Studies were also carried out in curved channels on scour and turbulence around impermeable groynes (Matsuura and Townsend 2004;Jamieson et al. 2013;Vaghefi et al. 2016;Biswas and Barbhuiya 2019) and permeable groynes (Kalamizadeh et al. 2021;Hajibehzad et al. 2020;Ferro et al. 2019). ...
... For experimental studies, groynes have been constructed using different materials like concrete (Kuhnle et al. 1999;Uijttewaal 2005;Yossef and Vriend 2011), metal plates (Kuhnle and Alonso, 2013), plexi glass (Yeo et al. 2005;Biswas and Barbhuiya 2019;Yarahmadi et al. 2020;Tripathi and Pandey et al. 2021b), plywood (Bhuiyan et al. 2010), gravels (Matsuura and Townsend 2004) and gabions (Klingeman et al. 1984) with different head shapes. The use of permeable dykes reduces the tip velocity resulting in less scouring (Yeo et al. 2005;Kang et al. 2011;Ferro et al. 2019;Mostafa et al. 2019;Kalamizadeh et al. 2021) and less turbulent kinetic energy (Iqbal et al. 2021) at the groyne head. ...
Article
The study investigates the role of cocolog, an eco-friendly material, as groyne, in protecting the river banks in meandering channels. Laboratory experiments were conducted to compare the flow pattern and sediment dynamics in a mobile-bed meandering channel for cocolog (permeable) and impermeable groynes. It is observed that as the permeability increases, the vertical and transverse velocities are seen to be reduced in the groyne fields. The results also indicate that places of maximum scour depth display amplified velocity, turbulent intensity and turbulent kinetic energy. However, considerable reduction in these characteristics is being observed in the case of permeable groynes that have densities varying between 140 and 160 kg/m³. The eroded volume is seen to be reduced by 60% for permeable groynes of medium density compared to the tests done without groynes. Also impermeable groynes exhibit a wider distribution of the maximum values for scour depth, turbulence intensity and turbulent kinetic energy than that of permeable groynes. It is found that the cocolog groynes with suitable density (or permeability) can perform best in meandering channels by dampening the velocity and turbulence of the groyne fields.
... Generally, the flow near around the spur dike can be further divided into three zones: main flow zone, wakeup zone, and mixing zone in between the two zones. The large-scale u and v velocity fluctuations are in phase-in the center of the mixing layer and out of phase for the points (Chabert and Engelding 1956) on the boundaries of the mixing layer (Yossef 2011). The main flow area extends from the top edge of the structure to the opposite side of the channel; the region extending to the bank from which the dike is protruded to the head of the dike is termed as the wake-up zone, and the mixing zone lies in the middle of two. ...
Article
A spur dike is a hydraulic structure, protruding in a river or channel used for several purposes like protection of river-bank erosion and deepening of the main channel. The present paper discusses pre-existing research work on flow pattern and prediction of temporal and maximum scours depth around the spur dikes placed in different locations at 90∘ and 180° curved channels. The equations having approximately 2.367, 4.47, 0.17, and 0.271 (average) times with their corresponding experimental data. The parameters, influencing the scour process and flow pattern, have been identified as the ratio of flow intensity to critical velocity (V/Vc ≥ 1) is below 1 and special kind of bedding material is approximately 10 % greater than under live-bed condition and many more. The numerical value of the Froude number and the geometry of the bed surface material are also discussed in this paper. Based on these parameters, the empirical formulations and experimental studies on local scours around the straight, L-shaped, T-shaped spurs, placed at 30°, 45°, 60°, 120°, and 180° azimuthal angles have been discussed. Various numerical schemes proposed in almost seventy-five literatures have been summarized. A critical review of numerical and experimental results found in different works related to temporal and maximum scour depth, flow characteristics, and bed topography around the dike shows that the data and accompanying results are insufficient for the design of spurs used as river structures in curved channels. There are needs to carry out extensive experiments, under various flow conditions, to examine the flow behavior and scouring processes around the spurs. Due to complex flow pattern and scouring processes, taking place around the spur, it becomes difficult to understand the real physics behind these phenomenon and therefore, data-driven models are suggested to arrive at more reasonable relationships required to be used for design purposes.
... It can not only clear bed sediments by constraining flow and serve for river training, but also provide a habitat for fishes by changing the surrounding hydrodynamic conditions. Studies by Arlinghaus et al. (2002), Armstrong et al. (2003), Yossef and Vriend (2010b), and other researchers have shown that groin fields, with shallow water and low flow velocities, allow sediment deposition and vegetation growth, and provide favorable habitats for fishes to spawn and forage. Chang et al. (2019) investigated the flow behavior around a groin system along the Yizheng section of the Yangtze River and found that this groin system provided abundant food and a stable migration route for fishes. ...
... Besides, there is also experimental research on this topic. For instance, Yossef and Vriend (2010b) investigated the flow near the Dutch River Waal by experimenting on a physical model on a 1:40 scale and found that the nature of the turbulence differs between submerged and emerged groins. Cai et al. (2018) experimentally investigated the distribution patterns of flow velocity and water depth around groins and explored the effects of groin length and submergence level on flow force coefficient. ...
Article
Over the past few decades, eco-friendly channel construction has gained wide attention around the world including China, as it considers both navigational and ecological functions. The Groin system is a hydraulic structure that can effectively increase the area of fish habitats. Groin shape has an influence on the local flow field and the scour and deposition pattern near the structure, thereby affecting the adaptability of fishes to the groin fields. In this study, a new groin structure with a notch in its middle part was proposed. A lab experiment was conducted on a physical model to investigate the characteristics of the flow field including patterns of water level distribution, flow velocity distribution, and scour and deposition on different submergence levels and incoming flows. This experiment used four notch depths: 0 (unnotched), 1/3 of groin height, 2/3 of groin height, and full groin height. Then the distribution of habitat suitability index (HSI) was calculated based on the assessment index system of the indicator fish species (Cyprinidae). The results showed that the notches created in the middle of localized overflow groins had a significant effect on the surrounding flow pattern. Due to localized overflow, both the flow velocity and the velocity gradient behind each groin increased. The flow velocity behind each groin was positively correlated with notch depth, while the flow velocity in the main channel was negatively correlated with notch depth. Under the complex flow conditions, depositional features did not occur along the edges of the groin fields, ensuring connectivity between groin fields and the main channel. An analysis of water depth and flow velocity suitability indexes of Cyprinidae suggests when notch depth is 2/3 of groin height, the notched groins have good ecological effects, especially in the groin fields.