This study examines the influence of emergent vegetation on flow dynamics in natural river systems using large eddy simulation. It specifically addresses alterations in flow velocity and turbulence characteristics within open channels with varying densities of vegetation. The findings reveal a marked reduction in flow velocity within vegetated areas due to the obstruction posed by vegetation and the dynamics of shear layers. The flow alteration manifests in two distinct phases: the formation of a stable wake region behind densely vegetated patches and a more gradual velocity recovery in areas with sparser vegetation. In non-vegetated downstream zones, flow velocity tends to stabilize over a distance. Regarding turbulence, the study identifies differing patterns: enhanced vortex structures and accelerated energy dissipation in sparsely vegetated areas, contrasted with reduced turbulence in densely vegetated patches. Quadrant analysis further elucidates that ejections and inward interactions are primary contributors to Reynolds stress within vegetation patches, whereas outward interactions and sweeps become dominant in downstream regions. These insights offer a deeper understanding of how aquatic vegetation shapes river hydrodynamics, providing valuable information for effective river management and ecological restoration strategies.

  • Application of large eddy simulation method to investigate open-channel flows containing emergent vegetation patches.

  • Analysis of the adjustment process of flow structures under the influence of patches with different densities.

  • Clarification of material transport trends after water obstruction by patches using quadrant analysis.

Natural river environments, characterized by fluctuating inundation levels during flood and drought cycles, lead to the formation of diverse river channels and floodplains. These ecosystems support a variety of vegetation, including trees, shrubs, and grasses (Li et al. 2019; Ren et al. 2020; Aydogdu 2023). Aquatic vegetation, either fully submerged or partially exposed, plays a pivotal role in altering flow structures, affecting sediment deposition and resuspension, and shaping riverbed morphology. This contributes to the stability and resilience of aquatic ecosystems (Szabo-Meszaros et al. 2018; Wu et al. 2020; Zhang et al. 2020). By retaining, intercepting, and absorbing water flow through its roots, leaves, and stems, aquatic vegetation effectively reduces flow velocity, rate, and kinetic energy. This can mitigate scouring and erosion of riverbanks and beds, foster the growth of riparian vegetation, and decelerate water and soil loss (Tal & Paola 2007; Cornacchia et al. 2022). Moreover, it significantly influences the transport of pollutants, sediments, microorganisms, and other entities, both biotic and abiotic (Li et al. 2023; Wu et al. 2023; Zhan et al. 2024). Thus, understanding the impacts of aquatic vegetation on flow characteristics is essential for effective water ecological management.

Aquatic vegetation often forms non-uniformly distributed patches in natural and artificial river landscapes (Zhao & Huai 2016; Schoelynck et al. 2018; Li et al. 2019). The complex boundary shapes of these patches lead to intricate flow structures and turbulent characteristics around them (Follett & Nepf 2012; Li et al. 2022). The flow adjustments made by these patches alter local flow conditions in the river channel, affecting shear stress profiles and turbulent characteristics near the vegetation communities and bed area. This plays a decisive role in local bed morphology (Luhar et al. 2008; Le Bouteiller & Venditti 2015). Moreover, reduced flow velocity within the wake region behind these patches can encourage local deposition (Gurnell et al. 2012; Yamasaki et al. 2021). During flood periods, water inundates floodplains, resulting in shallow depths and low-flow velocity structures. When vegetation patches are too dense, they impede the flow significantly, leading to sediment deposition within the patches and hindering flood discharge. Conversely, sparse vegetation patches can trigger bed erosion under certain conditions, causing instability in the riverbank slopes.

The majority of relevant experimental studies mainly focused on the wake structure of flows around patches (Zong & Nepf 2016; Hu et al. 2018), which have multiple limitations. For example, due to the presence of dense cylinders in a patch, it is difficult to obtain clear measurement results by particle image velocimetry. The local flow field directly behind a cylinder is confined to a space that is often too small and tortuous for traditional probes to pass through (Zhang et al. 2019). In contrast, numerical simulations offer detailed insights into flow patterns throughout the entire domain.

Modern computer technology has enabled various mathematical models based on Navier–Stokes equations to effectively simulate turbulence structures of flows, aligning well with flume experimental data (Keramaris et al. 2017; Kim et al. 2018; Sonnenwald et al. 2019). Researchers have explored the pros and cons of different computational models. Kološ et al. (2020) conducted numerical simulations of the flow around a cylinder with a high Reynolds number. They calculated the wake drag coefficient, lift coefficient, pressure coefficient, and wake velocity field, and then compared shear stress transport kω model, large eddy simulation (LES) model, scale adaptive simulation model, and detached eddy simulation model. Finally, the best evaluation results were obtained from the LES model. The LES model yielded the best results, proving its efficiency despite the high computational costs.

Although numerical simulation methods have shown potential in studying flows under vegetation, research has mainly focused on turbulent kinetic energy and turbulent intensity distribution. Comprehensively and accurately analyzing energy changes in vegetation patches remains challenging. Detailed analyses of turbulence structures in vegetation patches are currently insufficient (Sonnenwald et al. 2017; Truong & Uijttewaal 2019; Li et al. 2020). More effective implementation of river ecological restoration by artificial approaches calls for a better understanding of how vegetation distribution impacts turbulence structures in river channels (Nikora et al. 2008; Huai et al. 2021).

The current study was conducted to explore the effects of emergent vegetation patches on open-channel flows. LES was performed to thoroughly analyze the flow velocity, turbulent kinetic energy, and vorticity profiles under single-vegetation patches with various densities. Power spectral density and quadrant analyses were used to elaborate the rules of flow energy change and the trends of material transport after water obstruction by the patches from a further perspective. The results of this study could provide a new angle for unraveling the changes in complex turbulence structures in the patch areas of emergent vegetation, in order to contribute meaningfully to future ecological restoration and redevelopment efforts, it is imperative to undertake a comprehensive investigation into the impact of vegetation distribution characteristics on water flow.

