Download as pdf or txt
Download as pdf or txt
You are on page 1of 19

Journal of Hydraulic Research

ISSN: 0022-1686 (Print) 1814-2079 (Online) Journal homepage: https://www.tandfonline.com/loi/tjhr20

Hydrodynamics of vegetated channels

Heidi M. Nepf

To cite this article: Heidi M. Nepf (2012) Hydrodynamics of vegetated channels, Journal of
Hydraulic Research, 50:3, 262-279, DOI: 10.1080/00221686.2012.696559

To link to this article: https://doi.org/10.1080/00221686.2012.696559

Published online: 25 Jun 2012.

Submit your article to this journal

Article views: 10512

View related articles

Citing articles: 144 View citing articles

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=tjhr20
Journal of Hydraulic Research Vol. 50, No. 3 (2012), pp. 262–279
http://dx.doi.org/10.1080/00221686.2012.696559
© 2012 International Association for Hydro-Environment Engineering and Research

JHR Vision paper

Hydrodynamics of vegetated channels


HEIDI M. NEPF, Professor, Department of Civil and Environmental Engineering, Massachusetts Institute of Technology, 77
Massachusetts Avenue, Building 48-216D, Cambridge, MA 02139, USA.
Email: hmnepf@mit.edu

ABSTRACT
This paper highlights some recent trends in vegetation hydrodynamics, focusing on conditions within channels and spanning spatial scales from
individual blades, to canopies or vegetation patches, to the channel reach. At the blade scale, the boundary layer formed on the plant surface plays a
role in controlling nutrient uptake. Flow resistance and light availability are also influenced by the reconfiguration of flexible blades. At the canopy
scale, there are two flow regimes. For sparse canopies, the flow resembles a rough boundary layer. For dense canopies, the flow resembles a mixing
layer. At the reach scale, flow resistance is more closely connected to the patch-scale vegetation distribution, described by the blockage factor, than
to the geometry of individual plants. The impact of vegetation distribution on sediment movement is discussed, with attention being paid to methods
for estimating bed stress within regions of vegetation. The key research challenges of the hydrodynamics of vegetated channels are highlighted.

Keywords: blade boundary layer; channel resistance; sediment transport; turbulence; vegetation

1 Introduction better understand and protect these systems, the study of vege-
tation hydrodynamics has, over time, become interwoven with
Aquatic vegetation provides a wide range of ecosystem ser- other disciplines, such as biology (e.g. Koch 2001), fluvial
vices. The uptake of nutrients and production of oxygen improve geomorphology (e.g. Tal and Paola 2007), landscape ecol-
water quality (e.g. Wilcock et al. 1999). The potential removal ogy (e.g. Larsen and Harvey 2011), and geochemistry (e.g.
of nitrogen and phosphorous is so high that some researchers Clarke 2002). This integration will surely accelerate in the
advocate widespread planting in waterways (Mars et al. 1999). future, as hydraulics contributes to understanding and managing
Seagrasses form the foundation of many food webs (Green and environmental systems (Nikora 2010).
Short 2003), and in channels, vegetation promotes biodiversity The presence of vegetation alters the velocity field across sev-
by creating different habitats with spatial heterogeneity in the eral scales, ranging from individual branches and blades on a
stream velocity (e.g. Kemp et al. 2000). Marshes and mangroves single plant to a community of plants in a meadow or patch.
reduce coastal erosion by damping waves and storm surge (e.g. Flow structure at the different scales is relevant to different pro-
Turker et al. 2006), and riparian vegetation enhances bank sta- cesses. For example, the uptake of nutrients by an individual
bility (Pollen and Simon 2005). Through the processes described blade depends on the boundary layer on that blade, that is, on
above, aquatic vegetation provides ecosystem services with an the blade-scale flow (e.g. Koch 1994). Similarly, the capture of
estimated annual value of over $10 trillion (Costanza et al. 1997). pollen is mediated by the flow structure generated around indi-
These services are all influenced in some way by the flow field vidual stigma (e.g. Ackerman 1997). In contrast, the retention or
existing within and around the vegetated region. release of organic matter, mineral sediments, seeds, and pollen
In rivers, aquatic vegetation was historically considered only from a meadow or patch depends on the flow structure at the
as a source of flow resistance, and vegetation was frequently meadow or patch scale (e.g. Gaylord et al. 2004, Zong and Nepf
removed to enhance flow conveyance and reduce flooding. 2010). Furthermore, spatial heterogeneity in the canopy-scale
Because of this context, the earliest studies of vegetation hydro- parameters can produce complex flow patterns. For example, in
dynamics focused on the characterization of flow resistance a marsh or wetland, a branching network of channels cuts through
with a strictly hydraulic perspective (e.g. Ree 1949). However, regions of dense, largely emergent vegetation. While the chan-
as noted above, vegetation also provides ecological services nels provide most of the flow conveyance, the vegetated regions
that make it an integral part of coastal and river systems. To provide most of the ecosystem function and particle trapping.

Revision received 21 May 2012/Open for discussion until 30 November 2012.

ISSN 0022-1686 print/ISSN 1814-2079 online


http://www.tandfonline.com
262
Journal of Hydraulic Research Vol. 50, No. 3 (2012) Vegetated channels 263

Thus, to describe the functions of a marsh, one must describe the Because of the difference in magnitude between molecular
transport into and circulation within the vegetated regions. These diffusivity (Dm ) and molecular viscosity (ν), the concentration
examples tell us that to properly describe the physical role of veg- boundary layer, δc , is smaller than δs . Specifically, δc =
etation within an environmental system, one must first identify δs S−1/3 , with the Schmidt number S = ν/Dm (e.g. Boudreau and
the spatial scale relevant to a particular process and choose mod- Jorgensen 2001). The kinematic viscosity of water is of the order
els and measurements that are consistent with that scale. This ν = 10−6 m2 s−1 , and for most dissolved species, Dm is of the
paper focuses particularly on vegetation in channels, considering order 10−9 m2 s−1 , so that in water we generally find δc = 0.1δs .
both submerged and emergent vegetation. The following sections Within δc , transport perpendicular to the surface can only occur
review some fundamental aspects of flow structure at the blade, through molecular diffusion, so that this layer is also called the
canopy (patch), and reach scales. diffusive sub-layer.
The mass flux to the blade surface, ṁ, is described by Fick’s
law,
2 Processes at the scale of individual blades
ṁ ∂C
= −Dm (1)
A ∂n⊥
2.1 Blade boundary layers and nutrient fluxes
(e.g. Kays and Crawford 1993). Here, A is the surface area and
At the scale of individual blades and leaves, the hydrodynamic ∂C/∂n⊥ is the gradient in concentration perpendicular to the
response is dominated by boundary-layer formation on the plant surface. If the flux across δc is the rate-limiting step in transferring
surface. A flat plate boundary layer has often been used as a model dissolved species to the blade, the concentration at the surface
for flow adjacent to leaves oriented in the streamwise (x) direc- is assumed to be zero, that is, the plant is a perfect absorber.
tion (Fig. 1). A viscous boundary layer forms at the leading edge In addition, because transport across the sub-layer proceeds at
(x = 0), and its thickness, δ, grows with the streamwise distance,
√ the rate of molecular diffusion, it is several orders of magnitude
δ(x) = 5 νx/U , with U being the mean velocity and ν the kine- slower than the turbulent diffusion occurring outside this layer.
matic viscosity (e.g. White 2008). As the viscous boundary layer Therefore, it is reasonable to assume that the concentration at
grows, it becomes sensitive to perturbations caused by turbulent the outer edge of the sub-layer is the bulk fluid concentration
oscillations in the outer flow or by irregularities in surface texture. C. Then, ∂C/∂n⊥ ≈ C/δc , and Eq. (1) can be reduced to (e.g.
At some point along the blade, the boundary-layer transitions to a Boudreau and Jorgensen 2001)
turbulent boundary layer with a viscous sub-layer, δs (Fig. 1). The
transition occurs near Rx = Ux/ν ≈ 105 , but this can be mod- Dm AC ub∗ 1/3
ṁ = = S Dm AC ∼ U (2)
ified by surface roughness (White 2008). If the blade length is δc 10ν c
less than the transition length, the boundary layer is laminar over
the entire blade. If the boundary layer becomes turbulent, the where the relations for δc introduced above are used. This is
viscous sub-layer will have a constant thickness set by the fric- called the mass-transfer-limited flux, because it is controlled by
tion velocity on the blade, ub∗ . The viscous sub-layer thickness is the mass transfer across the diffusive sub-layer. Because ub∗ is
between δs = 5ν/ub∗ and 10ν/ub∗ (e.g. Boudreau and Jorgensen correlated with the velocity (U ), Eq. (2) suggests ṁ ∼ U . In
2001). Within this layer, the flow is essentially laminar. other words, as the velocity increases, the diffusive sub-layer
thins, and the flux to the blade increases. This behaviour has been
observed for nutrient uptake by seagrasses (e.g. Koch 1994).
However, as the velocity increases, at some point, the physi-
cal rate of mass flux matches and then surpasses the biological
rate of incorporation at the surface, at which point the uptake
rate is controlled biologically. This is called the biologically (or
kinetically) limited flux rate. This transition was observed for
seagrasses between U = 4 and 6 cm s−1 (Koch 1994).
A flat plate is not always a good geometric model for a plant
surface. However, a generalized version of Eq. (2) will hold for
surfaces of any shape or rigidity, and mass-transfer limitation
Figure 1 The evolution of a boundary layer on a flat plate. The ver- by a diffusive sub-layer can occur on any surface. Specifically,
tical coordinate is exaggerated. The momentum boundary layer, δ, the mass flux can be described at any point on the surface by
grows with distance from the leading edge (x = 0). Initially, the bound- ṁ = Dm AC/δc . The problem lies in describing δc , which can vary
ary layer is laminar (shaded grey). At distance x, corresponding to
along the surface due to the surface shape or texture. For exam-
Rx = xU /ν ≈ 5 × 105 , the boundary layer becomes turbulent, except
for a thin layer near the surface that remains laminar, called the vis- ple, on an undulated blade of the kelp Macrocystis integrifolia,
cous (or laminar) sub-layer, δs . In water, the diffusive sub-layer, δc , is the laminar sub-layer is thinned at the apex of each undulation,
much smaller than the viscous sub-layer, with δc = δs S−1/3 This figure and thickened on the downstream side, relative to a flat blade
is adapted from Nepf (2012a) under the same mean flow (Hurd et al. 1997). Furthermore, blade
264 H.M. Nepf Journal of Hydraulic Research Vol. 50, No. 3 (2012)

