Abstract
The need for operational models describing the friction factor f in streams remains undisputed given its utility across a plethora of hydrological and hydraulic applications concerned with shallow inertial flows. For small-scale roughness elements uniformly covering the wetted parameter of a wide channel, the Darcy-Weisbach f = 8(u*/Ub)2 is widely used at very high Reynolds numbers, where u* is friction velocity related to the surface kinematic stress, Ub = Q/A is bulk velocity, Q is flow rate, and A is cross-sectional area orthogonal to the flow direction. In natural streams, the presence of vegetation introduces additional complications to quantifying f, the subject of the present work. Turbulent flow through vegetation are characterized by a number of coherent vortical structures: (i) von Karman vortex streets in the lower layers of vegetated canopies, (ii) Kelvin-Helmholtz as well as attached eddies near the vegetation top, and (iii) attached eddies well above the vegetated layer. These vortical structures govern the canonical mixing lengths for momentum transfer and their influence on f is to be derived. The main novelty is that the friction factor of vegetated flow can be expressed as fv = 4Cd(Uv/Ub)2 where Uv is the spatially averaged velocity within the canopy volume, and Cd is a local drag coefficient per unit frontal area derived to include the aforemontioned layer-wise effects of vortical structures within and above the canopy along with key vegetation properties. The proposed expression is compared with a number of empirical relations derived for vegetation under emergent and submerged conditions as well as numerous data sets covering a wide range of canopy morphology, densities, and rigidity. It is envisaged that the proposed formulation be imminently employed in eco-hydraulics where the interaction between flow and vegetation is being sought.
Similar content being viewed by others
Introduction
Since its inception by Darcy (in 1857) and Weisbach (in 1845), the now called Darcy-Weisbach equation for determining frictional losses in open channels (and pipes) is considered ‘standard’ provided its associated friction factor coefficient \(f=8{({u}_{\ast }/{U}_{b})}^{2}\) is known, where \({u}_{\ast }=\sqrt{g{R}_{h}{S}_{f}}\) is the friction velocity related to the kinematic ‘bed stress’, g is the gravitational acceleration, Rh is the hydraulic radius, Sf is the frictional slope or energy grade-line that converges to the bed-slope So for uniform flow, \({U}_{b}=Q/A\) is the bulk or time and area-averaged velocity determined from the flow rate Q and cross-sectional area \(A=B{h}_{w}\) (assuming rectangular section) orthogonal to the flow direction, B is the channel width and hw is the water depth approximating Rh when \(B\gg {h}_{w}\). The definition of f can be combined with the estimate of u* to yield
In classical hydraulics, the work of Moody, Nikuradse and many others established f to vary with two dimensionless quantities: the relative roughness r/Rh and a bulk Reynolds number \({R}{{e}}_{b,h}={U}_{b}{R}_{h}/\nu \), where ν is the kinematic viscosity of water. This expression is generally accepted in pipe- and open channel- flows above small-scale roughness elements where \(r/{h}_{w}\ll 1\). At very high Reb,h, f becomes independent of Reb,h and is presumed to abide by the so-called Strickler scaling \(f\sim {(r/{R}_{h})}^{1/3}\)1,2,3. An immediate consequence of this result is that when r is a priori known, f and subsequently Sf can be determined from Eq. 1.
Operationally, such an expression for Sf can be used to mathematically close the combined continuity and unsteady shallow water flow equations (i.e., the Saint-Venant) to predict hw and Ub in a plethora of hydrological and hydraulics applications. Example applications include overland flow from bare to vegetated patches, dam-break and similar shallow inertial flows, flash-flood runoff in ephemeral streams, tsunamis on coastal plains, to name a few4,5,6,7,8,9,10,11. It is now recognized that such naive view of f cannot be juxtaposition to streams covered by large roughness values (\(r/{h}_{w} > 0.1\)) such as shallow flow over gravel beds12,13,14,15,16,17 or vegetation18,19,20 and frames the scope of the work here. In fact, a number of authors are calling for the abandonment of such an approach altogether in such situations21,22.
The main theoretical novelty is to arrive at a new expression for f whose generic form resembles the conventional \(f=8{({u}_{\ast }/{U}_{b})}^{2}\) for steady-uniform flow through or above dense canopies. Specifically, it is shown that under certain simplifying assumptions, the canopy-related f (hereafter referred to as fv) is given by
where Uv is the spatially averaged velocity within the canopy volume, and Cd is a local drag coefficient per unit frontal area derived to include the layer-wise effects of vortical structures within and above the canopy along with key vegetation properties. For this expression to be used in practice, an estimate of Uv/Ub and Cd are required. The derivation of Eq. 2 and estimates of Cd and Uv/Ub are first presented followed by a comparison between model predictions and measurements of fv. Comparisons with other widely used models of fv are also featured demonstrating that the proposed formulation appears superior to all prior formulations across a wide-range of experiments and canopy-flow configurations.
Background
To progress on the description of fv for canopy flows, a number of studies have been conducted to explore connections between the shape of the mean velocity profile u(z) and its depth-integrated value defined as
in wide rectangular channels, where z is the distance from the channel bottom23,24,25,26,27,28,29,30. These studies naturally bridge the dominant vortical structures31,32 to u(z)12,33,34,35 and subsequently to Ub and fv. Other studies have focused on the force balance and stresses acting on the vegetation elements as well as Rh to arrive at u* and fv36,37,38. Various resistance models for flow within and above vegetated elements were proposed that used bulk flow measurements or combination of models and measurements to determine f 9,39,40,41,42,43,44,45,46,47,48,49,50,51,52,53,54,55,56. The emerging picture from all these studies is that key vortical structures impact the shape of u(z) across various canopy layers and in the vegetation free layer for submerged vegetation. The relation between these vortical structures and u(z) is now reviewed.
Review of the mean velocity profile within and above canopies
For stationary and planar-homogeneous flow in a wide channel covered by a densely vegetated canopy, the flow region hw can be decomposed into three zones according to the dominant sizes of the vortices in each zone as shown in Fig. 1.
Zone I forms near the channel bottom where von Karman vortex streets dominate the energetics of turbulence in the vertical direction. The mean velocity in zone I can be approximated by a constant given as41,57,58
where Cd is a local drag coefficient as before, m is the number of rods or vegetation stems per unit ground area, and D is the frontal width of vegetation elements.
Zone II spans the vegetation top and is dominated by attached eddies to a zero-plane displacement as well as mixing-layer eddies (though those types of eddies do not co-exist in space but both impact time-averaged statistics). When explained by the mixing layer analogy, the mean velocity is given by
where uv,top is the mean velocity at the top of the vegetation elements, hv is the vegetation height, Lmix is a characteristic energetic eddy size generated from Kelvin-Helmholtz instabilities34.