Numerical methods

The Navier–Stokes equations in Cartesian tensor form are typically expressed as
(1)
(2)
where ui is the velocity field, xi, xj are the fluid positions, fi is the mass force, is the fluid density, p is the pressure, t is the time, and is the fluid viscosity.
The integral-filtered equation can be expressed as follows:
(3)
where is the filter function. Integrating over a cubic domain with side length , the integral filter function can be expressed in the following form:
(4)
By substituting the aforementioned two equations into the Navier–Stokes equations, the filtered Navier–Stokes equations can be obtained:
(5)
(6)
where xi, xj are the fluid positions, , are the filtered velocity, and is the filtered mass force.

Equation (6) consists of an unknown quantity , which is associated with small-scale fluctuations filtered after application of the filter function. represents the filtered value of momentum flux per unit mass of fluid in turbulent motion. Because numerical simulation cannot acquire flow information across all scales, is unknown and a model for should be developed to close the LES equations. Then, by substituting it into Equation (6), the simplified filtered equation can be obtained.

Let , then Equation (6) is expressed as follows:
(7)
Similar to the form of Reynolds equation, Equation (3) has a non-closed term on the right-hand side:
(8)
where , commonly known as the subgrid stress, represents the momentum transport that occurs between subgrid-scale fluctuations removed by the filter function and the resolvable-scale turbulence. To achieve the closure of the numerical LES equations, a subgrid stress closure model must be built.
This study adopted a Wall-Adapting Local Eddy-viscosity (WALE) model as the subgrid model (Nicoud & Ducros 1999); the WALE model is a composite model that considers turbulent wall effects and momentum transfer, among other factors. For laminar shear flow, the turbulent viscosity in the WALE model automatically approaches zero, ensuring accuracy in numerical simulations of near-wall flow fields. It is suitable for high-gradient regions such as boundary layer flow and turbulent separation, exhibiting better performance in regions with finer turbulent structures. In the model, the eddy-viscosity coefficient is defined as follows:
(9)
where is the subgrid-scale mixing length, , d is the distance to the closest wall, V is the volume of the computational cell, and is the rate-of-strain tensor for the resolved scale:
(10)
(11)
where is the von Kármán constant, .

This study referenced previous studies, setting the value of the constant of the wall-adapting local eddy-viscosity model to , aiming to achieve more accurate results within a wide range of flow rates (Kim et al. 2019).

Model development and computational methods

The computational domain for the simulation was 3.6 m long × 0.5 m wide × 0.2 m high. The emergent vegetation patch was generalized into an array of rigid smooth-surfaced cylinders with a diameter of 0.5 cm. These vegetation patches were centrally located in the computational domain, 0.5 m from the inlet, longitudinally. The coordinate system originated at the domain's lower-left corner, with axes oriented longitudinally (x-axis), transversally (y-axis), and vertically upward (z-axis) from the bed. The specific layout is shown in Figure 1.
Figure 1

Schematic diagram of the computational model: (a) three-dimensional view and (b) top view.

Figure 1

Schematic diagram of the computational model: (a) three-dimensional view and (b) top view.

Close modal
In the computational domain, the horizontal plane at a mid-height of the vegetation patch was selected as the monitoring sections. Additionally, five equidistant sections (#1–#5) were set in the patch area, with a spacing of 0.025 m, and one more section (#6) 1D downstream from the patch, as shown in Figure 2.
Figure 2

Setup of the monitoring surfaces from top view.

Figure 2

Setup of the monitoring surfaces from top view.

Close modal

Simulation calculation was carried out for two groups of circular emergent vegetation patch. To simulate the vegetation patch, rigid cylinders (diameter d = 0.5 cm, height h = 20 cm) were neatly arranged in a circular range (diameter D = 10 cm). At this time, the blockage ratio of the cross-section where the patch was located was β = D/B = 0.2, where B is the width of the flume. This ratio ensures a minimal impact of flume sidewall vortices on the study area's vortex shedding, according to Sahin & Owens (2004) and Kumar & Mittal (2006).

The patch density is defined as , where a is the upstream area of cylinders in per unit volume; a=nd, with units of cm−1, where n is the number of cylinders in per unit area. Here, n signifies the number of cylinders per unit area, with units of cm−2, and d denotes the diameter of individual rigid cylinders, with units of cm. The calculation of Reynolds number based on patch size (D) and the size of single cylinders in the patch (d) is as follows: and . Two patch densities, ϕ = 0.05 and ϕ = 0.17, were set for the simulation, as shown in detail in Figure 3 and Table 1.
Table 1

Model parameters

ConditionParameter
Patch density, ϕMean inlet velocity, V (m/s)Patch-scale Reynolds number, ReDPlant-scale Reynolds number, RedFroude number, Fr
0.05 0.50 99,403.58 2,485.09 0.36 
0.17 0.50 99,403.58 2,485.09 0.36 
ConditionParameter
Patch density, ϕMean inlet velocity, V (m/s)Patch-scale Reynolds number, ReDPlant-scale Reynolds number, RedFroude number, Fr
0.05 0.50 99,403.58 2,485.09 0.36 
0.17 0.50 99,403.58 2,485.09 0.36 
Figure 3

Top view of two vegetation patch arrangements with different densities: (a) patch arrangement with ϕ = 0.05 and (b) patch arrangement with ϕ = 0.17.