motion may disturb the diffusive sub-layer, replacing the fluid stiffness. For a blade of length l, width d, thickness t, and den-
next to the surface with the fluid from outside the boundary sity, ρv , and in a uniform flow of horizontal velocity U , these
layer, which in turn creates an instantaneously higher concen- parameters are defined as
tration gradient at the surface and thus higher flux. This process
can be represented by a surface renewal model (Stevens et al. (ρ − ρv )gdtl 3
B= (3)
2003, Huang et al. 2011). Recent studies have documented blade EI
motions associated with turbulence (Plew et al. 2008, Sinis- 1 ρCD dU 2 l 3
calchi et al. 2012), and future studies should examine how the C= (4)
2 EI
turbulence-induced motion may enhance flux.
Here, E is the elastic modulus for the blade, I (= dt3 /12) is the
second moment of area, ρ is the density of water, and g is the
2.2 Flexibility and reconfiguration
acceleration due to gravity.
Because many aquatic plants are flexible, they can be pushed over The impact of reconfiguration on drag can be described by an
by currents, resulting in a change in morphology called reconfig- effective blade length, le , which is defined as the length of a rigid,
uration (e.g. Vogel 1994). The change in blade posture can alter vertical blade that generates the same horizontal drag as a flexible
light availability in competing ways. When a blade is pushed blade of total length l. Based on this definition, the horizontal
over, its horizontal projected area increases, which creates a drag force is Fx = (1/2)ρCD dle U 2 , where the drag coefficient,
greater surface area for light interception, but also increases shad- CD , is identical to that for rigid, vertical blades. The following
ing among neighbouring blades (Zimmerman 2003). In addition, relationships for effective length, le , and meadow height, h, are
because light attenuates with distance from the surface, plants based on the model described in Luhar and Nepf (2011, 2012):
pushed down away from the surface receive less light. Recon-
figuration also reduces flow resistance through two mechanisms. le 1 − 0.9C−1/3
=1− (5)
First, reconfiguration reduces the frontal area of the vegetation l 1 + C−3/2 (8 + B3/2 )
and, second, the reconfigured shape tends to be more stream- h 1 − C−1/4
lined (de Langre 2008). Because of reconfiguration, the drag =1− −3/5
(6)
l 1+C (4 + B3/5 ) + C−2 (8 + B2 )
on a plant increases more slowly with velocity than that pre-
dicted by the quadratic law. To quantify this, the relationship When rigidity is the dominant restoring force (C  B), Eq. (6)
between the drag force (F) and velocity (U ) has been expressed reduces to h/l ∼ C−1/4 ∼ (EI/U 2 )1/4 , which is similar to the
as F ∝ U 2+γ , with γ called the Vogel exponent. The Vogel scaling suggested by Kouwen and Unny (1973). Although
exponent has been observed to vary between γ = 0 (rigid) and Eqs. (5) and (6) were developed for individual blades, Luhar
γ = −0.7 (e.g. Albayrak et al. 2011). and Nepf (2012) demonstrated how they can be used to pre-
In practice, predictions of drag have used the standard dict the height (h) of a submerged seagrass meadow and how
quadratic law, but allow the reference area and drag coefficient the predicted h and le can then be used to predict the channel-
to vary with velocity. There has been a significant debate as to scale resistance for plants with simple blade morphology. How
which reference area best characterizes drag as the vegetation is these relationships can be extended to plants of more complex
pushed over (see the discussion of Sand-Jensen 2003 by Green morphology is an open research question.
2005a, Sukhodolov 2005, Statzner et al. 2006). Some recent
studies have addressed this debate by developing drag relation-
2.3 Future research challenges at the blade and individual
ships that incorporate the change in posture (e.g. Luhar and Nepf
plant scales
2011).
A flexible body in flow will adjust its shape until there is a While recent laboratory studies have advanced our understand-
balance between the drag force and the restoring force due to ing of how buoyancy and rigidity impact the reconfiguration of
body stiffness, for which scaling predicts F ∝ U 4/3 , which cor- individual blades, there are several obstacles to application in
responds to γ = 2/3 (e.g. Alben et al. 2002, de Langre 2008). the field. First, there is inadequate information on plant mate-
Although derived for fibres, scaling laws close to F ∝ U 4/3 have rial properties in the field (E, ρV ). Second, current models of
been observed for the leaves of individual plants (Albayrak et al. plant–flow interaction are based on blade morphology and con-
2011). Because many aquatic species have gas-filled sacs or sider only the interaction with a steady flow. Work is needed to
material density less than that of water, buoyancy may also act extend these theories to describe the following: (1) more complex
as a restoring force (Green 2005a). Dijkstra and Uittenbogaard morphology, for example, consisting of branches and leaves with
(2010) and Luhar and Nepf (2011) considered buoyancy and different material properties, and (2) interaction with an unsteady
rigidity together, in which case reconfiguration depends on two flow (waves and turbulence). Finally, future studies should exam-
dimensionless parameters. The Cauchy number, C, is the ratio of ine the impact of turbulence- and wave-induced plant motion on
drag to the restoring force due to rigidity. The buoyancy param- light availability and nutrient fluxes, providing an important link
eter, B, is the ratio of the restoring forces due to buoyancy and between hydraulics and plant function.
Journal of Hydraulic Research Vol. 50, No. 3 (2012) Vegetated channels 265

3 Uniform meadows of submerged vegetation Here, θ is the angle of channel inclination. The spatially-averaged
Reynolds stress appears in term (i). Spatial correlations in the
In this section, we consider a community of individual plants time-averaged velocity give rise to a dispersive stress, which
within a uniform, submerged meadow. The meadow geometry appears in term (ii). This term is negligible for l = ah > 0.1
is defined by the size of individual stems and blades and their (Poggi et al. 2004). Term (iii) is the viscous stress associated
number per bed area. If the individual stems or blades have a with vertical variation in
ū . Dx is the spatially-averaged drag
characteristic diameter or width d, and an average spacing S, associated with the canopy elements, which is represented by a
then the frontal area per volume is a = d/S 2 . Note that a can quadratic drag law (e.g. Kaimal and Finnigan 1994):
only be defined as an average over length scales greater than S,
and using this representation for meadow geometry, we forfeit 1 CD a
Dx =
ū |
ū | (9)
the resolution of flow at scales less than S. The meadow density 2 n
can also be described by the solid volume fraction occupied by Based on the momentum balance, Belcher et al. (2003) defined
the canopy elements, φ, or the porosity, n = (1 − φ). If the indi- this canopy-drag length scale, Lc , as
vidual elements approximate a circular cylinder, for example,
reed stems, then φ = (π/4)ad. If the morphology is strap like, 2(1 − φ)
with blade width d and thickness b, then φ = db/S 2 = ab. Note Lc = (10)
CD a
that d and S, and therefore a, can vary spatially and specifically
over the height of a canopy. In addition, for flexible vegetation, as This represents the length scale over which the mean and turbu-
flow speed increases, the canopy height may decrease, increasing lent velocity components adjust to the canopy drag. Since most
both a and φ. Finally, a non-dimensional measure of the canopy aquatic canopies have high porosity (φ < 0.1), this scale is often
density is the frontal area per bed area, l, known as the roughness approximated by 2(CD a)−1 .
density (Wooding et al. 1973). For meadow height h and z = 0
at the bed, 3.1 Stem-scale turbulence
 h
l= adz = ah (7) Branches and stems with an orientation perpendicular to the flow
z=0 can generate turbulence. The stem diameter (or blade width) d
with the right-most expression being valid for vertically uni- defines the stem Reynolds number, Rd = Ud/ν. For Rd >≈ 100,
form a. the canopy elements will generate vortices of scale d, which is
Within a canopy, a flow is forced to move around each branch called stem-scale turbulence (e.g. Nepf 2012a). If the stem den-
or blade, so that the velocity field is spatially heterogeneous at sity is high, such that the mean spacing between stems (S) is
the scale of these elements. A double-averaging method is used less than d, the turbulence is generated at the scale S (Tanino
to remove the element-scale spatial heterogeneity, in addition to and Nepf 2008). Even for very sparse canopies, the production
the more common temporal averaging (e.g. Wilson and Shaw of turbulence within stem wakes is comparable to or greater than
1977, Nikora et al. 2007). The velocity vector − →u = (u, v, w) the production by bed shear (Lopez and Garcia 1998, Nepf and
corresponds to the coordinates (x, y, z), respectively. The instan- Vivoni 2000). Therefore, turbulence level cannot be predicted
taneous velocity and pressure (p) fields are first decomposed into from the bed-friction velocity, as it is done for open-channel
a time average (overbar) and deviations from the time average flows. Instead, it is a function of the canopy drag. Vortex genera-
(single prime). The time-averaged quantities are further decom- tion in stem wakes drains energy from the mean flow (expressed
posed into a spatial mean (angle bracket) and deviations from as mean canopy drag) and feeds it into the turbulent kinetic
the spatial mean (double prime). The spatial averaging volume energy. If this conversion is 100% efficient, then the rate at which
is thin in the vertical plane, to preserve vertical variation, and turbulent energy is produced is equal to the rate of work done by
large enough in the horizontal plane to include several stems the flow against the canopy drag. If we assume that the energy
(> S). is extracted at the length scale
, the turbulent kinetic energy (k)
Applying this averaging scheme to a homogeneous canopy, in the canopy may be estimated as (Tanino and Nepf 2008)
the momentum equation in the streamwise direction is (e.g. 
 1/3
Nikora et al. 2007)


≈ CD (11)