Zone III resembles a canonical turbulent boundary layer above a rough ground surface and is commonly represented by
where the friction velocity \({u}_{\ast ,sur}\approx \sqrt{g{h}_{s}{S}_{f}}\) for the surface layer (as before) whose thickness is hs, zo is the momentum roughness length (linked to r or hv), d is the zero-plane displacement height and is related to the penetration depth, κ is the von Karman constant, and \({h}_{s}={h}_{w}-{h}_{v}\) when \({h}_{w}/{h}_{v} > 1\) (submerged vegetation case).
Due to the presence of vegetation, the addition of zones I and II introduce major distortions to the expected u(z) shape when compared to a typical rough boundary (i.e. zone III) where fv primarily varies with r/hw and Reb,h as before. When \(r/{h}_{w}\ll 1\) (in zone III), r may be linked to z0 using the scaling relation \(r\sim {z}_{o}^{1/6}\) discussed elsewhere12. The aforementioned distortions to u(z) and the possibility of deriving a general expression for fv to be used in operational models for emergent and submerged vegetation is the main goal of the work here.
Derivation of a Friction Factor for Vegetated Flow
The goal of the derivation here is to arrive at a new formula for fv whose generic form resembles the conventional \(f=8{({u}_{\ast }/{U}_{b})}^{2}\) and is given by
This definition necessitates estimates of the friction velocity for the vegetated flow \({u}_{\ast v}=\sqrt{g{S}_{v}{R}_{v}}\), where Sv denotes the energy grade-line slope caused by the presence of vegetation, Rv denotes a new vegetation-related hydraulic radius, and Ub is, as before, the bulk flow velocity averaged over the entire cross-sectional area but adjusted for the finite porosity of the vegetation medium.
Section 2.1 derives the vegetation-related energy slope Sv and Section 2.2 derives the vegetation-related hydraulic radius Rv. The outcome is then compared to a wide range of published data sets and prior formulations derived from fitting to other published data sets.
Derivation of vegetation-related energy slope
Throughout, it is assumed that the flow is steady and uniform occurring within a wide rectangular open channel characterized by width B and bed slope So. The case considered here is when vegetation is sufficiently dense so that the overall friction factor is mainly due to fv not bed and side-wall friction and the frictional slope is Sv. The vegetation is further assumed to be cylindrical with height hv, diameter D, and density m defined by the number of cylinders per unit area as before. The vegetation can be in one of two-states: submerged (\(\alpha < 1\)) or emergent (\(\alpha > 1\)) as characterized by the degree of submergence \(\alpha ={h}_{v}/{h}_{w}\). For a control volume above a unit bed area extending from the channel bed to the water surface, the mean momentum balance in the streamwise direction reduces to a force-balance between (i) the gravitational contribution of the water weight along the streamwise direction for a unit ground area (=Fw), and (ii) the drag force per unit ground area (=Fv) caused by the presence of vegetation stems resisting the flow. The Fw is given by
where ρ is the water density, Vw is the volume of flowing water per unit ground area determined from the entire domain volume reduced by the volume occupied by the vegetation elements. The \({V}_{w}={h}_{w}\,(1-\alpha \varphi )\) for submerged conditions (i.e. \(\alpha < 1\)) and \({V}_{w}={h}_{w}\,(1-\varphi )\) for emergent vegetation (i.e. \(\alpha > 1\)), where \(\varphi \) is the volume fraction of the vegetation. When ignoring wall stresses relative to the drag force imposed by the dense vegetation and assuming a quadratic-law for Fv yields
where Av is frontal area of vegetation given by \({A}_{v}=mD{h}_{v}\) for \(\alpha < 1\) and \({A}_{v}=mD{h}_{w}\) for \(\alpha > 1\), Cd in Eq. (9) is a ‘local spatially-averaged’ non-constant drag coefficient that must account for the effects of the vegetation on the flow as discussed elsewhere55, and Uref is a reference velocity representing the flow within the vegetation section sensing the drag elements. A number of plausible velocities are now introduced to represent Uref: (a) The bulk velocity in the vegetation layer (VL, zones I + II) sensing the drag effects that can be estimated from \({U}_{v,b}={Q}_{v}/(B{h}_{v})\)59,60,61. (b) The pore velocity in VL calculated based on a spatially-averaged value \({U}_{v,p}={Q}_{v}/({B}_{p}{h}_{v})\), where the effective width used for the pore velocity is \({B}_{p}=B(1-\varphi )\)36,51,62,63,64,65. (c) The constricted velocity in a classical staggered array \({U}_{v,c}={Q}_{v}/({B}_{c}{h}_{v})\) with a characteristic constricted width \({B}_{c}=B(1-D/{L}_{s,stag})\)55, where Ls,stag is the spacing distance defined in Etminan et al.’ work55. (d) The separation velocity \({U}_{v,s}={k}_{p}{U}_{v,p}\) where kp is a kinetic energy of a subgrid scale (SGS) for a Smagorinsky model55. Due to different vegetation arrangements used in the experiments analyzed here (Section 3 and 4), the spatially-averaged pore velocity was adopted (i.e. \({U}_{ref}={U}_{v,p}\)). Detailed analysis shows that this choice of Uref along with a vegetation-related hydraulic radius resulting in a new Reynolds number perform reasonably when compared to many experiments (discussed later). Hence, from the force balance \({F}_{w}={F}_{v}\), the vegetation-related friction slope is given by
where Cd and Uv,p remain to be determined.
Derivation of a vegetation-related hydraulic radius R v
Many studies treat flow resistance of a vegetated canopy as equivalent to a ‘bed stress’ assuming energy losses occur due to a ‘vegetated bed’. This equivalence is convenient as it leads to a hydraulic radius related to flow depth hw or surface layer depth hs9,41,44,66 without considering the details of the vegetation configuration such as frontal area associated with \(\varphi \), m and D. However, the flow resistance by vegetation is mainly dominated by form drag and is directly linked to the frontal area of the rods. The hydraulic radius may be revised to account for these vegetation-related features. Here, a vegetation-related hydraulic radius Rv is proposed by extending the work of Cheng36 from emergent condition to submerged cases, where the vegetation-related hydraulic radius Rv is now defined by the ratio of the whole flow volume Vw per unit ground area to the vegetation-fluid contact Aresistance per unit ground area and can be expressed as
From Eq. (9), the vegetation resistance is caused by its form drag, which is a function of the frontal vegetated area Av per unit ground area. Hence, the vegetation-related hydraulic radius is
Since both Vw and Av are defined per unit ground area, the ground area cancels in the definition of Rv.