Figure 3

Top view of two vegetation patch arrangements with different densities: (a) patch arrangement with ϕ = 0.05 and (b) patch arrangement with ϕ = 0.17.

Close modal

In this study, the LES turbulence model, wall-adapting local eddy-viscosity subgrid-scale turbulence model, and finite volume method were used to discretely solve Navier–Stokes equations. In the numerical calculation, the smallest grid is located at the position of the vegetation patch stem, with a scale of 0.005 m. The velocity in the vicinity is ∼0.05 m/s, resulting in a ratio of 0.02 s. To ensure computational accuracy and convergence, the time step should be less than or equal to one-fifth of the ratio between the smallest grid size and the surrounding flow velocity. Therefore, a time step of 0.01 s is set in this study. Additionally, considering that in previous studies, the minimum residual value for iterations is typically set to 1 × 10−4 to achieve strict accuracy, the minimum residual for each equation is set to 1 × 10−4 in this study (Xiang et al. 2020). Furthermore, the computational model was configured to perform a maximum of 60 iterations per time step. Iterations were designed to cease when the equation's convergence criteria, defined as the residual falling below the aforementioned minimum threshold, were satisfied. This process then progressed to the subsequent time step for further iterations. The boundary conditions were set as follows: velocity inlet boundary at the inlet, pressure outlet boundary at the downstream outlet, free water surface boundary at the top water surface, and no-slip wall boundary at the bottom and sides, including vegetation surface.

For the LES calculations, precise mesh generation was imperative. Adhering to the boundary layer requirement of y+ ≤ 1, all-hexahedral meshes with specialized refinement in the vegetation patch area were generated. This approach included dividing the walls of individual cylinders within the patch into five distinct layers of boundary mesh to ensure detailed and accurate LES simulations. In this calculation, the inlet velocity is V = 0.5 m/s, and the model length is 3.6 m. The total calculation time is 40 s, which corresponds to the water flowing through five model lengths to ensure stability in the computational model. The computed data are then sampled, and for subsequent flow field calculations and analyses, physical quantities such as flow velocity are averaged over 1-min sampling intervals.

Model verification

To validate the reliability of the numerical model used in this study, validation experiments were conducted under conditions approximating those of the numerical model, concentrating on vegetation patches of two different densities (Figure 4).
Figure 4

Experimental instrument and equipment.

Figure 4

Experimental instrument and equipment.

Close modal

Based on the research content of the simulation, the mean velocity at the inlet was set to U0 = 0.2 m/s as a verification example. The model verification experiment was carried out in the Multiphase Flow and Flow Visualization Laboratory of Xi'an University of Technology (Xi'an, Shanxi Province, China). The experiments were conducted in a rectangular flume of dimensions 7.50 m × 0.30 m × 0.45 m (length × width × height), manufactured by Armfield (MODEL S6-MKII), with the sidewalls made of glass and a fixed bed slope of 0.1%. Polymethyl methacrylate (PMMA) plates, forming a stable base, were installed across the flume's bottom for rigid vegetation simulation, using PMMA rods. Two types of vegetation patches with different densities (dense patch: ϕ = 0.17, sparse patch: ϕ = 0.05) were arranged, with the diameter of the patches adjusted to achieve a blockage ratio (β) of 0.2 relative to the width of the cross-section. The inflow rate was set to 9.15 L/s, with an average flow velocity (U0) of 0.2 m/s at the cross-section. Flow velocities were measured using an acoustic Doppler velocimeter.

For verification purposes, the mid-height of the longitudinal measuring surface was selected for observation. The numerical simulation results were then juxtaposed with experimental results for mean streamwise velocity profiles. Figure 5 illustrates the comparative analysis of longitudinal mean velocity profiles derived from both these approaches.
Figure 5

Comparison of mean streamwise velocity profiles along the longitudinal direction: (a) ϕ = 0.05, U0 = 0.20 m/s and (b) ϕ = 0.17, U0 = 0.20 m/s.

Figure 5

Comparison of mean streamwise velocity profiles along the longitudinal direction: (a) ϕ = 0.05, U0 = 0.20 m/s and (b) ϕ = 0.17, U0 = 0.20 m/s.

Close modal

Analysis of the acquired data revealed that the relative errors between the numerical simulations and experimental results were below 5%, demonstrating a high degree of concordance in velocity profiles. This alignment suggests the appropriateness of the turbulence model and calculation parameters employed in the numerical study for accurately simulating open-channel flows with vegetation patches. It was noted, however, that the numerical simulation results were marginally higher than the experimental outcomes. This discrepancy can be attributed to the use of plexiglass for vegetation simulation in the experiments, contrasting with the no-slip solid wall boundary condition applied to the vegetation surface in the simulations. Despite employing the wall-function method for near-wall region corrections, slight variances from the experimental results persisted.

Velocity analysis

Figure 6 displays the mean longitudinal velocity profiles of the horizontal section at the mid-height of the patch (Z = 0.1 m), where the vector arrow only represents the velocity direction. At lower patch densities (ϕ = 0.05), the sparse distribution of vegetation minimally obstructs the flow, leading to independent wakes and rapid velocity recovery. This suggests that in sparsely vegetated areas, the impact on flow dynamics is localized and less pronounced. Conversely, at higher densities, the flow encounters significant resistance, creating complex flow patterns around and through the vegetation patches. When the upstream inflow encounters the patch, it forms the flow around the patch and through flow. A patch-scale longitudinal shear layer is generated on both sides of the vegetation patch, which develops in the streamwise direction and gradually widens until intersecting and interacting with the shear layer on both sides. This is mainly attributable to gradually enhanced lateral outflow through the patch, which has been defined as the ‘analog mixing layer’ by Li et al. (2020), and this region is controlled by local coherent vortices. The emergence of patch-scale longitudinal shear layers and the subsequent formation of recirculation regions highlight the substantial influence of vegetation density on flow modification. These results underline the critical role of vegetation structure in shaping river flow dynamics and underscore the importance of accurately modeling vegetation in hydrodynamic studies for effective river management and ecological restoration planning.
Figure 6

Horizontal profiles of longitudinal mean velocity under different conditions: (a) condition 1 and (b) condition 2.