ū d (1 − φ)π
D
ū 1 ∂n
p̄ 1 ∂ 1 ∂
= g sin θ − − n
u w − n
ū w̄
Dt nρ ∂x n ∂z n ∂z Here,
is the smaller than d or S. In fact, only the form drag is
(i) (ii) converted into turbulent kinetic energy. The viscous drag is dis-
sipated directly to heat. For stiff canopies, or near the rigid base
1 ∂ ∂

+ ν n − Dx of most stems, the drag is mostly form drag, and Eq. (11) is a rea-
n ∂z ∂z
sonable approximation. However, in the streamlined portion of
(iii) (8) flexible submerged plants, the drag is predominantly viscous, and
266 H.M. Nepf Journal of Hydraulic Research Vol. 50, No. 3 (2012)

Eq. (11) would be an overestimate of the stem-scale turbulence 2007). Since the assumption of CD ≈ 1 is often reasonable (e.g.
production (Nikora and Nikora 2007). Table 1 in Nepf 2012a), the measured and theoretical limits are
An interesting nonlinear behaviour emerges when we com- consistent.
pare conditions of different stem densities under the same driving If the velocity profile contains an inflection point, it is unstable
force (i.e. the same potential and/or pressure gradient). The to the generation of Kelvin–Helmoltz (KH) vortices (e.g. Rau-
details of this comparison are given in Nepf (1999). Because the pach et al. 1996). These structures dominate the vertical transport
vegetation offers resistance to flow, the velocity within a canopy at the canopy interface (e.g. Ghisalberti and Nepf 2002). These
is always less than the velocity over a bare bed under the same vortices are called canopy-scale turbulence, to distinguish it from
external forcing, and the canopy velocity decreases monoton- the much larger boundary-layer turbulence, which may form
ically with increasing stem density (or φ). However, changes above a deeply submerged or unconfined canopy, and the much
in turbulent kinetic energy,
k̄ , reflect competing effects as smaller stem-scale turbulence. Over a deeply submerged (or ter-
stem density (φ) increases, that is, turbulence intensity,
k̄ /
ū 2 , restrial) canopy (H /h > 10), the canopy-scale vortices are highly
increases (Eq. 11), but
ū 2 decreases, which together produce a three dimensional due to their interaction with boundary-layer
nonlinear response. As stem density (or φ) increases from zero, turbulence, which stretches the canopy-scale vortices, enhanc-

k̄ initially increases, but eventually decreases as φ increases fur- ing secondary instabilities (Finnigan et al. 2009). However, with
ther. The fact that at some stem densities the near-bed turbulence shallow submergence (H /h ≤ 5), which is common in aquatic
level within a meadow can be higher than that over the adjacent systems, large-scale boundary-layer turbulence is not devel-
bare bed has important implications for sediment transport. This oped, and the canopy-scale vortices dominate the turbulence both
is discussed further in the next section. within and above the meadow (Ghisalberti and Nepf 2005, 2009).
Within a distance of about 10h from the canopy’s leading edge,
the canopy-scale vortices reach a fixed scale and a fixed pene-
3.2 Sparse and dense meadows tration into the canopy (δe in Fig. 2; Ghisalberti and Nepf 2002,
2004, 2009). The final vortex and shear-layer scale is reached
We now consider a submerged meadow of height h in water
when the shear production that feeds energy into the canopy-scale
of depth H (Fig. 2). For a submerged meadow, there are two
vortices is balanced by the dissipation by the canopy drag. This
limits of flow behaviour, depending on the relative importance
balance predicts the following scaling, which has been verified
of the bed shear and meadow drag. If the meadow drag is
with observations (Nepf et al. 2007),
smaller than the bed drag, then the velocity follows a turbulent
boundary-layer profile, with the vegetation contributing to the 0.23 ± 0.6
bed roughness. This is the sparse canopy limit (Fig. 2a). In this δe = (12)
CD a
limit, the turbulence near the bed will increase as stem density
increases. Alternatively, in the dense canopy limit, the canopy Note that Eq. (12) only applies to canopies that form a shear
drag is larger than the bed stress, and the discontinuity in drag layer (i.e. CD ah ≥ 0.1). For CD ah = 0.1–0.23, the canopy-scale
at the top of the canopy generates a region of shear resembling turbulence penetrates to the bed, δe = h, creating a highly tur-
a free shear layer, including an inflection point near the top of bulent condition over the entire canopy height (Fig. 2b). At
the canopy (Fig. 2b,c). From scaling arguments, the transition higher values of CD ah, the canopy-scale turbulence does not
between sparse and dense limits occurs at l = ah = 0.1 (Belcher penetrate to the bed, δe < h (Fig. 2c). If the submergence ratio
et al. 2003). From measured velocity profiles, a boundary-layer H /h < 2, Eq. (12) for δe is not applicable, as the interaction
form with no inflection point is observed for CD ah < 0.04, and a with the water surface diminishes the strength and size of the
pronounced inflection point appears for CD ah > 0.1 (Nepf et al. canopy-scale vortices (Nepf and Vivoni 2000). Canopies for

Figure 2 The mean velocity profiles through submerged meadows of increasing roughness density (ah). The meadow height is h. Water depth is H .
(a) For ah < 0.1 (sparse regime), the velocity follows a rough boundary-layer profile. (b) For ah ≥ 0.1, a region of strong shear at the top of the canopy
generates canopy-scale turbulence. The canopy-scale turbulence penetrates a distance δe = [0.23 ± 0.06](CD a)−1 into the canopy. (c) For ah > 0.23
(dense regime), δe < h, and the bed is shielded from the canopy-scale turbulence. Stem-scale turbulence is generated throughout the meadow. This
figure is adapted from Nepf (2012b)
Journal of Hydraulic Research Vol. 50, No. 3 (2012) Vegetated channels 267

which δe /h < 1 (Fig. 2c) shield the bed from strong turbu- flexible vegetation, by solving iteratively for the meadow height
lence and turbulent stress. Because turbulence near the bed plays and velocity profile (Dijkstra and Uittenbogaard 2010, Luhar and
a role in resuspension, these dense canopies are expected to Nepf 2012).
reduce resuspension and erosion. Consistent with this, Moore
(2004) observed that resuspension within a seagrass meadow
was reduced, relative to bare-bed conditions, only when the 4 Emergent canopies of finite width and length
above-ground biomass per area was greater than 100 g/m2 (dry
mass), which corresponds to ah = 0.4 (Luhar et al. 2008). In The previous section described the flow near a submerged canopy
a similar study, Lawson et al. (2012) measured sediment ero- that was fully developed and uniform along and across the flow.
sion in beds of different stem densities. Using the blade length While the fully developed case is important, it is not represen-
(8 cm) and width (3 mm) provided in that paper, we convert tative of all the field conditions. In this section, we consider
the stem density into a roughness density ah. Between 80 and geometries that are finite in length and width.
300 stems m−2 (ah = 0.02–0.07), erosion increased with increas-
ing stem density, consistent with sparse canopy behaviour, that 4.1 Long emergent canopies of finite width
is, the stem-scale turbulence augmented the near-bed turbulence
and increased with increasing stem density. However, above In river channels, emergent vegetation often grows along the
500 stems m−2 (ah = 0.12), bed erosion was essentially elim- bank, creating long regions of vegetation of finite width b (Fig. 3).
inated (Lawson et al. 2012). Both the Moore and Lawson studies Long patches may also exist at the centre of a channel, and to rec-
demonstrate a stem density threshold, above which the near- ognize the geometric similarity with bank vegetation, we define
bed turbulence becomes too weak to generate resuspension and b as the half-width for in-channel vegetation (Fig. 4). Let the
erosion. The threshold is roughly consistent with the roughness streamwise coordinate be x, with x = 0 at the leading edge. The
density transition suggested by Eq. (12) and depicted in Fig. 2. lateral coordinate is y, with y = 0 at the side boundary for bank
The regimes depicted in Fig. 2 give rise to a feedback between vegetation (Fig. 3) or at the centreline for in-channel vegetation
optimum meadow density and substrate type. Because dense (Fig. 4). The streamwise and lateral velocities are (u, v), respec-
canopies reduce near-bed turbulence, they promote sediment tively. Because the vegetation provides such high drag, relative
retention. In sandy regions, which tend to be nutrient poor, the to the bare bed, much of the flow approaching from upstream is
preferential retention of fines and organic material, that is, mud- deflected away from the patch. The deflection begins upstream of
dification, enhances the supply of nutrients to the canopy, so that the patch over a distance that is set by the scale b, and it extends a
dense canopies provide a positive feedback to canopy health in distance xD into the vegetation (Zong and Nepf 2010). Rominger
sandy regions (van Katwijk et al. 2010). In contrast, in regions and Nepf (2011) showed that xD scales with the larger of the two
with muddy substrate, which is more susceptible to anoxia, sparse length scales b or Lc = 2(CD a)−1 . It is only after the deflection is
meadows (CD ah ≤ 0.1) may be more successful, because the complete (x > xD ) that the shear layer with KH vortices develops
enhanced near-bed turbulence removes fines, leading to a sandier along the lateral edge of the vegetation. The KH vortices domi-
substrate that is less prone to anoxia. nate the mass and momentum exchange between the vegetation
Both the boundary-layer profile of a sparse canopy and the and the adjacent open flow (White and Nepf 2007).
mixing layer profile of a dense canopy have been observed in the The initial growth and the final scale of the horizontal shear-
field, in seagrass meadows (Lacy and Wyllie-Echeverria 2011), layer vortices and their lateral penetration into the patch, δL , are
and in river canopies (Sukhodolov and Sukhodolova 2010).
Although both profiles have been observed, modelling efforts
have focused on the dense canopy limit. Many methods divide
the flow into a uniform layer within the vegetation and a logarith-
mic profile above the vegetation. However, if there is a poor scale
separation between plant height and flow depth, it is unlikely that
a genuine logarithmic layer exists.
A number of studies have proposed models for the full velocity
profile, that is, both within and above the canopy. These studies
have utilized three general approaches: (i) simple momentum Figure 3 Top view of a channel with a long patch of emergent vegeta-
balances that segregate the flow into a vegetated layer of depth h tion along the right bank (grey shading). The width of the vegetated zone
and an overflow of depth H − h (e.g. Huthoff et al. 2007, Cheng is b. The flow approaching from upstream has uniform velocity U0 . The
flow begins to deflect away from the patch at a distance b upstream and
2011); (ii) analytical descriptions using an eddy viscosity model,
continues to decelerate and deflect until distance xD . After this point, a
νt , to define the turbulent stress (e.g. Baptiste 2007, Poggi et al. shear layer forms on the flow-parallel edge and shear-layer vortices form
2009); and (iii) numerical models with first- or second-order by KH instability. These vortices grow downstream, but subsequently
turbulence closures (e.g. Shimizu and Tsujimoto 1994, Lopez reach a fixed width and fixed penetration distance into the vegetation,
and Garcia 2001). Some models reflect the bending response of δL . This figure is adapted from Zong and Nepf (2010)
268 H.M. Nepf Journal of Hydraulic Research Vol. 50, No. 3 (2012)