Derivation of the friction formula for flow through submerged and emergent vegetation
The friction factor caused by the presence of vegetation is now obtained by substituting Eqs (10 and 12) into Eq. (7) to yield
This expression is similar to a prior result given by9
derived from a different set of assumptions. To illustrate the similarities between the two expressions, consider the canopy-level adjustment length scale \({L}_{c}={({C}_{d}mD)}^{-1}\) as defined by the aforementioned study and \({\rm{\Delta }}U={U}_{s}-{U}_{v}\) to be the velocity difference between the vegetation zone Uv and the free water zone Us. Equation (14) can be arranged to read \({f}_{v,poggi}=4{C}_{d}mD{h}_{v}\,{({U}_{v}/{U}_{b})}^{2}\), where Cd is a local drag coefficient per unit frontal area (as shown next). More relevant here is that Eq. (13) allows for multiple effects to be conveniently included in a dimensionless Cd (instead of a dimensional Lc) and Uv/Ub. Before doing so, it is to be noted that when submergence \(\alpha ={h}_{v}/{h}_{w}\ll 1\)44, \({U}_{v}\sim {u}_{\ast }\) and the conventional expression for friction factor is recovered from Eq. (13). To show explicitly similarities and differences between Eqs (13 and 14), the force balance for submerged conditions is rewritten per unit ground area as
so that
where Cd is a local drag coefficient per unit frontal area as before. Equation (14) further assumes the hydraulic radius to be the flow depth, or \({R}_{h}={h}_{w}(1-\alpha \varphi )\approx {h}_{w}\). Using this definition of \({\tau }_{v,poggi}={C}_{d}mD{h}_{v}\rho {U}_{v}^{2}/2\) and inserting this definition into the definition of \(f=8\tau /(\rho {U}_{b}^{2})\) yields \({f}_{v,poggi}=4{C}_{d}mD{h}_{v}\,{({U}_{v}/{U}_{b})}^{2}\). However, when using a vegetation-related hydraulic radius \({R}_{v}={V}_{w}/{A}_{v}={h}_{w}(1-\alpha \varphi )/(mD{h}_{v})\) instead of Rh as proposed here, Eq. (16) can be transformed as
and the expression
is recovered where Cd remains a local drag coefficient per unit frontal area (as before). Naturally, when setting \({\tau }_{v,present}={C}_{d}\rho {U}_{v}^{2}/2\) into the definition of \(f=8\tau /(\rho {U}_{b}^{2})\), the outcome \({f}_{v,present}=4{C}_{d}\,{({U}_{v}/{U}_{b})}^{2}\) of Eq. (13) is recovered. It can be surmised that different Rh definitions yield expressions for f that appear to be different. However, not withstanding these apparent differences, the Cd remains a local drag coefficient per unit frontal area in all of them.
Because Cd is to be related to a local Reynolds number, several possibilities are now reviewed for defining a local Reynolds number in the presence of vegetation. Etminan et al.55 compared Reynolds numbers using various characteristic velocity scales (Uv,b, Uv,p, Uv,c and Uv,s) with a fixed characteristic length scale D giving Rev,b, Rev,p, Rev,c and Rev,s. The aforementioned work then showed that typical Cd formulation for a single cylinder can be employed to predict a drag coefficient for staggered vegetation when using Uv,c as the reference velocity to form \({R}{{e}}_{v,c}={U}_{v,c}D/\nu \) and resulting in a best-fit expression applicable for \(0 < {R}{{e}}_{v,c} < 6000\) given as
where suffix ‘EGL’ denotes the first letter of the surname of each author in Etminan et al.55. Here, vegetation effects are introduced in both velocity and length scales by selecting the pore velocity (Uv or Ub) and the vegetation-related hydraulic radius Rv for defining a Reynolds number. For emergent canopies, only the canopy layer exists and the Reynolds number is labeled as Rev,v (in the suffix, the first letter ‘v’ is vegetation layer, and second letter ‘v’ denotes the characteristic length in the Reynolds number to be the vegetation-related hydraulic radius) given as
where Rv is the vegetation-related hydraulic radius as before, given as
For a submerged vegetation, the Reynolds number Reb,v can also be defined as (in the suffix the first letter ‘b’ is bulk flow, and second letter ‘v’ denotes the characteristic length in the Reynolds number to be vegetation-related hydraulic radius)
where Rv is, as before, the vegetation-related hydraulic radius now given as
Friction Factor for Flow Through Emergent Vegetation
For emergent vegetation in flow (hv > hw)
Cheng’s result36 for emergent vegetated flow is recovered. In the aforementioned study, f was shown to be a monotonically decreasing function of Reynolds number Rev,v consistent with several studies29,39,59,62,63,67,68. The relation between f and Reynolds number is analyzed using several experiments described next.
Laboratory experiments
Present study
Flume experiments were conducted in a 10 m long and B = 0.3 m wide glass flume at the State Key Laboratory of Water Resources and Hydro-power Engineering Science at Wuhan University in China. The flume bed was set flat and the vegetation stems were arranged in linear configuration to ensure a locally uniform resistance. The vegetation was composed of cylinders with diameter D = 8 mm and height hv = 250 mm. The cylindrical vegetation array was then positioned on a 10 mm thick plastic board covered with holes to allow for variable cylinder density variation (or m). Let \({\varphi }_{board}={m}_{0}\pi {D}^{2}/4\) be the fractional area covered by holes on the bare board with m0 being the number of holes on the bare board per unit board area being used for each run. Eight vegetation densities were then used, labeled Runs A to H, with \(\varphi \) = 0.419, 0.291, 0.206, 0.163, 0.073, 0.041, 0.018 and 0.010, respectively. For all the runs, steady non-uniform flow was conducted with constant flow rate \(Q=0.00384\,\,{{\rm{m}}}^{3}\,{s}^{-1}\).
A local Cd,local – Re function in steady non-uniform flow was derived and shown to be parabolic in shape (i.e. not monotonic) within the vegetation zone for a given vegetation density. Here, the focus is on the averaged drag for the entire vegetation zone, which can be expressed as
where Cd,local is a local drag coefficient that varies in the streamwise direction along the vegetation zone, the normalized distance is \({x}^{+}=x/{L}_{veg}\) along the streamwise direction, and Lveg is the streamwise length of the vegetation zone. A summary of the data used here is given in Table 1 and the details are shown in Supplementary Information (Table S1). Also, further details about the experimental setup can be found elsewhere51.