Figure 6

Horizontal profiles of longitudinal mean velocity under different conditions: (a) condition 1 and (b) condition 2.

Close modal
Considering the flow impacts due to vegetation patches, sections every 0.1 m along the x-axis, starting from x = 0.2 m, were analyzed. The lateral profiles of mean streamwise velocity both in the vegetated and non-vegetated areas are depicted (Figure 7), with vegetation patches marked as shaded areas. Notably, velocity adjustment begins near the patch entrance, with deceleration starting 0.2 m upstream, a phenomenon identified as the ‘upstream impact region’ (Belcher et al. 2003). Similar to the findings of Li et al. (2019) and Zong & Nepf (2016), we also observe significant variations in the longitudinal velocity distribution under the influence of vegetation patches of two different densities. Upon entering the patch, there is a notable deceleration, particularly in the upstream area. In low-density patches, the minimum velocity occurs at the patch exit, decelerating to ∼21.7% of the inflow velocity, before rapidly recovering to 91% within a distance of 0.4 m downstream, followed by a gradual return to normal. Conversely, high-density patches exhibit the lowest velocity 0.1 m downstream of the patch, experiencing a maximum deceleration of ∼42.0%, and a brief period of sustained low velocity before a gradual recovery.
Figure 7

Lateral profiles of mean streamwise velocity at specific sections in different areas.

Figure 7

Lateral profiles of mean streamwise velocity at specific sections in different areas.

Close modal

Zong & Nepf (2016) defined the region downstream of a patch, where longitudinal mean velocity continues to decrease along the streamwise direction as the ‘steady wake region’. In this region, both the mean velocity and turbulence intensity are significantly reduced. The flow in the steady wake region mainly originates from the longitudinal outflow through the patch. The formation of the steady wake region is additionally tied to the horizontal shear layers formed at the outer edges on both sides of the patch. These two shear layers act as barriers for the wake region. They separate the high-velocity flow past both sides of the vegetation patch from the low-velocity flow in the core zone of the wake, allowing the wake region to maintain low momentum levels within certain longitudinal distances. The above results indicate that no steady wake region is formed downstream of the sparse patch, whereas a ∼0.1-m-long steady wake region exists in the dense patch. Similar to the results of Chang & Constantinescu (2015), we also observed a patch-scale steady wake region downstream of the dense patch, indicating pronounced cylinder–wake interactions within this area.

The velocity in the non-vegetated area increases slightly as a consequence of the flow around the patch and lateral outflow through the patch. This region is defined as the ‘uniform region in the non-vegetated area’ by Li et al. (2020). With low patch density, the highest velocity in the non-vegetated area emerges at the exit of the patch, and the maximum acceleration rate is ∼6.5% of the inflow velocity. After passing through the patch, the velocity slowly decreases within a distance of 0.3 m and then levels off. With a high patch density, the highest flow velocity is found at the exit of the patch, with a maximum acceleration rate of ∼14.0%. After passing through the patch, the velocity slowly recovers at low rates over a long distance. Consistent with the findings of Zong & Nepf (2016), the velocity continues to decelerate after entering the patch, with the velocity at the patch exit being lower than that at the patch entrance. In principle, the wake flow cannot fully recover as the momentum loss due to cylinder drag remains constant downstream of the patch.

Figure 8 depicts the mean velocity profiles at different characteristic sections, with the patch area denoted by dotted lines. At the entrance (section #1) of patches with varying densities, velocity profiles exhibit similar characteristics. In low-density patches, the inflow is primarily disturbed at the patch site due to the substantial spacing between internal vegetation, resulting in minimal impact on the non-vegetated areas within the patch. The flow converges at section #6 downstream, facilitating rapid velocity recovery. Conversely, in the regions flanking the patch, the impact is negligible, with no significant velocity changes observed. At section #6, located 1D downstream, the velocity generally resumes to the inflow velocity level. Conversely, at a patch density of 0.17, the inflow is significantly dispersed to both sides of the patch, leading to a noticeable increase in velocity in these lateral channels. This condition creates a distinct demarcation between the analog mixing layer and the uniform region in the non-vegetated area. Velocity recovery at section #6 downstream of the vegetation patch under this high-density condition () is slower compared with the low-density scenario. As previously mentioned, the horizontal shear layers expand in width along the streamwise direction and intersect at a certain distance downstream of the patch. This intersection point aligns with the terminus of the steady wake region, where the wake's velocity commences its recovery.
Figure 8

Velocity profiles at different monitoring surfaces in the patch area under various conditions: condition 1 (ϕ = 0.05, V = 0.50 m/s) and condition 2 (ϕ = 0.17, V = 0.50 m/s).

Figure 8

Velocity profiles at different monitoring surfaces in the patch area under various conditions: condition 1 (ϕ = 0.05, V = 0.50 m/s) and condition 2 (ϕ = 0.17, V = 0.50 m/s).