If the patch width, b, is greater than the penetration distance,


δL (CD ab > 0.5, according to Eq. 13), turbulent stress does not
penetrate to the centreline of the patch, and the velocity within
the patch (U1 , Fig. 3) is set by a balance of potential gradient (bed
and/or water surface slope) and vegetation drag. In contrast, for
CD ab < 0.5, turbulent stress can reach the patch centreline, and
U1 is set by the balance of turbulent stress and vegetation drag.
Detailed formulations for U1 are given in Rominger and Nepf
(2011).
The centre of each vortex is a point of low pressure, which, for
shallow flows, induces a wave response across the entire patch
and specifically beyond δL from the edge (White and Nepf 2007,
2008). The wave response within the vegetation has been shown
to enhance the lateral (y) transport of suspended particles, above
that predicted from stem turbulence alone (Zong and Nepf 2011).
For in-channel patches, shear layers develop along both flow-
parallel edges, and the vortices along each edge interact across
the canopy width (Fig. 4a). The low-pressure core associated with
Figure 4 (a) Top view of emergent vegetation with two flow-parallel each vortex produces a local depression in the water surface, such
edges. The patch width is 2b. The coherent structures on either side of the that the passage of individual vortices can be recorded by a sur-
patch are out of phase. The passage of each vortex core is associated with face displacement gage. A time record of surface displacement
a depression in surface elevation, which is measured at the patch edges measured on opposite sides of a patch (A1 and A2 in Fig. 4b)
(A1 and A2). The velocity is measured mid-patch (square). (b) Data
measured for a patch of width b = 10 cm in a channel with flow veloc-
shows that there is a half-cycle phase shift (π radians) between
ity U0 = 10 cm s−1 . The patch centreline velocity is U1 = 0.5 cm s−1 . the vortex streets that form on either side of the patch. Because
The surface displacements measured at A1 (heavy dashed line) and at A2 the vortices are a half cycle out of phase, when the pressure (sur-
(heavy solid line) are a half cycle (π radians) out of phase. The resulting face elevation) is at a minimum on side A1, it is at a maximum
transverse pressure gradient imposed across the patch generates trans- at side A2. The resulting cross-canopy pressure gradient induces
verse velocity within the patch (thin line), which, as in a progressive
a transverse velocity within the canopy (Fig. 4b) that lags the
wave, lags the lateral pressure gradient by a quarter cycle (π/2 radians).
This figure is adapted from Rominger and Nepf 2011 lateral pressure gradient by π/2, that is, a quarter cycle. The syn-
chronization of the vortex streets occurs even when the vortex
penetration is less than the patch width, δL /b < 1, and it signifi-
depicted in Fig. 3. The vortices extend into the open channel
cantly enhances the vortex strength and the turbulent momentum
over length δ0 ∼ H /Cf , where Cf is the bed friction (White and
exchange between the open channel and vegetation (Rominger
Nepf 2007). There is no direct relation between δL and δ0 . As
and Nepf 2011). More importantly, the vortex interaction intro-
expected from the discussion of vertical canopy-shear layers,
duces significant lateral transport across the patch. For example,
δL ∼ (CD a)−1 . However, the scale factor observed for lateral
the data shown in Fig. 4(b) correspond to a patch with centre-
shear layers (denoted by subscript L) is twice that measured
line velocity U1 = 0.5 cm s−1 . The lateral velocity induced by
for vertical shear layers above submerged meadows (δe , Fig. 2,
the vortex pressure field is nearly one order of magnitude larger,
Eq. 12). Based on the study of White and Nepf (2007, 2008),
with maximum values of 3.5 cm s−1 (vrms = 2.2 cm s−1 , Fig. 4b).
0.5 ± 0.1 Using the period of vortex passage (T = 10 s), the lateral excur-
δL = (13) sion of a fluid parcel during each vortex cycle was found to be
CD a
10 cm (= vrms T /2), which is comparable to the half-width of the
The difference between δL and δe may be due to the difference patch, b = 10 cm, indicating that fluid parcels in the centre of
in flow geometry relative to the model canopy. Specifically, in the patch can be drawn into the free stream and vice versa, dur-
experiments with vertical circular cylinders (as in White and ing each vortex passage. This cycle of flushing can significantly
Nepf 2007), the cylinder presents a different geometry to vor- reduce the patch retention time and may even control it.
tices rotating in the horizontal plane than to the vortices rotating
in the vertical plane. Also note that a wider range of canopy
4.2 Circular patches of vegetation
morphology, including field measurements with real vegetation,
and a wider range of flow speeds were used to determine the A circular patch with diameter D (Fig. 5) is used as a model
scale factor for δe (Nepf et al. 2007). An estimate of 0.5 for δL for a patch of vegetation with length and width smaller than
is based only on one set of flume experiments with rigid circular the channel width. We consider emergent patches, so that the
cylinders. Whether, or not, the difference in the scale factor is flow field is roughly two dimensional (x–y). Because the patch
significant for field conditions is yet to be determined. is porous, the flow passes through it, and this alters the wake
Journal of Hydraulic Research Vol. 50, No. 3 (2012) Vegetated channels 269

Figure 5 Top view of a circular patch of emergent vegetation with patch diameter D. The upstream, open-channel velocity is U0 . Stem-scale
turbulence is generated within the patch, but dies out quickly behind the patch. The flow coming through the patch (U1 ) blocks interaction between
the shear layers at the two edges of the patch, which delays the onset of the patch-scale vortex street by a distance L1 . Tracer (grey line) released from
the outermost edges of the patch comes together at a distance L1 downstream from the patch and reveals the von Karman vortex street

structure relative to that of a solid body (Castro 1971, Zong and


Nepf 2012). Directly behind a solid body, there is a region of
recirculation, followed by a von Karman vortex street. The wake-
scale mixing provided by the von Karman vortices allows the
velocity in the wake to quickly return (within a few diameters) to
a velocity comparable to the upstream velocity (U0 ). In contrast,
the wake behind a porous obstruction (patch of vegetation) is
much longer, because the flow entering the wake through the
patch (called the bleed flow) delays the onset of the von Karman
vortex street until a distance L1 behind the patch. As a result,
the velocity at the centreline of the wake, U1 , remains nearly
constant over distance L1 . Within this region, both the velocity
and turbulence are reduced, relative to the adjacent bare bed, so
that it is a region where deposition is likely to be enhanced (see
the discussion in Section 6).
The delayed onset of the von Karman vortex street is visual-
ized using traces of dye injected at the outermost edges of the
patch (grey streaks in Fig. 5). Because the near wake is fed only Figure 6 The flow blockage (CD aD, with CD assumed to be 1) deter-
by water entering from upstream through the patch, it contains no mines (a) the velocity behind the patch U1 and (b) the length of the near
dye and appears as a clear region behind the patch, in between wake, L1 . (a) For low flow blockage, the velocity ratio, U1 /U0 , fits a
simple, linear relationship (Eq. 14, shown with solid and dashed (SD)
the two dye streaks. After distance L1 , the dye streaks come
lines). For high flow blockage, the exit velocity is a small fraction of
together, and a single, patch-scale, von Karman vortex street is U1 , but non-zero, until aD > 10, at which point U1 is indistinguishable
formed. Note that Fig. 5 is a snapshot in time, capturing one from zero. (b) For low flow blockage, L1 can be predicted from Eqs. (14)
phase of the unsteady vortex cycle. As the vortex cores migrate and (15) and becomes constant (L1 /D = 2.5) for high flow blockage.
downstream, the flow field at any fixed point oscillates with fre- Model predictions are shown as black lines. Black circles indicate mea-
quency, f , which is set by the patch scale D. The patch-scale surements obtained using a circular array of circular cylinders (Chen
et al. 2012)
vortex street follows the same scaling as a solid body, with the
Strouhal number S = fD/Uo ≈ 0.2 (Zong and Nepf 2012).
Both U1 and L1 depend on the patch diameter, D, and the does become zero, and the flow field around the porous patch
drag length scale, Lc ∼ (CD a)−1 , which together form a dimen- becomes identical to that around a solid obstruction (Nicolle and
sionless parameter, CD aD, called the flow blockage (Chen et al. Eames 2011, Zong and Nepf 2012). This transition is also seen
2012). For low flow blockage (small CD aD), U1 /U0 decreases in the length scale, L1 , discussed below.
linearly with CD aD (Fig. 6a) Using CD = 1, a reasonable linear Zong and Nepf (2012) suggested that L1 may be predicted
fit is from the linear growth of the shear layers located on either side
U1
= 1 − [0.33 ± 0.08]CD aD (14) of the near-wake region, from which they derived
U0
For high flow blockage, U1 is negligibly small (U1 /U∞ ≈ 0.03), L1 1 (1 + U1 /U2 ) 1 (1 + U1 /U0 )
= ≈ (15)
but not zero. However, at some point around CD aD = 10, U1 D 4S1 (1 − U1 /U2 ) 4S1 (1 − U1 /U0 )
270 H.M. Nepf Journal of Hydraulic Research Vol. 50, No. 3 (2012)