Two other published data sets from Ishikawa et al.59 and Tanino et al.62 are used here and are briefly summarized in Table 1.
Experiments conducted by Ishikawa et al
The first data set was generated from a 15 m long, 0.3 m wide, and 0.3 m deep channel using two types of steel cylinders with diameters 0.004 m and 0.0064 m, and a height of 0.2 m as described elsewhere59. The cylinder spacing was uniform in each case (=0.0632 m and 0.0316 m, respectively). The flow was nearly uniform and the steel canopy array was arranged in staggered configuration. The hydraulic parameters required here are given in Supplementary Information (Table S1).
Experiments conducted by Tanino et al
Tanino et al.62 conducted their experiments in two Plexiglas recirculating flumes. Cylindrical maple dowels with diameter D = 0.0064 m were used as laboratory models for vegetation. The vegetation array fractional volume varied as \(\varphi \) = 0.091, 015, 0.20, 0.27, 0.35. The following data points (shown in Supplementary Information Table S1) were tracked from their best-fitting curve to yield an approximate expression
where suffix ‘tn’ denotes the expression of Tanino et al.62, T1 and T2 are obtained from linear regression. The Rev,d in their formulation was based on a stem Reynolds number given as
which uses D as characteristic length for their Reynolds number.
Empirical expressions
For flow through emergent vegetation, a large synthesis study proposed a Cd − Re expression given as36
The linkage between vegetation-array and stem related Reynolds number is
Here, a simplified expression that summarizes all the aforementioned expressions may be expressed as a function of Rev,v (desirable for the purposes of the work here) as
This summary expression is shown in Fig. 2, which an be re-arranged to be related to the more commonly reported Rev,d as
Based on this drag coefficient, the f for emergent vegetation is given as a function of Rev,v
or as a function of Rev,d
Friction Factor for Flow Through Submerged Vegetation
Laboratory experiments
For submerged vegetated flow, data were used from several published experiments and classified as either rigid or flexible. Data for the rigid vegetation flow are described elsewhere29,31,38,57,69,70,71,72,73,74,75,76. Data for flexible vegetation are also described elsewhere24,69,76,77,78,79,80,81. A summary of all the data points are in Table 2 and the details can be found in Supplementary Information (Tables S2 and S3).
Determination of depth-averaged velocity in the vegetation and surface layers
In the vegetation layer (VL), the depth-averaged velocity can be derived based on the force balance between vegetation drag and flow gravity in streamwise direction
where Cd is calculated from Eqs (30 or 31), indicating iterations are needed to obtain the velocity and drag coefficient. Moreover, the effective width of the canopy layer for the pore velocity \({B}_{p} < B\) is briefly discussed. Figure 3 shows side and top views of the flume with cylindrical vegetation, where the green zone indicates the area occupied by vegetation. The effective width Bp is expressed as \({B}_{p}=B\,(1-\varphi )\).
For surface layer (SL), the depth-averaged velocity can be determined by the linkage between the vegetation-layer velocity Uv and the bulk velocity Ub. The bulk velocity is defined by the total discharge Q to the effective cross-sectional area and given as (pore velocity for bulk flow)
From the continuity equation,
where Qs, As,flow are the flow rate and the effective cross-sectional area in the surface layer, and Qv, Av,flow are the flow rate and the effective cross-sectional area in the vegetation layer. The relation between Ub, Us and Uv is derived as
Arranging the above equation, the bulk velocity is given by
then the depth-averaged velocity of surface layer gives
Methods to determine friction factor in the submerged cases
The measured friction fv,measure can be obtained from Uv (using Eq. 34), Ub (using Eq. 35) and Cd (using Eqs 30 or 31) when setting \({f}_{v}=4{C}_{d}\,{({U}_{v}/{U}_{b})}^{2}\). All data points and their variation with the Reynolds number Reb,v or Reb,d are given in Fig. 4. The results suggest no obvious trends between measured fv and estimated Reb,v or Reb,d.
The absence of a unique relation between fv and Reynolds number requires further inquiry. The key to calculating fv is Uv/Ub. Previous studies9 showed that fv is linked to two dimensionless groups: \(\alpha ={h}_{v}/{h}_{w}\) and \(\eta ={h}_{v}/{L}_{c}\). As mentioned before, the adjustment canopy length scale is defined as
The aforementioned study9 did show a nonlinear increase in fv with increasing α at a given \(\eta \), and a nonlinear increase in fv with increasing \(\eta \) at a given α. Based on these results, we propose a 3-parameter mathematical function to describe this behavior without focusing on the detailed mean velocity profile in each zone. This 3-parameter function is given as
where p1, p2, p3 are constants to be determined from regression analysis. The corresponding friction fv can now be expressed as
This equation shows fv increases with increasing submergence \(\alpha ={h}_{v}/{h}_{w}\) when other parameters are held constant, and increases with increasing \(\eta ={h}_{v}/{L}_{c}\) when other parameters are held constant. With a local Cd determined from Eq. (31), parameters \({p}_{1}=1.198\), \({p}_{2}=0.681\), \({p}_{3}=0.416\) where determined using non-linear regression with a correlation coefficient \({R}_{cc}=0.803\) as shown in Fig. 5.
Friction factor derived from previous studies
Some studies report expressions for the velocity within the vegetated zone, which can also be used to obtain a friction factor. A summary of these expressions are briefly reviewed. Stone et al.74 estimated the bulk velocity as
Baptist et al.41 report a bulk velocity as
where Cd is the bed-related Chezy coefficient (=60 m1/2 s−1 for smooth bed).
Huthoff et al.28 proposed a
Yang et al.82 showed a bulk velocity given as
where \({C}_{u}=1\) for \(4\varphi /(\pi D)\le 5\) and \({C}_{u}=2\) for \(4\varphi /(\pi D) > 5\).
Cheng et al.83 derived the representative roughness height of the vegetation for the surface layer and proposed
Katul et al.44 proposed an analytical model for Ub linked to three canonical length scales (Lc, hv, hw) and is given as
where UKCL, UKSL are layer-averaged velocity for the canopy and surface layers, respectively, given as
and
where \(\beta =\,{\min }(0.135\sqrt{mD},0.33)\) is a momentum absorption coefficient, and z0 is given by
With these mean velocity profile formulations, friction factor can be estimated. The bulk velocity of the vegetation layer is determined using Eq. (34) and the velocity across hw is calculated based on Eqs (43 to 48).