Close modal

Analysis of turbulent characteristics

Vorticity analysis

When the flow passes a single cylinder, it would generate flows on both sides of the cylinder, accompanied by boundary layer separation on both sides (if the Reynolds number is high enough). Then, the Kelvin–Helmholtz instability due to shear begins to develop, which causes periodic vortex shedding downstream of the cylinder, resulting in a Karman vortex street. Generally, when the cylinder has a slenderness ratio h/D < 2, it is thought that the periodic vortex shedding behavior downstream of the cylinder would be significantly suppressed (Lee et al. 2007). Considering the overall slenderness ratio of the circular vegetation patch simulated in this study (h/D = 2), patch-scale Karman vortex street can be observed downstream of the vegetation patch.

To clarify the generation and dissipation processes of vortex structures in the flow field, this study introduces the Q-criteria based on vortex identification methods. The flow field with Q-criterion values of 0–1 is selected to generate vorticity contours, and the distributions of vorticity contours under vegetation patches with two different densities are shown (Figure 9). It is observed that the flow around a cylinder is generated under the action of vegetation patch and develops downstream to form different wake flow patterns. In the presence of vegetation, the vorticity field of flow has strongly enhanced intensity, and a relatively remarkable coherent structure is formed behind the patch. Figure 9 (left) shows the condition of low patch density, under which large-scale vortex shedding is observed downstream of the patch due to the long distance between vegetation cylinders in the patch, forming a relatively long wake region. Under the condition of high patch density (Figure 9, right), shedding vortices generated by the vegetation patch would be suppressed due to the dense distribution and small distance of vegetation cylinders. Meanwhile, more and denser vortices would be generated in the flow field, resulting in more complicated flow structures in the flow field and posing more intensive impacts on the walls of both sides.
Figure 9

Distributions of vorticity contours based on Q-criteria under characteristic conditions.

Figure 9

Distributions of vorticity contours based on Q-criteria under characteristic conditions.

Close modal

Reynolds stress profiles

In this study, flow field conditions are less impacted in the vertical direction, so the analysis of Reynolds stress is focused in the x and y directions. Figure 10 displays the Reynolds stress profiles in the patch area under various conditions. In both cases, large Reynolds stress emerges in the areas on both lateral sides of the patch, corresponding to the ‘analog mixing layer’ mentioned in Section 3.1, which indicates the occurrence of strong shear. In the interior and downstream areas of the patch, the Reynolds stress differs considerably under various patch densities. When the patch density is low, different single cylinders in the patch exhibit are relatively independent effects, and Reynolds stress only shows a relatively evident shear effect in the lateral direction. Under high patch density, the upstream face on both sides of the patch would have a greater impact range and stronger shear effect on the inflow. Moreover, the generation of remarkable Reynolds stress is observed in the interior of the patch area. In summary, the Reynolds stress profile in the flow near the dense patch is more dispersed and intensive, which indicates more complicated momentum exchange processes.
Figure 10

Reynolds stress profiles in the patch area under different conditions: (a) condition 1 (ϕ = 0.05, V = 0.50 m/s) and (b) condition 2 (ϕ = 0.17, V = 0.50 m/s).

Figure 10

Reynolds stress profiles in the patch area under different conditions: (a) condition 1 (ϕ = 0.05, V = 0.50 m/s) and (b) condition 2 (ϕ = 0.17, V = 0.50 m/s).

Close modal

Analysis of energy structure changes

Power spectral density analysis

As stated previously, significant Reynolds stress occurs both upstream and downstream of the vegetation patch, as depicted in Figure 10. Consequently, these areas have been identified as key characteristic points for detailed velocity monitoring within the vegetation patch zone, illustrated in Figure 11.
Figure 11

Layout of monitoring points.

Figure 11

Layout of monitoring points.

Close modal
Figure 12 depicts the power spectral density curves of the longitudinal velocity time series at various characteristic points, examining vortex frequency changes under different vegetation patch densities. Figure 12(a) shows that the greater obstruction effect on flow results in higher velocity diversion. In contrast, lower patch densities allow faster passage of flow through the patch, leading to vigorous momentum exchange and rapid energy dissipation, with high-frequency water adhering to Kolmogorov's ‘ − 5/3’ power law. Figure 12(c) presents the spectral curves of longitudinal velocity at the center of the patch. Based on the above analysis, the velocity at the center position is relatively low under the action of the patch, causing no evident disturbances. This is manifested in the spectral curves, which exhibit relatively steady trends of energy changes with no intense energy dissipation process. Figure 12(b) shows that the velocity changes substantially at the exit of the patch, where there are remarkably high Reynolds stress and turbulence intensity. This accounts for intense momentum exchange between water and fast energy dissipation rate. Moreover, the flow velocity starts to recover gradually at the downstream characteristic point, where there is no strong momentum exchange between water, and as such, the energy dissipation rate is low. Consequently, high-frequency water also carries great energy.
Figure 12

Power spectral density curves of longitudinal velocity time series at the characteristic points of vegetation patch under different conditions: (a) front end of the patch, (b) exit and 1D downstream of the patch, and (c) central point of the patch.

Figure 12

Power spectral density curves of longitudinal velocity time series at the characteristic points of vegetation patch under different conditions: (a) front end of the patch, (b) exit and 1D downstream of the patch, and (c) central point of the patch.

Close modal

Overall, when the patch density is relatively low, the energy carried by water with low-frequency bands decreases linearly under the log–log coordinate system. In contrast, the decrease in the energy carried by water with low-frequency bands slows down under relatively high patch density. This provides evidence that the water is subjected to greater resistance under relatively high patch density. The underlying reason is that under low vegetation density, the gaps between vegetation allow the vortices generated by disturbances to fully develop. Due to drastic momentum exchange and energy dissipation, the high energy carried by large vortices is rapidly transferred to small vortices, or they are decomposed into high-frequency small vortices. The mutual transfer and conversion processes of energy are accompanied by considerable energy consumption. In particular, a large amount of energy is consumed by the separation of boundary layers and the diffusion of wake vortex at the downstream exit of the patch. With a high density of vegetation patch, the turbulence of water is limited and the energy dissipation occurs slowly, so that the momentum exchange between water does not reach a very high level.