where S1 is a constant (0.10 ± 0.02) across a wide range of D and begun to describe the flow structure near the leading and trail-
φ (Zong and Nepf 2012). If the channel width is much greater than ing edges of a canopy and within the gaps between canopies
the patch diameter, we may assume that U2 ≈ U0 , resulting in (e.g. Sukhodolov and Sukhodolova 2010, Zong and Nepf 2010,
the right-most expression in Eq. (15). Predictions for L1 /D based Folkard 2011, Siniscalchi et al. 2012). However, there is much
on Eqs. (14) and (15) do a good job representing the observed work to be done. Future studies should enable the prediction of
variation in L1 with CD aD (Fig. 6b). Note that even as the velocity velocity and turbulence within and adjacent to a patch of any
behind the patch approaches zero, the delay in the vortex street size and depth of submergence and to use this information to
persists, with L1 /D = 2.5. However, when CD aD becomes high understand the patch-scale drag as well as the feedbacks between
enough that there is no bleed flow (U1 = 0), the wake resembles patch geometry and channel bed evolution (see the discussion
that observed for a solid body, with a recirculation zone and in Section 6). Finally, more studies are needed to explore the
vortex street forming directly behind the patch, L1 ≈ 0. The data transition between flow resistance dominated by the stem (leaf)-
shown in Fig. 6 suggest that this occurs for CD aD > 10. scale drag to flow resistance dominated by the patch-scale drag.
The wake transition described above has implications for the In the next section, we consider flow resistance at the channel
characterization of drag contributed by finite patches. As noted reach scale and again find that the patch-scale geometry is more
by Folkard (2010), drag is produced at two distinct scales: the important than the leaf-scale geometry.
leaf and stem scale and the patch scale. For low flow blockage
patches, there is sufficient flow through the patch that the stem-
and leaf-scale drag dominates the flow resistance, that is, the flow 5 Reach-scale hydraulic resistance
resistance can be represented by the integral of CD au2 over the
patch interior, with u being the velocity within the patch. How- 5.1 Recent trends
ever, for high flow-blockage patches, there is negligible flow At the scale of the channel reach, flow resistance due to vege-
through the patch, and the integral of CD au2 over the patch inte- tation is determined primarily by the blockage factor, Bx , which
rior is irrelevant. The flow response to a high flow-blockage patch is the fraction of the channel cross-section blocked by vegeta-
is essentially identical to the flow response to a solid obstruc- tion (Green 2005b, Luhar et al. 2008, Nikora et al. 2008). For
tion of the same patch frontal area, Ap . Thus, the flow resistance a patch of height h and width w in a channel of width W and
provided by the patch should be represented by the patch-scale depth H , Bx = wh/WH. These studies show strong correlations
geometry, that is, CD Ap U 2 , with U being the channel velocity. between Bx and Manning’s roughness coefficient, nM , noting that
This idea is supported by measurements of flow resistance pro- the relationship is nonlinear. These observations are in agreement
duced by sparsely distributed bushes (Righetti 2008). A bush with those of Ree (1949) and Wu et al. (1999), who showed that
consists of a distribution of stems and leaves and so is a form of resistance to flow in channels lined with vegetation is influenced
vegetation patch. The flow resistance generated by the bushes fit primarily by the submergence ratio, H /h. For vegetation that
the quadratic model, ρCD Ap U 2 , and notably CD was O(1), sim- fills the channel width, Bx = h/H . A few studies suggest that
ilar to a solid body. Thus, although porous, the bush generated the vegetation distribution may also influence the resistance and
drag that was comparable to that of a solid object of the same size specifically that greater resistance is produced by distributions
(Ap ). It is worth noting that CD decreased somewhat (from 1.2 to with a greater interfacial area between vegetated and unvegetated
0.8) as the channel velocity increased. This shift is most likely regions (e.g. Bal et al. 2011). Luhar and Nepf (2012) quantified
due to the reconfiguration of stems and leaves that reduced Ap . the impact of interfacial area by considering channels with the
Since this reconfiguration was not accounted for in the analysis, same blockage factor (Bx ), but a different number (N ) of patches.
it shows up as an apparent decrease in CD . They showed that for realistic values of N , the resistance is
increased by at most 20%, so that N = 1 is a reasonable sim-
4.3 Future research challenges at the canopy and patch plifying assumption. For N = 1, the momentum balance leads
scales to the following equations for Manning’s roughness coefficient
(Luhar and Nepf 2012):
We have a fairly complete conceptual picture of a uniform flow
through a dense canopy (CD ah > 0.1) and a reasonable under-  
 
CD aH 1/2
g 1/2
standing of the transition from sparse (CD ah  0.1) to dense For Bx = 1 : nM = (16)
KH 1/6 2
flow behaviours. However, much of this understanding has come  1/2   1/2
from laboratory studies, in which CD ah is easy to quantify. To g C∗
For Bx < 1 : nM 1/6
= (1 − Bx )−3/2 (17)
connect these models to the field, we need to understand how best KH 2
to measure and predict both CD a and h for real vegetation, noting
that both of these may vary with channel discharge (for flexible The constant K = 1 m1/3 s−1 is required to make the equations
plants) and seasonally (with plant growth and decay). Second, dimensionally correct. Note that Eq. (16) is valid when Bx = 1,
for canopies of finite width and length, the flow transition at the which indicates that vegetation covers the entire cross-section,
boundaries must also be understood. A few recent studies have width, and depth. The coefficient C∗ parameterizes the shear
Journal of Hydraulic Research Vol. 50, No. 3 (2012) Vegetated channels 271

stress at the interface between vegetated and unvegetated regions, 10


and C∗ = 0.05–0.13, based on fits to field data (Luhar and Nepf
Eq. 16
2012). While Eq. (17) seems attractively simple, remember that
for flexible vegetation, Bx (=wh/WH ) will be a function of flow Eq.
nM 1
speed, because the meadow height, h, decreases as flow speed 18
increases.
It is instructive to consider the case of submerged vegetation
0.1
that fills the channel width, such that the resistance is a func- 0.1 1 10
tion only of the submergence depth (H /h). This case has been H/h
considered in many papers on channel resistance (e.g. Ree 1949,
Figure 7 Manning’s coefficient (nM ) versus depth ratio (H /h). Most
Wu et al. 1999). For this case, Manning’s coefficient may be channel vegetation is flexible, so that increasing velocity is associated
represented as (Luhar and Nepf 2012) with a decrease in vegetation height (h), that is, h ∼ 1/V , and the previ-
ously noted nonlinear trend of nM with VR (e.g. Ree 1949) is captured
  by the trends of nM with H /h, as expressed through Eq. (16), for emer-
g 1/2
For H /h > 1 : nM gent conditions, and Eq. (18), for submerged conditions, based on the
KH 1/6 work of Luhar and Nepf (2012)
     1/2 −1
2 1/2 h 3/2 2 h
= 1− + (18)
C∗ H CD ah H expect Bx to be a function of channel discharge. Although the
reconfiguration of individual blades has been described by pre-
If CD ah > C∗ , a common field condition, the second term drops vious research (discussed in Section 2), the reconfiguration of
out and Eq. (18) reverts to Eq. (17), because for vegetation plants and patches of more complex morphology requires further
covering the full channel width, Bx = h/H . study. In addition, the breakage and/or dislodgement of plants
Several researchers have noted a nonlinear relationship during high-flow conditions may create significant shifts in flow
between nM and a surrogate of channel Reynolds number, VR, resistance between the rising and falling limbs of a flood (e.g.
with V being the channel average velocity and R the hydraulic Boelscher et al. 2010). Finally, we need to understand at what
radius (e.g. Ree 1949). Folkard (2011) provided a useful dis- spatial scale should Bx be resolved for accurate prediction of
cussion of this relationship, noting that the peak in hydraulic flow resistance and how best to make these measurements to
resistance occurs at the transition from emergent to submerged capture seasonal variation. These points are highlighted again in
conditions. Because most channel vegetation is flexible, an Section 7.
increase in velocity is associated with a decrease in vegetation
height, that is, h ∼ 1/V . In addition, for wide channels, R = H ,
so that H /h ∼ VR. This suggests that the observed trends of 6 Sediment transport and channel morphodynamics
nM with VR can be mostly explained by the trends of nM with
H /h, as expressed through Eq. (16), for emergent conditions, and By baffling the flow and reducing the bed stress, vegetation
Eq. (18), for submerged conditions. As an example, nM was cal- creates regions of sediment retention (e.g. Abt et al. 1994, Cot-
culated from Eqs. (16) and (18) using CD ah = 10 and C∗ = 0.1 ton et al. 2006). Vegetation can also enhance channel stability
(Fig. 7). If the plants are emergent (H /h < 1), the vegetation drag and reduce bank erosion (Afzalimehr and Dey 2009, Pollen-
increases with increasing depth ratio (H /h), because the total Bankhead and Simon 2010). Because of the positive impacts
vegetation area per bed area (aH ) increases as H /h increases, that vegetation provides for water quality, habitat, and channel
Eq. (16). However, if the plants are submerged (H /h > 1), the stability, researchers now advocate replanting and maintenance
hydraulic resistance decreases as H /h increases. This is made of vegetation in rivers (e.g. Mars et al. 1999, Pollen and Simon
more obvious by noting that as H /h increases above 1, the sec- 2005). However, to design restoration schemes that are sustain-
ond term in Eq. (18) quickly becomes negligible, reducing to able, we need a better understanding of how the distribution
nM = (C∗ /2)1/2 (1 − H /h)−3/2 . The curve shown in Fig. 7 is and density of vegetation determine channel stability. Similarly,
visually similar to the many empirical curves presented for nM numerous publications (e.g. NRC 2002) and government poli-
versus VR (e.g. Ree 1949, Wu et al. 1999). cies (CBEC 2003) advocate for fluvial vegetation as traps for
sediments and other pollutants, but few studies have measured
the actual storage rates. These gaps in understanding must be
5.2 Future research challenges at the reach scale
addressed through collaborations between fluvial hydraulics and
The observations and theory discussed above suggest that the geomorphology.
blockage factor, Bx , is the geometric description of vegetation While most previous studies observed enhanced deposition
that is most relevant to the reach-scale resistance. However, there within regions of vegetation, the opposite trend has also been
are still questions to be addressed before reliable predictions of observed, that is, the removal of fines from within a patch. Specif-
channel resistance will be possible. For flexible vegetation, we ically, van Katwjk et al. (2010) observed that sparse patches of
272 H.M. Nepf Journal of Hydraulic Research Vol. 50, No. 3 (2012)