Comparison between measured and modeled friction factor using published bulk velocities
The different methods are now used to predict fv and are then compared with measured fv in Fig. 6. To assess the agreement between prediction and measurement, Rcc is computed and reported in Table 3. Figure 6 suggests that the present model performs ‘no worse’ than other (more elaborate) methods. In many cases, it even outperforms prior methods.
Discussion and Conclusion
Equation \({f}_{v}=4{C}_{d}\,{({U}_{v}/{U}_{b})}^{2}\) encodes much of the recent developments about vegetation effects on local drag (e.g. blockage and sheltering, weak Reynolds number dependence, etc) and their up-scaled contribution on fv via Uv/Ub. These effects are now briefly discussed. For isolated cylinders, the local Cd can be determined from Cheng43
where
and
Many studies found that Cd in a vegetated array differs from Cd,iso. When only focusing on variations in vegetation density, Cd appears to increase62,68 and then decrease60,84 with increasing vegetation density55. In the present study, we go beyond vegetation density and introduce all the key vegetation effects in a Reynolds number \({R}{{e}}_{v,v}={U}_{v}{R}_{v}/\nu \) formed by Uv and Rv. These velocity and length scales contain the vegetation density \(\varphi \) and appear to collapse the experiments onto a single curve.
To accommodate the mechanics of sheltering, delayed separation and blockage effects arising from an array of cylinders instead of isolated cylinders, Cd of Eq. (31) can be used. Sheltering effect indicates that some elements were located in the wake region of the upstream elements85, resulting in a lower velocity than their upstream counterparts and generate a lower drag compared with the isolated cylinder case. Delayed separation can be explained by the enhancement of the mean separation angle that is larger than that for the isolated cylinder, resulting in a decreasing drag coefficient (compared with the isolated cylinder). Both sheltering and delayed separation tend to reduce drag when compared to the isolated cylinder case. For the blockage effect, which leads to a local increase in the drag coefficient, it can by explained by two main factors55, one is that the velocity between cylinders is enhanced by the presence of the vegetation in flow. The other factor is reduced wake pressure86.
The Cd,iso is a local drag coefficient acting on a single stem per unit frontal area without considering the interaction among elements in a vegetation array. To illustrate the role of sheltering, delayed separation and blockage on Cd, a ‘bifurcation’ type Reynolds number \({R}{{e}}_{v,d}^{+}\) may be introduced. When \({R}{{e}}_{v,d} < {R}{{e}}_{v,d}^{+}\), the viscous boundary layer formed around cylinders creates a slow-moving flow zone with a path smaller than the spacing between adjacent stem. This effect leads to an equivalent local drag coefficient by the array of cylinders to be larger than the one associated with isolated cylinders (i.e. a blockage effect). However, when \({R}{{e}}_{v,d} > {R}{{e}}_{v,d}^{+}\), vegetation stems become a new source of turbulent kinetic energy (wake production) spawning horizontal vortices resembling von Karman streets that grow in size and ‘fill’ the space between the stems. This effect is akin to a decreasing drag coefficient when compared with Cd,iso (i.e. sheltering effect). Blockage and sheltering effects can be distinguished by calculating
When \(E > 1\), blockage dominates and when \(E < 1\), sheltering dominates Cd51. Using Eqs (31 and 52) with different vegetation concentration (\(\varphi \) = 0.01, 0.05, 0.1, 0.3 and 0.5), blockage and sheltering can be delineated for Rev,d ranging from 101 to 105 in Fig. 7. As expected, Cd increases with increasing \(\varphi \) for small \(\varphi \). The threshold Reynolds number \({R}{{e}}_{v,d}^{+}\) also increases with increasing \(\varphi \). For example, at \({R}{{e}}_{v,d}^{+}\approx 4000\), separation between blockage and sheltering occurs for \(\varphi =0.01\). This threshold \({R}{{e}}_{v,d}^{+}\) becomes much larger, around 30000 for \(\varphi =0.5\). For \(\varphi =0.01\), when \({R}{{e}}_{v,d} < {R}{{e}}_{v,d}^{+}\), the Cd for isolated stems are comparable but not identical to their array counterpart.
To conclude, the expression of friction factor proposed here does accommodate blockage or sheltering through a local drag coefficient Cd (Eq. 31) and any distortions to the shape of the mean velocity profile (Eq. 41). Evaluated using numerous data sets covering a wide range of canopy morphology, densities and rigidity, the friction factor for emergent vegetation is shown to be proportional to the drag coefficient. This finding shows why the friction factor monotonically decreases with increasing Reynolds number for emergent vegetation. For submerged vegetation, the friction factor is shown to vary with submergence and adjustment length scale. Also, this variation is not monotonic with increasing Reynolds number.
References
Strickler, A. Beitrge zur frage der geschwindigkeitsformel und der rauhigkeitszahlen fr strme, kanle und geschlossene leitungen. Mitteilungen des Amtes fr Wasserwirtschaft (1923).
Gioia, G. & Bombardelli, F. Scaling and similarity in rough channel flows. Physical review letters 88, 014501 (2002).
Bonetti, S., Manoli, G., Manes, C., Porporato, A. & Katul, G. G. Mannings formula and stricklers scaling explained by a co-spectral budget model. Journal of Fluid Mechanics 812, 1189–1212 (2017).
Ajayi, A. E., van de Giesen, N. & Vlek, P. A numerical model for simulating hortonian overland flow on tropical hillslopes with vegetation elements. Hydrological Processes 22, 1107–1118, https://doi.org/10.1002/hyp.6665 (2008).
Chanson, H., Jarny, S. & Coussot, P. Dam break wave of thixotropic fluid. Journal of Hydraulic Engineering 132, 280–293 (2006).
Chanson, H. Application of the method of characteristics to the dam break wave problem. Journal of Hydraulic Research 47, 41–49 (2009).
Daly, E. & Porporato, A. Similarity solutions of nonlinear diffusion problems related to mathematical hydraulics and the fokker-planck equation. Physical Review E 70, 056303 (2004).
Hogg, A. J. & Pritchard, D. The effects of hydraulic resistance on dam-break and other shallow inertial flows. Journal of Fluid Mechanics 501, 179–212 (2004).
Poggi, D., Krug, C. & Katul, G. G. Hydraulic resistance of submerged rigid vegetation derived from first-order closure models. Water Resources Research 45, https://doi.org/10.1029/2008WR007373 (2009).
Thompson, S., Katul, G., Konings, A. & Ridolfi, L. Unsteady overland flow on flat surfaces induced by spatial permeability contrasts. Advances in Water Resources 34, 1049–1058, https://doi.org/10.1016/j.advwatres.2011.05.012 (2011).