Quadrant analysis

The concept of quadrant analysis of Reynolds stress, first proposed by Lu & Willmarth (1973), is utilized to characterize additional shear stress at the viscous sublayer in turbulent flow. Based on two components of instantaneous fluctuating velocity, and , this method depicts velocity fluctuations in a four-quadrant Cartesian diagram, with the longitudinal fluctuating velocity as the abscissa and the transverse fluctuating velocity as the ordinate. It also allows for evaluating the contribution of each event to the mean Reynolds stress, i.e., . Then, each quadrant is associated with a different type of turbulence event and a specific form of momentum transfer: Q1: , , which belongs to the flow pattern of outward interactions; Q2: , , which represents the flow pattern of ejections; Q3: , , which pertains to the flow pattern of inward interactions; and Q4: , , which belongs to the flow pattern of sweeps. In addition to the above-mentioned four flow patterns, a fifth action called the ‘hole’ is usually defined in the quadrant analysis of Reynolds stress. For the shaded area in Figure 13, the ‘hole’ region is the excluded area of the (, ) plane. The shaded area is bounded by the hyperbolas , and its size H is the threshold value. This allows quadrant analysis to determine the relative importance of intermittent events. Its definition is expressed as follows:
(12)
where and are the root mean squares of fluctuating velocity components u and v, respectively.
Figure 13

Schematic diagram of the ‘hole’ region (Carvalho & Aleixo 2014).

Figure 13

Schematic diagram of the ‘hole’ region (Carvalho & Aleixo 2014).

Close modal
As H increases, larger parts of the plane are excluded from the analysis, and the conditions for Reynolds stress are averaged over larger and less frequent values. This approach can eliminate minor contributions within hyperbolic regions, and only major contributions from events associated with each quadrant are considered. Outside the hyperbolic region of size H, the contribution to the total Reynolds stress in quadrant i is given by the following equation (Poggi et al. 2004):
(13)
where is located in the quadrant, i represents the ith quadrant, T is the sampling period of each monitoring point, and is the detection function, which is defined as follows:
(14)
Under different threshold parameters, the contribution of various flow patterns to Reynolds stress, , is expressed by the following equation:
(15)
At H = 0, the sum of the percentage contributions of all events at the monitoring point is equal to 1:
(16)
Based on the principle of quadrant analysis, Figure 14 illustrates the distribution of the contribution ratio of each quadrant to Reynolds stress at specific monitoring points at H = 0. It can be observed that at the entrance of the patch (Figure 14(a), monitoring points 1 and 2), the inflow is obstructed by the patch, resulting in strong shearing. Its Reynolds stress is equally distributed in opposite directions at the edges on both sides of the patch. This distribution is reflected in the graphs of quadrant contribution ratios: under the two conditions, the Reynolds stress at the monitoring point #1 is contributed by Q1, whose flow pattern belongs to outward interactions and is favorable for the upward diffusion of momentum. At the monitoring point #2, the Reynolds stress is contributed by Q4, whose flow pattern belongs to sweeps, with downward diffusion of momentum.
Figure 14

Contribution ratios of various quadrants to Reynolds stress at the monitoring points under different conditions.

Figure 14

Contribution ratios of various quadrants to Reynolds stress at the monitoring points under different conditions.

Close modal

At the central position of the patch (Figure 14(b), monitoring point #3), the large difference in patch density between the two conditions results in remarkable variation in the major quadrants contributing to momentum exchange. In the sparse patch, the dominant flow pattern is outward interactions, whereas in the dense patch, ejections are predominant with a contribution ratio of ∼57%, in agreement with the observations of Poggi et al. (2004). At the junction of the patch exit and the flow (Figure 14(c), monitoring point #4), in contrast to the sparse patch dominated by outward interactions, intensive momentum exchange occurs in the flow in the dense patch, and the greatest contribution comes from ejections. Sweeps also affect the generation of shear stress, with a contribution ratio of 12%. In the downstream area of the patch (Figure 14(d), monitoring point #5), the flow field starts to recover gradually and its momentum exchange is primarily contributed by outward interactions and sweeps. Moreover, sweeps play a more prominent role under the condition of high patch density.

To further understand the momentum exchange process in the patch area, the representative distributions of (, ) based on quadrant analysis are shown for the monitoring points #3–#5 (Figure 15). Under the two conditions, the momentum exchange that occurs within the patch and at the interface between the patch and the water is mainly governed by ejections and inward interactions, which mirrors the patterns reported previously (Pang 2016). However, in addition to outward interactions, the contribution of sweeps is notable in the development region downstream of the patch, especially under the condition of dense patch. This also provides further evidence supporting the important role of sweeps.
Figure 15

Representative distributions of (, ) at selected monitoring points of vegetation patch: (a) point #3 under condition 1, (b) point #3 under condition 2, (c) point #4 under condition 1, (d) point #4 under condition 2, (e) point #5 under condition 1, and (f) point #5 under condition 2.

Figure 15

Representative distributions of (, ) at selected monitoring points of vegetation patch: (a) point #3 under condition 1, (b) point #3 under condition 2, (c) point #4 under condition 1, (d) point #4 under condition 2, (e) point #5 under condition 1, and (f) point #5 under condition 2.