vegetation were associated with sandification, a decrease in fine


particles and organic matter, which is most likely attributed to
higher levels of turbulence within the sparse patch, relative to the
adjacent bare regions (see the discussion in Section 3). Elevated
turbulence levels have also been observed within the leading
edge of a patch, resulting in net deposition that is lower within
the leading edge than in the adjacent bare bed, despite the fact
that the mean flow is reduced (Cotton et al. 2006, Zong and
Nepf 2011, 2012). At the same time, deposition of fine sediment
has been observed in the wake behind a patch (Tsujimoto 1999,
Chen et al. 2012), which, together with the diminished deposi-
tion near the leading edge, may explain why patches grow in
length predominantly in the downstream direction (Sand-Jensen Figure 8 The transition in the bed form, from (a) migrating dunes to (b)
and Madsen 1992). Furthermore, observations given in Chen a fixed pattern of scour associated with individual plants, in a sand-bed
river upon the addition of vegetation to the point bars. Photographs
et al. (2012) suggest that the deposition of fine material is lim-
taken by Jeff Rominger during the Outdoor StreamLab experiment at
ited to the near-wake region defined by L1 (Fig. 5), where both Saint Anthony Falls Laboratory 2008 (Rominger et al. 2010)
the mean and turbulent velocities are depressed. The formation
of the von Karman vortex street at the end of the near wake sig-
nificantly elevates the turbulence level, inhibiting deposition. By
There is also significant spatial variability in bed stress at the
extension, we conjecture that the onset of the von Karman vor-
scale of individual stems, for example, similar to that observed
tex street may set the maximum streamwise extension of a patch.
around piers (Escauriaza and Sotiropoulos 2011). The spatial
The lateral growth of a patch may also be influenced by a hydro-
pattern of bed stress imposed by the stems is revealed, in part, by
dynamic control. Specifically, the deflection of a flow around a
the scour holes observed around individual stems (e.g. Bouma
vegetated region produces a locally enhanced flow at its edges
et al. 2007). Indeed, in sand-bed rivers, the addition of vegetation
that leads to erosion that may inhibit lateral expansion (Bouma
can lead to a transition in bed forms, from migrating dunes to a
et al. 2007, Bennett et al. 2008, Rominger et al. 2010). These
fixed pattern of scour associated with individual plants or stems
examples of the interplay between flow and patch growth demon-
(Fig. 8). To the extent that migrating dunes contribute to sedi-
strate feedbacks between vegetation, flow, and morphodynamics.
ment transport, the elimination of this migration will certainly
There is much to be learned about these feedbacks, and this
impact the bed-load transport.
understanding is vital in the planning of successful restoration
projects.
Setting aside the complexity of spatially heterogeneous veg-
6.1 Bed-shear stress within a uniform canopy of vegetation
etation discussed above, even for homogeneous regions of
vegetation, we lack a good description of sediment transport. This If we compare channels with and without vegetation, but with
is currently hampered by two problems. First, while it is tempting the same potential forcing, the mean bed stress, τbed = ρu∗2 , is
to apply sediment transport models developed for open-channel reduced in the presence of vegetation. This is reflected in the

flows to predict sediment transport in regions of vegetation, it ratio u∗ / gHS, with S describing the slope of the bed and/or
is not clear whether this is a valid approach. Open-channel flow water surface. This ratio is 1 for open-channel flows and less
models relate sediment transport to the mean bed stress (e.g. than 1 in a vegetated channel. Using a k–ε model to represent
Julien 2010). However, new studies point to the important role a flow through rigid vegetation (H /h = 3), Lopez and Garcia
of turbulence in initiating sediment motion (e.g. Nino and Gar- (1998) showed that this ratio drops off steadily with increasing

cia 1996, Vollmer and Kleinhans 2007, Celik et al. 2010). In an aH (and ah), approaching u∗ / gHS = 0.1 at aH = 3 (ah = 1).
open channel, the turbulence is linked to the mean bed stress, That is, the bed stress with vegetation is reduced to just 10% of
so that traditional sediment transport models, based on the mean the bare-bed value.
bed stress, may empirically incorporate the role of turbulence While it is not yet clear whether sediment transport within
into their parameterization. However, in vegetated regions, the vegetation can be predicted from the mean bed stress alone, it is
turbulence level is set by the vegetation drag and has little or reasonable to expect that the bed stress will play a contributing
no link to the bed stress (e.g. Nepf 1999). If turbulence has any role. Therefore, it is useful to consider methods for estimating
role to play in sediment transport, then we cannot expect that this parameter in the field. Several methods have been developed
the relationships developed for open-channel flows will hold in and tested for open-channel flows. However, most of these do not
regions with vegetation. apply or do not work well in the presence of vegetation, because
The second problem that we face in trying to characterize sedi- the vegetation profoundly alters the vertical profiles of turbu-
ment transport within vegetation is that we lack a reliable method lence and mean flow. In the following paragraphs, we discuss
for estimating the mean bed stress within a region of vegetation. five methods.
Journal of Hydraulic Research Vol. 50, No. 3 (2012) Vegetated channels 273

First, the bed stress can be defined by the spatial average of


the viscous stress at the bed:


∂ ū
τbed = ρu∗2 = ν (19)
∂z z=0

However, to properly define ∂ ū/∂z at the bed, the measurement


of velocity must be within the laminar sub-layer. While this is
possible in a laboratory setting, it is rarely possible (or practical)
to make this fine-scale measurement in the field.
Second, for open-channel flows, the bed stress can be esti-
mated from the maximum, near-bed Reynolds stress or by
extrapolating the linear profile of Reynolds stress to the bed (e.g.
Nezu and Rodi 1986). We might adapt this method to vege-
tated regions by imposing the spatial average described above,
τbed = ρu∗2 ≈
u w max . However, in many vegetated flows, the
near-bed turbulent stress is zero or close to it (e.g. Lopez and Gar-
cia 1998, Siniscalchi et al. 2012), making this estimator difficult
to resolve in the field.
Third, turbulence in an open channel is produced by the
Figure 9 The measurements of bed stress in an array of emergent,
boundary shear, so that there is a direct link between τbed and
rigid cylinders. Friction velocity is estimated from the spatial average
near-bed turbulent kinetic energy, TKE (= 0.5(u 2 + v 2 + w 2 )). of near-bed viscous stress, as in Eq. (19). White circles indicate the data
Observations over a bare bed suggest that τbed /ρ = u∗2 ≈ 0.2 of F. Kerger (unpublished data). Black circles indicate the data of Zav-
TKE (Stapleton and Huntley 1995). Although this method has istoski (1992). (a) Ratio of TKE to bed stress. Over a bare bed, this ratio
been used to estimate τbed within regions of vegetation (e.g. Wid- is 5 (e.g. Stapleton and Huntley 1995). (b) Bed-friction velocity nor-
dows et al. 2008), it is questionable whether it is valid over malized by the bed-stress estimator given in Eq. (21). For a sufficiently
dense array (aH > 0.3), the ratio has a constant value
vegetated surfaces. Within vegetation, turbulence is produced
predominantly in the wakes of individual stems and branches
and within the shear layer at the top of submerged meadows is generally much smaller than either term on the right-hand
(Section 3). There is no physical reason that τbed and TKE should side, making this estimator prone to large errors. In addition,
be correlated, because the contribution of bed shear to turbulence the method relies on accurate estimates of frontal area (a) and
generation within the canopy is small to negligible (e.g. Nepf and the drag coefficient CD . These values are not known for many
Vivoni 2000). The lack of correlation between TKE and u∗ is sug- plant species and can be functions of velocity, as discussed in
gested by recent measurements (F. Kerger, unpublished data). Section 2.
Using a laser Doppler velocimeter (LDV) positioned to achieve A possible new estimator for the mean bed stress within veg-
high vertical resolution near the bed, the bed stress in a chan- etation is based on the following observations. If vegetation
nel with rigid emergent dowels was estimated using Eq. (19). density is high enough (ah > 0.23, or δe /h < 1), the velocity
The ratio TKE/u∗2 is plotted in Fig. 9(a). If an extension from near the bed is vertically uniform and is set by the vegeta-
open-channel conditions were valid, we should see TKE/u∗2 ≈ 5. tion drag (e.g. Liu et al. 2008). Specifically, the velocity is set
However, within the emergent arrays, TKE/u∗2 varies between 3 by a balance of vegetation drag and potential forcing, yielding

and 67, showing no clear trend with roughness density, ah. This Uv = 2gS/CD a (e.g. Nepf 2012b). In some cases, a velocity
suggests that the estimator u∗2 = 0.2 TKE is not valid within overshoot is observed near the bed, associated with the junc-
regions of vegetation. tion vortex at the stem base (Liu et al. 2008). For the purpose
Fourth, when vegetation is present, the total flow resistance of this simple analysis, we neglect this overshoot. Because the
can be partitioned between the bed stress and the vegetation drag stem turbulence has scale d, we may reasonably assume that
(e.g. Raupach 1992). Integrating the momentum equation (8) this turbulence is damped by viscous stress near the bed within
over the flow depth, we can infer the bed stress by subtracting a region z < d. This implies that the velocity deviates from its
the vegetation drag from the total potential forcing, ρgSH . For uniform value at a distance from the bed that scales with d. If the
steady, uniform flow conditions, flow conditions within this region (z < d) are laminar, then we
h can estimate Eq. (1) using the scale ∂u/∂z|z=0 ∼ Uv /d. Then,
τbed = ρu∗2 = ρgSH − z=0 12 ρn CD a
ū |
ū | dz Eq. (19) reduces to
(20)
bed stress vegetation drag