Paschalis, A., Katul, G. G., Fatichi, S., Manoli, G. & Molnar, P. Matching ecohydrological processes and scales of banded vegetation patterns in semi-arid catchments. Water Resources Research 1–68, https://doi.org/10.1002/2015WR017679 (2016).
Katul, G. G., Wiberg, P., Albertson, J. & Hornberger, G. A mixing layer theory for flow resistance in shallow streams. Water Resources Research 38, https://doi.org/10.1029/2001WR000817 (2002).
Manes, C., Pokrajac, D. & McEwan, I. Double-averaged open-channel flows with small relative submergence. Journal of Hydraulic Engineering 133, 896–904 (2007).
Manes, C., Poggi, D. & Ridolfi, L. Turbulent boundary layers over permeable walls: scaling and near-wall structure. Journal of Fluid Mechanics 687, 141–170 (2011).
Manes, C., Ridolfi, L. & Katul, G. A phenomenological model to describe turbulent friction in permeable-wall flows. Geophysical research letters 39 (2012).
Lamb, M. P., Brun, F. & Fuller, B. M. Hydrodynamics of steep streams with planar coarse-grained beds: Turbulence, flow resistance, and implications for sediment transport. Water Resources Research 53, 2240–2263 (2017).
Ali, S. Z. & Dey, S. Impact of phenomenological theory of turbulence on pragmatic approach to fluvial hydraulics. Physics of Fluids 30, 045105 (2018).
Nepf, H. M. Flow and transport in regions with aquatic vegetation. Annual Review of Fluid Mechanics 44, 123–142, https://doi.org/10.1146/annurev-uid-120710-101048 (2012).
Wang, W.-J. et al. Roughness height of submerged vegetation in flow based on spatial structure. Journal of Hydrodynamics 30, 754–757, https://doi.org/10.1007/s42241-018-0060-3 (2018).
Wang, W.-J. et al. Derivation of canopy resistance in turbulent flow from first-order closure models. Water 10, 1782, https://doi.org/10.3390/w10121782 (2018).
Bjerklie, D. M., Dingman, S. L. & Bolster, C. H. Comparison of constitutive flow resistance equations based on the manning and chezy equations applied to natural rivers. Water resources research 41 (2005).
Ferguson, R. Time to abandon the manning equation? Earth Surface Processes and Landforms 35, 1873–1876 (2010).
Cheng, N.-S. Single-layer model for average flow velocity with submerged rigid cylinders. Journal of Hydraulic Engineering 141, 06015012, https://doi.org/10.1061/(ASCE)HY.1943-7900.0001037 (2015).
Ciraolo, G. & Ferreri, G. Log velocity profile and bottom displacement for a flow over a very flexible submerged canopy. In Proceedings of 32nd Congress of IAHR Harmonizing the Demands of Art and Nature in Hydraulics, 1–13 (International Association for Hydro-Environment Engineering and Research, Venice, Italy, 2007).
Dijkstra, J. & Uittenbogaard, R. Modeling the interaction between flow and highly flexible aquatic vegetation. Water Resources Research 46, https://doi.org/10.1029/2010WR009246 (2010).
Wang, W.-J., Huai, W.-X., Zeng, Y.-H. & Zhou, J.-F. Analytical solution of velocity distribution for flow through submerged large deflection flexible vegetation. Applied Mathematics and Mechanics 36, 107–120, https://doi.org/10.1007/s10483-015-1897-9 (2015).
Huai, W.-X., Wang, W.-J. & Zeng, Y.-H. Two-layer model for open channel flow with submerged flexible vegetation. Journal of Hydraulic Research 51, 708–718, https://doi.org/10.1080/00221686.2013.818585 (2013).
Huthoff, F., Augustijn, D. & Hulscher, S. J. Analytical solution of the depth-averaged flow velocity in case of submerged rigid cylindrical vegetation. Water Resources Research 43, https://doi.org/10.1029/2006WR005625 (2007).
Liu, D., Diplas, P., Fairbanks, J. D. & Hodges, C. C. An experimental study of flow through rigid vegetation. Journal of Geophysical Research: Earth Surface 113, https://doi.org/10.1029/2008JF001042 (2008).
Liu, Z. W., Chen, Y. C., Zhu, D. J., Hui, E. Q. & Jiang, C. B. Analytical model for vertical velocity profiles in flows with submerged shrub-like vegetation. Environmental Fluid Mechanics 12, 341–346, https://doi.org/10.1007/s10652-012-9243-6 (2012).
Ghisalberti, M. & Nepf, H. M. The limited growth of vegetated shear layers. Water Resources Research 40, https://doi.org/10.1029/2003WR002776 (2004).
Brunet, Y. & Irvine, M. The control of coherent eddies in vegetation canopies: streamwise structure spacing, canopy shear scale and atmospheric stability. Boundary-Layer Meteorology 94, 139–163, https://doi.org/10.1023/A:1002406616227 (2000).
Rowinski, P. & Kubrak, J. A mixing-length model for predicting vertical velocity distribution in flows through emergent vegetation. Hydrological sciences journal 47, 893–904, https://doi.org/10.1080/02626660209492998 (2002).
Raupach, M., Finnigan, J. & Brunei, Y. Coherent eddies and turbulence in vegetation canopies: the mixing-layer analogy. Boundary-Layer Meteorology 78, 351–382, https://doi.org/10.1007/BF00120941 (1996).
Ghisalberti, M. & Nepf, H. M. Mixing layers and coherent structures in vegetated aquatic flows. Journal of Geophysical Research 107, 3011, https://doi.org/10.1029/2001JC000871 (2002).
Cheng, N.-S. & Nguyen, H. T. Hydraulic radius for evaluating resistance induced by simulated emergent vegetation in open-channel flows. Journal of Hydraulic Engineering 137, 995–1004, https://doi.org/10.1061/(ASCE)HY.1943-7900.0000377 (2010).
Nepf, H. M., Ghisalberti, M., White, B. & Murphy, E. Retention time and dispersion associated with submerged aquatic canopies. Water Resources Research 43, https://doi.org/10.1029/2006WR005362 (2007).
Nezu, I. & Sanjou, M. Turburence structure and coherent motion in vegetated canopy open-channel flows. Journal of Hydro-Environment Research 2, 62–90, https://doi.org/10.1016/j.jher.2008.05.003 (2008).