Close modal

The study enhances the understanding of how emergent vegetation affects flow in open channels. It is important to note that in the natural world, aquatic vegetation, due to its diverse modes of existence such as submerged, emergent, and floating forms, and its morphological attributes such as stems, stalks, and leaves, typically exhibits a range of distinct external morphologies and structural resilience, accurately characterizing real vegetation presents significant challenges. The scope of this study is confined to emergent rigid vegetation in shallow, low-velocity channels. A reasonable approach is to represent vegetation as idealized cylinders. With this simplification, this study is confined to idealized vegetation elements, characterized solely by their diameter and height, without considering additional structural complexities. While the artificial vegetation patches employed in this study do not exhibit the intricate characteristics of natural vegetation, this methodological approach provides a controlled experimental setting that enables the differentiation of the effects of diverse vegetation parameters on hydrodynamic processes. The findings of this study provide foundational insights into the general principles of flow–vegetation interactions, aiding in a broader understanding of plant effects on fluid dynamics and laying the groundwork for more detailed investigations in the real world.

This study demonstrates the impacts of emergent vegetation patches on velocity, turbulence, and energy exchange characteristics of open-channel flows using the numerical method of LES. The major conclusions are as follows:

  • 1. Impact on flow velocity: The study demonstrates that vegetation patches markedly influence the flow field's velocity profiles. Specifically, a substantial decrease in flow velocity within vegetated areas was observed, with the dense patch showing a maximum deceleration rate of ∼42.0% at 0.1 m downstream. This starkly contrasts with the sparse patch, where the maximum deceleration rate is ∼21% at the patch's exit, signifying the profound impact of vegetation density on flow behavior.

  • 2. Turbulence and vorticity: The presence of vegetation notably intensifies the vorticity field, leading to coherent structure formation behind the patch. This effect varies between sparse and dense patches, with the former facilitating large-scale vortex shedding and the latter inhibiting it. The study's detailed analysis reveals a significant increase in Reynolds stress in the patch area, especially under dense vegetation, highlighting the intricate relationship between vegetation density and turbulence.

  • 3. Dominant flow patterns and Reynolds stress: The analysis identifies ejections and inward interactions as the predominant flow patterns at the patch site, contributing to about 80% of the total Reynolds stress. Interestingly, in the dense patch, the interior Reynolds stress is mainly due to ejections (∼57%), while in the downstream development region, outward interactions and sweeps are more prominent.

In the future, it will be of great significance to study how flow characteristics vary with changes in patch shape. Relevant research results could have practical implications for determining appropriate vegetation arrangement methods (e.g., patch size, density, interior shape) in ecological restoration and river channel management.

This research was supported by the National Key Technologies R&D Program of China (No. 2022YFF1300803), the National Natural Science Foundation of China (No. U2243201), and the Joint Open Research Fund Program of State Key Laboratory of Hydro-science and Engineering and Tsinghua – Ningxia Yinchuan Joint Institute of Internet of Waters on Digital Water Governance (sklhse-2023-Iow06).

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

The authors declare there is no conflict.