This method has been used by several authors (e.g. Larsen et al. νUv ν 2gS
u∗2 ∼ = (21)
2009). The problem with this method is that the bed stress d d CD a
274 H.M. Nepf Journal of Hydraulic Research Vol. 50, No. 3 (2012)

The scale relation given in Eq. (21) was verified with measure- is a central feature in many stream restoration and bank stabiliza-
ments collected in uniform arrays of rigid, emergent cylinders. tion efforts. For example, Bennett et al. (2002) showed that the
For simplicity, Uv is approximated by the depth-averaged veloc- introduction of emergent vegetation at fixed spatial intervals, set
ity, U . In two studies (Zavistoski 1992, F. Kerger, unpublished at the estimated equilibrium meander interval, could provoke the
data), the friction velocity was estimated from multiple vertical evolution of a straight channel towards a natural state of mean-
profiles using Eq. (1). For arrays of sufficient density (ah >≈ 0.3, dering. Similarly, Larsen and Harvey (2010) explained how feed-
Fig. 9b), a constant scale factor is suggested by the observations, backs between vegetation and sediment transport drive landscape
u∗2 = [2.0 ± 0.2]νUv /d. However, note that the data shown are evolution in the Everglades. Future research should continue to
limited to conditions with Rd < 1000 and RH < 15, 000, which explore the feedbacks between vegetation, flow, and landscape
covers only a small fraction of field conditions. evolution, because they are critical components in the design of
conservation and restoration strategies for many aquatic systems.
6.2 Future research challenges in sediment transport
7.2 Hydraulic resistance and flood management
We conclude from the above discussion that there is much work
needed to understand sediment transport within regions of veg- Vegetation was historically considered only as a source of flow
etation. We lack a reliable method for estimating the mean bed resistance and was frequently removed to reduce flooding. How-
stress and, frankly, we are not even sure that the mean bed stress ever, vegetation provides ecological services that make it an
is the sole relevant parameter (e.g. Vollmer and Kleinhans 2007). integral part of river systems. The trade-off between flood and
We must also consider the role of turbulence and, relevant to this, ecological management underlines the need for a reliable method
the turbulent structure in regions of vegetation is quite different to predict channel resistance in the presence of vegetation. The
from that over a bare bed. Furthermore, unlike open-channel con- problem is particularly pressing given that over half of the
ditions, the bed stress varies spatially at the stem scale. We need world’s major river networks are regulated to manage water
to understand how this spatial variability may impact bulk trans- resources and reduce flooding (Nilsson et al. 2005), and the
port rates and how best to parameterize it. Finally, we need to frequency and magnitude of storms are projected to increase
understand the feedbacks between vegetation, flow, and morpho- due to climate change (Oki and Kanae 2006). For many years,
dynamics, in particular, with regard to the planning of restoration researchers have focused on characterizing flow resistance in
efforts (see further discussion in Section 7). channels with uniform distributions of vegetation, emphasizing
the drag contributed at the stem and leaf scale (e.g. Kouwen and
Unny 1973). However, this approach cannot work at the reach
7 Conclusions and future directions scale, because at the reach scale, vegetation is rarely distributed
uniformly, and the scale and spatial distribution of patches have
This paper has covered a lot of ground, and still it has not touched been shown to play an important role in the setting reach-scale
on many important areas, including the interaction between flow resistance (e.g. Section 5). It is the reach-scale resistance
waves and vegetation (e.g. Kobayashi et al. 1993, Lowe et al. that is most relevant for flood and watershed management. To
2005, Bradley and Houser 2009); the impact of vegetation on properly address the reach-scale flow resistance, we should focus
mass transport (e.g. Ghisalberti and Nepf 2005, Sukhodolova our efforts in two key areas. First, we need to develop and val-
et al. 2006, Murphy et al. 2007); and the dispersion and capture idate methods to rapidly characterize the spatial heterogeneity
of pollen and seeds (e.g. Ackerman 1997, Chambert and James of vegetation at the reach scale. As noted by Folkard (2010),
2009, Defina and Peruzzo 2010). Indeed, the volume of research some promising methods are emerging within the fields of air-
in vegetation hydrodynamics has exploded in recent years, as borne remote sensing (e.g. Mertes 2002); light detection and
we realize that many environmental functions are influenced by ranging (LIDAR) imaging (e.g. Heritage and Milan 2009); and
vegetation. In conclusion, I will note three areas in which veg- other high-resolution optical methods (Feurer et al. 2008). Sec-
etation hydrodynamics plays an important role: environmental ond, we need to understand what scale of morphologic detail is
restoration, resource management, and carbon cycling. relevant in the characterization of flow resistance. Recent stud-
ies point to spatial distribution at the patch scale, characterized
by the blockage factor, as the key geometric element in char-
7.1 Restoration
acterizing drag at the reach scale (Section 5). But we do not
River and stream restoration seeks to return ecological func- know the scale at which patch distributions must be resolved.
tion and biodiversity to channels by stabilizing stream banks, In other words, how sensitive is the prediction of the reach-
improving water quality, and restoring in-stream habitat. In the scale flow and resistance to the resolution at which vegetation
USA alone, over $1 billion per year is spent on river restoration distribution is described? Similarly, when are the gaps between
projects (Bernhardt et al. 2005). Studies of previous restoration patches sufficiently wide to producing channelling flow? What
efforts point to the need for collaboration between disciplines to scale of channel must be resolved to properly model the circula-
design sustainable projects (e.g. Wohl et al. 2005), and vegetation tion within a marsh (e.g. Lightbody et al. 2008)? These questions
Journal of Hydraulic Research Vol. 50, No. 3 (2012) Vegetated channels 275

could be addressed through numerical experiments that exam- C = concentration


ine the impact of vegetation spatial scale on mean flow. Finally, C = Cauchy number, ratio of drag force to
because reconfiguration impacts the meadow height, and thus the restoring force due to rigidity
blockage factor, we must understand what level of morphologi- CD = drag coefficient
cal detail is needed to properly predict reconfiguration, which in C∗ = coefficient describing the interfacial shear
turn will require more detailed measurements of plant material between vegetated and unvegetated flow
density and rigidity. regions
D = diameter of a circular patch of vegetation
d = stem diameter or blade width
7.3 Blue carbon Dm = molecular diffusivity
E = elastic modulus
Salt marshes, mangrove forests, and seagrass meadows cover
g = gravitational acceleration
less than 0.5% of the seabed, but account for 50–70% of the car- H = water depth
bon storage in ocean sediments (Nellemann et al. 2009). How H = canopy or meadow height
will the size of these habitats, and their potential for carbon stor- I = second moment of area
age, change with sea level rise, with changes in coastal land | = length scale of stem-generated turbulence
use, and with changes in sediment supply? Should we intention- l = length of a blade
ally build more marsh, mangrove, and seagrass habitats? The le = effective length of a blade
answer to these questions will require knowledge of vegetation Lc = canopy-drag length scale
hydrodynamics. For example, the potential carbon capture within L1 = length of near wake, distance from patch to
a seagrass meadow depends on the photosynthetic rate, which initiation of the von Karman vortex street
ṁ = mass flux
in turn depends on the blade-scale hydrodynamics (which sets
n = porosity of a canopy
nutrient flux) and blade/meadow scale reconfiguration (which
n⊥ = vector normal to surface
influences light availability). The potential to build new marshes R = Reynolds number
will depend on our understanding of the feedback between S = bed or water surface slope
vegetation, flow, and sediment dynamics discussed in Section 6. S = Schmidt number = Dm /ν
In conclusion, the proper management of many aquatic sys- t = blade thickness
tems depends on our understanding of the impact of vegetation U = characteristic channel velocity
on flow at different scales (blade, canopy, patch, and chan- U0 = depth-averaged upstream velocity
nel reach), which in turn impact the processes that establish (Figs. 3 and 5)
and maintain the ecosystem. Through collaborations in ecology, U1 = velocity within or just behind a region of
biology, geomorphology, and geochemistry, the field of envi- vegetation (Figs. 3 and 5)
ronmental hydraulics can answer many important questions in U2 = velocity adjacent to a patch of vegetation
environmental management. (Figs. 3 and 5)
ub∗ = shear velocity on blade surface
u∗ = bed-shear velocity = (τbed /ρ)1/2
Acknowledgements (u, v, w) = streamwise, lateral, and vertical velocity
components, respectively
Some of this material is based upon the work supported by the US (x, y, z) = streamwise, lateral, and vertical coordinate
National Science Foundation under Grant Nos. EAR0309188, directions, respectively
EAR 0125056, EAR 0738352, and OCE 0751358. Any opinions, w = width of a vegetation patch
conclusions, or recommendations expressed in this material are W = channel width
those of the author and do not necessarily reflect the views of the S = mean spacing between stems
δ = boundary-layer thickness
US National Science Foundation. The author thanks her students;
δe = length scale of vortex penetration for a
A. Lightbody, M. Ghisalberti, B. White, Y. Tanino, M. Luhar, L.
submerged canopy
Zong, J. Rominger, Z. Chen.
δL = length scale of vortex penetration at the
lateral edge of an emergent canopy
Notation δs = laminar sub-layer thickness
δC = concentration diffusive sub-layer
A = Unit surface area γ = Vogel exponent
a = frontal area per volume φ = solid volume fraction of a canopy
B = buoyancy parameter is the ratio of restoring l = roughness density = ah
forces due to rigidity and buoyancy ρ = density of water
b = half-width of a rectangular patch of vegetation ρv = density of plant material
Bx = blockage factor, fraction of channel cross-section ν = molecular viscosity
occupied by vegetation τbed = shear stress acting on the bed
276 H.M. Nepf Journal of Hydraulic Research Vol. 50, No. 3 (2012)