James, C., Birkhead, A., Jordanova, A. & O’sullivan, J. Flow resistance of emergent vegetation. Journal of Hydraulic Research 42, 390–398, https://doi.org/10.1080/00221686.2004.9728404 (2004).
Järvelä, J. Flow resistance of flexible and stiff vegetation: a flume study with natural plants. Journal of Hydrology 269, 44–54, https://doi.org/10.1016/S0022-1694(02)00193-2 (2002).
Baptist, M. et al. On inducing equations for vegetation resistance. Journal of Hydraulic Research 45, 435–450, https://doi.org/10.1080/00221686.2007.9521778 (2007).
Carollo, F., Ferro, V. & Termini, D. Flow resistance law in channels with flexible submerged vegetation. Journal of Hydraulic Engineering 131, 554–564, https://doi.org/10.1061/(ASCE)0733-9429(2005)131:7(554) (2005).
Cheng, N.-S. Calculation of drag coefficient for arrays of emergent circular cylinders with pseudofluid model. Journal of Hydraulic Engineering 139, 602–611, https://doi.org/10.1061/(ASCE)HY.1943-7900.0000722 (2012).
Katul, G. G., Poggi, D. & Ridolfi, L. A flow resistance model for assessing the impact of vegetation on flood routing mechanics. Water Resources Research 47, https://doi.org/10.1029/2010WR010278 (2011).
Klopstra, D., Barneveld, H., Van Noortwijk, J. & Van Velzen, E. Analytical model for hydraulic roughness of submerged vegetation. In 27th IAHR Congress, 775–780 (HKV Consultants, 1996).
Konings, A. G., Katul, G. G. & Thompson, S. E. A phenomenological model for the flow resistance over submerged vegetation. Water Resources Research 48, https://doi.org/10.1029/2011WR011000 (2012).
Nepf, H. M. & Ghisalberti, M. Flow and transport in channels with submerged vegetation. Acta Geophysica 56, 753–777, https://doi.org/10.2478/s11600-008-0017-y (2008).
Schoneboom, T., Aberle, J. & Dittrich, A. Hydraulic resistance of vegetated flows: Contribution of bed shear stress and vegetative drag to total hydraulic resistance. In Proceedings of the international conference on fluvial hydraulics river flow (2010).
Van Velzen, E., Jesse, P., Cornelissen, P. & Coops, H. Hydraulic resistance of vegetation in floodplains, part 2: Background document version 1–2003. Ministry of Transport, Public Works and Water Management, Institute for Inland Water Management and Waste Water Treatment, report (2003).
Chapman, J. A., Wilson, B. N. & Gulliver, J. S. Drag force parameters of rigid and flexible vegetal elements. Water Resources Research 51, 3292–3302, https://doi.org/10.1002/2014WR015436 (2015).
Wang, W.-J., Huai, W.-X., Thompson, S. & Katul, G. G. Steady nonuniform shallow flow within emergent vegetation. Water Resources Research 51, 10047–10064, https://doi.org/10.1002/2015WR017658 (2015).
Boothroyd, R. J., Hardy, R. J., Warburton, J. & Marjoribanks, T. I. Modeling complex flow structures and drag around a submerged plant of varied posture. Water Resources Research 53, 2877–2901, https://doi.org/10.1002/2016WR020186 (2017).
Antonarakis, A. S., Richards, K. S., Brasington, J. & Muller, E. Determining leaf area index and leafy tree roughness using terrestrial laser scanning. Water Resources Research 46, W06510, https://doi.org/10.1029/2009WR008318 (2010).
Tinoco, R. O., Goldstein, E. B. & Coco, G. A data-driven approach to develop physically sound predictors: Application to depth-averaged velocities on flows through submerged arrays of rigid cylinders. Water Resources Research 51, 1247–1263, https://doi.org/10.1002/2014WR016380 (2015).
Etminan, V., Lowe, R. J. & Ghisalberti, M. A new model for predicting the drag exerted by vegetation canopies. Water Resources Research 53, 3179–3196, https://doi.org/10.1002/2016WR020090 (2017).
Kim, S. J. & Stoesser, T. Closure modeling and direct simulation of vegetation drag in flow through emergent vegetation. Water Resources Research 47, https://doi.org/10.1029/2011WR010561 (2011).
Poggi, D., Porporato, A., Ridolfi, L., Albertson, J. & Katul, G. The effect of vegetation density on canopy sub-layer turbulence. Boundary-Layer Meteorology 111, 565–587, https://doi.org/10.1023/B:BOUN.0000016576.05621.73 (2004).
Huai, W.-X., Wang, W.-J., Hu, Y., Zeng, Y.-H. & Yang, Z.-H. Analytical model of the mean velocity distribution in an open channel with double-layered rigid vegetation. Advances in Water Resources 69, 106–113, https://doi.org/10.1016/j.advwatres.2014.04.001 (2014).
Ishikawa, Y., Mizuhara, K. & Ashida, S. Effect of density of trees on drag exerted on trees in river channels. Journal of Forest Research 5, 271–279, https://doi.org/10.1007/BF02767121 (2000).
Lee, J., Roig, L., Jenter, H. & Visser, H. Drag coefficients for modeling flow through emergent vegetation in the florida everglades. Ecological Engineering 22, 237–248, https://doi.org/10.1016/j.ecoleng.2004.05.001 (2004).
Wu, F., Shen, H. & Chou, Y. Variation of roughness coefficients for unsubmerged and submerged vegetation. Journal of hydraulic Engineering 125, 934–942, https://doi.org/10.1061/(ASCE)0733-9429(1999)125:9(934) (1999).
Tanino, Y. & Nepf, H. M. Laboratory investigation of mean drag in a random array of rigid, emergent cylinders. Journal of Hydraulic Engineering 134, 34–41, https://doi.org/10.1061/(ASCE)0733-9429(2008)134:1(34) (2008).
Kothyari, U. C., Hayashi, K. & Hashimoto, H. Drag coefficient of unsubmerged rigid vegetation stems in open channel flows. Journal of Hydraulic Research 47, 691–699, https://doi.org/10.3826/jhr.2009.3283 (2009).
Zhao, K., Cheng, N., Wang, X. & Tan, S. Measurements of fluctuation in drag acting on rigid cylinder array in open channel flow. Journal of Hydraulic Engineering 140, 48–55, https://doi.org/10.1061/(ASCE)HY.1943-7900.0000811 (2013).
Wang, W.-J., Huai, W.-X., Thompson, S., Peng, W.-Q. & Katul, G. G. Drag coefficient estimation using flume experiments in shallow non-uniform water flow within emergent vegetation during rainfall. Ecological Indicators 92, 367–378, https://doi.org/10.1016/j.ecolind.2017.06.041 (2018).