Belcher
S. E.
,
Jerram
N.
&
Hunt
J. C. R.
2003
Adjustment of a turbulent boundary layer to a canopy of roughness elements
.
Journal of Fluid Mechanics
488
,
369
398
.
Carvalho
E.
&
Aleixo
R.
2014
Quadrant method application to the study of the beginning of sediment motion of sedimentary particles
.
MEFTE
2014
,
189
.
Cornacchia
L.
,
Riviere
N.
,
Soundar Jerome
J. J.
,
Doppler
D.
,
Vallier
F.
&
Puijalon
S.
2022
Flow and wake length downstream of live submerged vegetation patches: How do different species and patch configurations create sheltering in stressful habitats?
Water Resources Research
58
(
3
),
e2021WR030880
.
Follett
E. M.
&
Nepf
H. M.
2012
Sediment patterns near a model patch of reedy emergent vegetation
.
Geomorphology
179
,
141
151
.
Huai
W. X.
,
Li
S.
,
Katul
G. G.
,
Liu
M. Y.
&
Yang
Z. H.
2021
Flow dynamics and sediment transport in vegetated rivers: A review
.
Journal of Hydrodynamics
33
(
3
),
400
420
.
Keramaris
E.
,
Kasiteropoulou
D.
,
Liakopoulos
A.
,
Michalolias
N.
&
Pechlivanidis
G.
2017
A study of flow in open channels with vegetation: Experiments and numerical models
.
Special Topics & Reviews in Porous Media: An International Journal
8
(
2
), 109–120.
Kim
M.
,
Lim
J.
,
Kim
S.
,
Jee
S.
,
Park
J.
&
Park
D.
2019
Large-eddy simulation with parabolized stability equations for turbulent transition using OpenFOAM
.
Computers & Fluids
189
,
108
117
.
Kološ
I.
,
Lausová
L.
&
Michalcová
V.
2020
RANS and SRS simulations of the flow around a smooth cylinder
. In
MATEC Web of Conferences
, Vol.
310
.
EDP Sciences
, p.
00031
.
Kumar
B.
&
Mittal
S.
2006
Effect of blockage on critical parameters for flow past a circular cylinder
.
International Journal for Numerical Methods in Fluids
50
(
8
),
987
1001
.
Le Bouteiller
C.
&
Venditti
J. G.
2015
Sediment transport and shear stress partitioning in a vegetated flow
.
Water Resources Research
51
(
4
),
2901
2922
.
Lee
T.
,
Lin
C. L.
&
Friehe
C. A.
2007
Large-eddy simulation of air flow around a wall-mounted circular cylinder and a tripod tower
.
Journal of Turbulence
8 (29),
29
.
Li
W. Q.
,
Wang
D.
,
Jiao
J. L.
&
Yang
K. J.
2019
Effects of vegetation patch density on flow velocity characteristics in an open channel
.
Journal of Hydrodynamics
31
(
5
),
1052
1059
.
Li
J.
,
Claude
N.
,
Tassi
P.
,
Cordier
F.
,
Vargas-Luna
A.
,
Crosato
A.
&
Rodrigues
S.
2022
Effects of vegetation patch patterns on channel morphology: A numerical study
.
Journal of Geophysical Research: Earth Surface
127
(
5
),
e2021JF006529
.
Li
G.
,
Gong
Z.
,
Jiang
W.
,
Zhan
J.
,
Wang
B.
,
Fu
X.
,
Xu
M.
&
Wu
Z.
2023
Environmental transport of gyrotactic microorganisms in an open-channel flow
.
Water Resources Research
59
(
4
),
e2022WR033229
.
Lu
S. S.
&
Willmarth
W. W.
1973
Measurements of the structure of the Reynolds stress in a turbulent boundary layer
.
Journal of Fluid Mechanics
60
(
3
),
481
511
.
Luhar
M.
,
Rominger
J.
&
Nepf
H.
2008
Interaction between flow, transport and vegetation spatial structure
.
Environmental Fluid Mechanics
8
,
423
439
.
Nicoud
F.
&
Ducros
F.
1999
Subgrid-scale stress modelling based on the square of the velocity gradient tensor
.
Flow, Turbulence and Combustion
62
(
3
),
183
200
.
Nikora
V.
,
Larned
S.
,
Nikora
N.
,
Debnath
K.
,
Cooper
G.
&
Reid
M.
2008
Hydraulic resistance due to aquatic vegetation in small streams: Field study
.
Journal of Hydraulic Engineering
134
(
9
),
1326
1332
.
Pang, C.
2016
Application of quadrant analysis in analyzing sediment transport in flow with submerged plants
.
Hydro-Science and Engineering
37, (
3
),
20
226
.
Poggi
D.
,
Porporato
A.
,
Ridolfi
L.
,
Albertson
J. D.
&
Katul
G. G.
2004
The effect of vegetation density on canopy sub-layer turbulence
.
Boundary-Layer Meteorology
111
,
565
587
.
Ren
B. L.
,
Wang
D.
,
Li
W. Q.
&
Yang
K. J.
2020
The velocity patterns in rigid and mobile channels with vegetation patches
.
Journal of Hydrodynamics
32
,
561
569
.
Schoelynck
J.
,
Creëlle
S.
,
Buis
K.
,
De Mulder
T.
,
Emsens
W. J.
,
Hein
T.
&
Folkard
A.
2018
What is a macrophyte patch? Patch identification in aquatic ecosystems and guidelines for consistent delineation
.
Ecohydrology & Hydrobiology
18
(
1
),
1
9
.
Sonnenwald
F.
,
Hart
J. R.
,
West
P.
,
Stovin
V. R.
&
Guymer
I.
2017
Transverse and longitudinal mixing in real emergent vegetation at low velocities
.
Water Resources Research
53
(
1
),
961
978
.
Sonnenwald
F.
,
Guymer
I.
&
Stovin
V.
2019
A CFD-based mixing model for vegetated flows
.
Water Resources Research
55
(
3
),
2322
2347
.
Szabo-Meszaros
M.
,
Navaratnam
C. U.
,
Aberle
J.
,
Silva
A. T.
,
Forseth
T.
,
Calles
O.
&
Alfredsen
K.
2018
Experimental hydraulics on fish-friendly trash-racks: An ecological approach
.
Ecological Engineering
113
,
11
20
.
Truong
S. H.
&
Uijttewaal
W. S. J.
2019
Transverse momentum exchange induced by large coherent structures in a vegetated compound channel
.
Water Resources Research
55
(
1
),
589
612
.
Wu
Y. J.
,
Jing
H. F.
,
Li
C. G.
&
Song
Y. T.
2020
Flow characteristics in open channels with aquatic rigid vegetation
.
Journal of Hydrodynamics
32
(
6
),
1100
1108
.
Wu
Z.
,
Jiang
W.
,
Zeng
L.
&
Fu
X.
2023
Theoretical analysis for bedload particle deposition and hop statistics
.
Journal of Fluid Mechanics
954
,
A11
.
Xiang
K.
,
Yang
Z.
,
Wu
S.
,
Gao
W.
,
Li
D.
&
Li
Q.
2020
Flow hydrodynamics of the mixing layer in consecutive vegetated groyne fields
.
Physics of Fluids
32
(
6
), 065110.
Yamasaki
T. N.
,
Jiang
B.
,
Janzen
J. G.
&
Nepf
H. M.
2021
Feedback between vegetation, flow, and deposition: A study of artificial vegetation patch development
.
Journal of Hydrology
598
,
126232
.
Zhang
J.
,
Liang
D.
,
Fan
X.
&
Liu
H.
2019
Detached eddy simulation of flow through a circular patch of free-surface-piercing cylinders
.
Advances in Water Resources
123
,
96
108
.
Zhang
Y.
,
Lai
X.
,
Zhang
L.
,
Song
K.
,
Yao
X.
,
Gu
L.
&
Pang
C.
2020
The influence of aquatic vegetation on flow structure and sediment deposition: A field study in Dongting Lake, China
.
Journal of Hydrology
584
,
124644
.
Zhao
F.
&
Huai
W.
2016
Hydrodynamics of discontinuous rigid submerged vegetation patches in open-channel flow
.
Journal of Hydro-Environment Research
12
,
148
160
.
Zong
L.
&
Nepf
H.
2016
Vortex development behind a finite porous obstruction in a channel
.
Journal of Fluid Mechanics
691
,
368
391
.
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/).