Nepf, H. M. & Vivoni, E. R. Flow structure in depth-limited, vegetated flow. Journal of Geophysical Research: Oceans 105, 28547–28557, https://doi.org/10.1029/2000JC900145 (2000).
Ferreira, R. M., Ricardo, A. M. & Franca, M. J. Discussion of “laboratory investigation of mean drag in a random array of rigid, emergent cylinders” by yukie tanino and heidi m. nepf. Journal of Hydraulic Engineering 135, 690–693, https://doi.org/10.1061/(ASCE)HY.1943-7900.0000021 (2009).
Stoesser, T., Kim, S. & Diplas, P. Turbulent flow through idealized emergent vegetation. Journal of Hydraulic Engineering 136, 1003–1017, https://doi.org/10.1061/(ASCE)HY.1943-7900.0000153 (2010).
Dunn, C. Experimental determination of drag coefficients in open channel with simulated vegetation. Master’s thesis (1996).
López, F. & García, M. Mean flow and turbulence structure of open-channel flow through non-emergent vegetation. Journal of Hydraulic Engineering 127, 392–402, https://doi.org/10.1061/(ASCE)0733-9429(2001)127:5(392) (2001).
Meijer, D. Flumes studies of submerged vegetation. In PR121.10, HKV, Lelystad (in Dutch) (1998).
Murphy, E., Ghisalberti, M. & Nepf, H. M. Model and laboratory study of dispersion in flows with submerged vegetation. Water Resources Research 43, https://doi.org/10.1029/2006WR005229 (2007).
Shimizu, Y., Tsujimoto, T., Nakagawa, H. & Kitamura, T. Experimental study on flow over rigid vegetation simulated by cylinders with equi-spacing. In Proceedings of the Japan Society of Civil Engineers, vol. 438, 31–40 (1991).
Stone, B. M. & Shen, H. T. Hydraulic resistance of flow in channels with cylindrical roughness. Journal of Hydraulic Engineering 128, 500–506, https://doi.org/10.1061/(ASCE)0733-9429(2002)128:5(500) (2002).
Yan, J. Experimental study of flow resistance and turbulence characteristics of open channel flow with vegetation. Thesis (2008).
Yang, S.-Q. & Chow, A. T. Turbulence structures in non-uniform flows. Advances in Water Resources 31, 1344–1351, https://doi.org/10.1016/j.advwatres.2008.06.006 (2008).
Kubrak, E., Kubrak, J. & Rowiński, P. M. Vertical velocity distributions through and above submerged, flexible vegetation. Hydrological Sciences Journal 53, 905–920 (2008).
Okamoto, T. & Nezu, I. Flow resistance law in open-channel flows with rigid and flexible vegetation. In River Flow 2010, Dittrich, A., Koll, K. A., Aberle, J. & Geisenhainer, P. (eds), 261–268 (Proceedings of the International Conference on Fluvial Hydraulics, Braunschweig, Germany, 2010).
Järvelä, J. Effect of submerged flexible vegetation on flow structure and resistance. Journal of Hydrology 307, 233–241, https://doi.org/10.1016/j.jhydrol.2004.10.013 (2005).
Carollo, F., Ferro, V. & Termini, D. Flow velocity measurements in vegetated channels. Journal of Hydraulic Engineering 128, 664–673, https://doi.org/10.1061/(ASCE)0733-9429(2002)128:7(664) (2002).
Kouwen, N., Unny, T. & Hill, H. M. Flow retardance in vegetated channels. Journal of the Irrigation and Drainage Division 95, 329–344 (1969).
Yang, W. & Choi, S.-U. A two-layer approach for depth-limited open-channel flows with submerged vegetation. Journal of Hydraulic Research 48, 466–475, https://doi.org/10.1080/00221686.2010.491649 (2010).
Cheng, N.-S. Representative roughness height of submerged vegetation. Water Resources Research 47, https://doi.org/10.1029/2011WR010590 (2011).
Nepf, H. M. Drag, turbulence, and diffusion in flow through emergent vegetation. Water Resources Research 35, 479–489, https://doi.org/10.1029/1998WR900069 (1999).
Raupach, M. Drag and drag partition on rough surfaces. Boundary-Layer Meteorology 60, 375–395, https://doi.org/10.1007/BF00155203 (1992).
Zdravkovich, M. Flow around circular cylinders: Applications 2 (Oxford University Press, Oxford, England, 2000).
Acknowledgements
The authors acknowledge the support from the following projects: National Water Pollution Control and Treatment Science and Technology Major Project of China (2017ZX07101004-001); National Natural Science Foundation of China (51809286, 51439007, 11672213, 41671048, 51479219); China Postdoctoral Science Foundation (2017M610949, 2018T110122); U.S. National Science Foundation (NSF-EAR-1344703, NSF-AGS-1644382, NSF-DGE-1068871); U.S. Department of Energy (DE-SC-0011461); IWHR Research and Development Support Program (WE0145B062019,WE0145B532017). The data used in this paper are summarized in the Tables 1 and 2 and shown in Supplementary Information document in detail.
Author information
Authors and Affiliations
Contributions
Wang contributed to the main work of study design, experiments, data collection, analyses and manuscript writing. Peng, Huai, Katul and Liu provided comments on the study design, experiments, analyses and improved the manuscript. Qu and Dong contributed to the data collection and data analyses.
Corresponding author
Ethics declarations
Competing Interests
The authors declare no competing interests.
Additional information
Publisher’s note: Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Supplementary information
Rights and permissions
Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons license, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons license and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this license, visit http://creativecommons.org/licenses/by/4.0/.
About this article
Cite this article
Wang, WJ., Peng, WQ., Huai, WX. et al. Friction factor for turbulent open channel flow covered by vegetation. Sci Rep 9, 5178 (2019). https://doi.org/10.1038/s41598-019-41477-7
Received:
Accepted:
Published:
Version of record:
DOI: https://doi.org/10.1038/s41598-019-41477-7
This article is cited by
-
Impact of mixed-height vegetation patches on energy loss in open-channel flow
Scientific Reports (2025)
-
Flow structures in asymmetric compound channels with emergent vegetation on divergent floodplain
Acta Geophysica (2022)
-
Estimation of the longitudinal dispersion coefficient using a two-zone model in a channel partially covered with artificial emergent vegetation
Environmental Fluid Mechanics (2021)
-
Predictions of bulk velocity for open channel flow through submerged vegetation
Journal of Hydrodynamics (2020)