References environments. J. Geophys. Res. 114, F01004, 13 pp.,


doi:10.1029/ 2007JF000951.
Abt, S., Clary, W., Thornton, C. (1994). Sediment deposition and Castro, I.P. (1971). Wake characteristics of two-dimensional per-
entrapment in vegetated streambeds. J. Irrig. Drain E 120(6), forated plates normal to an air-stream. J. Fluid Mech. 46,
1098–1110. 599–609.
Ackerman, J. (1997). Submarine pollination in the marine CBEC, Chesapeake Bay Executive Council (2003). Expanded
angiosperm Zostera marina. Am. J. Bot. 84(8), 1110– Riparian forest buffer goals. Directive 03-01. Annapolis,
1119. Maryland 21403.
Afzalimehr, H., Dey, S. (2009). Influence of bank vegetation and Celik, A., Diplas, P., Dancey, C., Valyrakis, M. (2010). Impulse
gravel bed on velocity and Reynolds stress distributions. Int. and particle dislodgement under turbulent flow conditions.
J. Sediment Res. 24(2), 236–246. Phys. Fluids 22, 046601. doi: 10.1063/1.3385433.
Albayrak, I., Nikora, V., Miler, O., O’Hare, M. (2011). Flow- Chambert, S., James, C. (2009). Sorting of seeds by hydrochory.
plant interactions at a leaf scale: Effects of leaf shape, ser- River Res. Applic. 25, 4861.
ration, roughness and flexural ridity. Aquatic Sci. 73(1). doi: Chen, Z., Ortiz, A., Zong, L., Nepf, H. (2012). The wake structure
10.1007/s00027-011-0220-9. behind a porous obstruction with implications for deposition
Alben, S., Shelley, M., Zhang, J. (2002). Drag reduction near a finite patch of emergent vegetation. In revision for Water
through self-similar bending of a flexible body. Nature 420, Resour Res.
479–481. Cheng, N. (2011). Representative roughness height of sub-
Bal, K., Struyf, E., Vereecken, H., Viaene, P., De Doncker, L., merged vegetation. Water Resour. Res. 47, W08517, 18 pp.,
de Deckere, E., Mostaert, F., Meire, P. (2011). How do macro- doi: 10.1029/2011WR010590.
phyte distribution patterns affect hydraulic resistances? Ecol. Clarke, S. (2002). Vegetation growth in rivers: Influences upon
Eng. 37(3), 529–533. sediment and nutrient dynamics. Prog. Phys. Geog. 26(2),
Baptiste, M., Babovic, V., Uthurburu, J.R., Keijzer, M., Uitten- 159–172.
bogaard, R.E., Mynett, A., Verwey, A. (2007). On inducing Costanza, R., d’Arge, R., de Groot, R., Farber, S., Grasso, M.,
equations for vegetation resistance. J. Hydraulic Res. 45(4), Hannon, B., Limburg, K., Naeem, S., O’Neill, R.V., Paruelo,
435–450. J., Raskin, R.G., Sutton, P., van den Belt, M. (1997). The value
Belcher, S., Jerram, N., Hunt, J. (2003). Adjustment of a turbulent of the world’s ecosystem services and natural capital. Nature
boundary layer to a canopy of roughness elements. J. Fluid 387, 253–260.
Mech. 488, 369–398. Cotton, J., Wharton, G., Bass, J., Heppell, C., Wotton, R.
Bennett, S., Pirim, T., Barkdoll, B. (2002). Using simulated emer- (2006). The effects of seasonal changes to in-stream vegeta-
gent vegetation to alter stream flow direction within a straight tion cover on patterns of flow and accumulation of sediment.
experimental channel. Geomorphology 44, 115–126. Geomorphology 77, 320–334.
Bennett, S., Wu, W., Alonso, C., Wang, S. (2008). Modeling flu- Defina, A., Peruzzo, P. (2010). Floating particle trapping and
vial response to in-stream woody vegetation: Implications for diffusion in vegetated open channel flow. Water Resour. Res.
stream corridor restoration. Earth Surf. Process. Landforms 46, W11525, doi:10.1029/2010WR009353.
33, 890–909. Dijkstra, J., Uittenbogaard, R. (2010). Modeling the interaction
Bernhardt, E., Palmer, M., Allan, J., and the National between flow and highly flexible aquatic vegetation. Water
River Restoration Science Synthesis Working Group (2005). Resour. Res. 46, W12547, doi:10.1029/2010WR009246.
Restoration of U.S. rivers: A national synthesis. Science 308, Escauriaza, C., Sotiropoulos, F. (2011). Initial stages of
636–637. erosion and bed form development in a turbulent flow
Boelscher, J., Schulte, A., Huppmann, O. (2010). Long-term around a cylindrical pier. J. Geophys. Res. 116, F03007,
flow field monitoring at the Upper Rhine floodplains. In River doi:10.1029/2010JF001749.
flow 2010, 477–485, A. Dittrich, et al., ed. Bundesanstalt fr Feurer, D., Bailly, J.-P., Puech, C., Le Coarer, Y., Viau, A. (2008).
Wasserbau, Braunschweig, Germany. Very high-resolution mapping of river-immersed topography
Boudreau, B., Jorgensen, B. (2001). The benthic boundary layer: by remote sensing. Prog. Phys. Geog. 32(4), 403–419.
Transport and biogeochemistry. Oxford: Oxford University Finnigan, J., Shaw, R., Patton, E. (2009). Turbulence struc-
Press. ture above a vegetation canopy. J. Fluid Mech. 637,
Bouma, T., van Duren, L., Temmerman, S., Claverie, T., Blanco- 387–424.
Garcia, A., Ysebaert, T., Herman, P. (2007). Spatial flow and Folkard, A. (2010). Vegetated flows in their environmental con-
sedimentation patterns within patches of epibenthic structures: text: A review. Proc. Inst. Civil Eng. – Eng. Comp. Mech.
Combining field, flume and modelling experiments. Cont. 164(1), 3–24.
Shelf Res. 27, 1020–1045. Folkard, A. (2011). Flow regimes in gaps within stands of flexible
Bradley, K., Houser, C. (2009). Relative velocity of seagrass vegetation: Laboratory flume simulations, Env. Fluid Mech.
blades: Implications for wave attenuation in low-energy 11, 289–386.
Journal of Hydraulic Research Vol. 50, No. 3 (2012) Vegetated channels 277

Gaylord, B., Reed, D., Washburn, L., Raimondi, P. (2004). Koch, E. (2001). Beyond light: Physical, geological, and geo-
Physical-biological coupling in spore dispersal of kelp forest chemical parameters as possible submersed aquatic vegetation
macroalgae. J. Mar. Syst. 49, 19–39. habitat requirements. Estuaries 24(1), 1–17.
Ghisalberti, M., Nepf, H. (2002). Mixing layers and coher- Kouwen, N., Unny, T. (1973). Flexible roughness in open
ent structures in vegetated aquatic flow. J. Geophys. Res. channels. J. Hydrology 99(HY5), 713–728.
107(C2), doi: 10.1029/2001JC000871. Lacy, J., Wyllie-Echeverria, S. (2011). The influence of cur-
Ghisalberti, M., Nepf, H. (2004). The limited growth of rent speed and vegetation density on flow structure in two
vegetated shear-layers. Water Resour. Res. 40, W07502, macrotidal eelgrass canopies. Limnol Oceanogr. Fluids Envi-
doi:10.1029/2003WR002776. ron. 1, 38–55.
Ghisalberti, M., Nepf, H. (2005). Mass transfer in vegetated shear de Langre, E. (2008). Effects of wind on plants. Ann. Rev. Fluid
flows. Env. Fluid Mech. 5(6): 527–551. Mech. 40, 141–168.
Ghisalberti, M., Nepf, H. (2009). Shallow flows over a permeable Larsen, L., Harvey, J. (2010). How vegetation and sediment
medium: The hydrodynamics of submerged aquatic canopies. transport feedbacks drive landscape change in the Everglades.
Transport. Porous Med. 78, 385–402. Am. Nat. 176, E66–E79.
Green, E.P., Short, F.T. (2003). World atlas of seagrasses, Larsen, L., Harvey, J. (2011). Modeling of hydroecological
University California Press, Berkeley. feedbacks predicts distinct classes of landscape pattern, pro-
Green, J. (2005a). Further comment on drag and reconfiguration cess, and restoration potential in shallow aquatic ecosystems.
of macrophytes. Freshwater Biol. 50, 2162–2166. Geomorphology 126, 279–296.
Green, J. (2005b). Comparison of blockage factors in modelling Larsen, L., Harvey, J., Crimaldi, J. (2009). Predicting bed shear
the resistance of channels containing submerged macrophytes. stress and its role in sediment dynamics and restoration poten-
River Res. Applic. 21, 671–686. tial of the Everglades and other vegetated flow systems. Ecol.
Heritage, G., Milan, D. (2009). Terrestrial laser scanning of grain Eng. 35, 1773–1785.
roughness in a gravel-bed river. Geomorphology 113(2), 1–11. Lawson, S., McGlathery, K., Wiberg, P. (2012). Enhance-
Huang, I., Rominger, J., Nepf, H. (2011). The motion of kelp ment of sediment suspension and nutrient flux by benthic
blades and the surface renewal model. Limnol. Oceanogr. macrophytes at low biomass. Mar. Ecol. Prog. Ser. 448,
56(4), 1453–1462. 259–270.
Hurd, C., Stevens, C., Laval, B., Lawrence, G., Harrison, P. Lightbody, A., Avener, M., Nepf. H. (2008). Observations of
(1997). Visualization of seawater flow around morphologi- short-circuiting flow paths within a constructed treatment wet-
cally distinct forms of the giant kelp Macrocystis inregrifolia land in Augusta, Georgia, USA. Limnol. Oceanogr. 53(3),
from wave-sheltered and exposed sites. Limnol. Oceanogr. 1040–1053.
41(1), 156–163. Liu, D., Diplas, P., Fairbanks, J., Hodges, C. (2008). An exper-
Huthoff, F., Augustijn, D., Hulscher, S. (2007). Analytical imental study of flow through rigid vegetation. J. Geophys.
solution of the depthaveraged flow velocity in case of sub- Res. 113, F04015, doi:10.1029/2008JF001042.
merged rigid cylindrical vegetation. Water Resour. Res. 43(6), Lopez, F., Garcia, M. (1998). Open-channel flow through sim-
W06413, doi:10.1029/2006WR005625. ulated vegetation: Suspended sediment transport modeling.
Julien, P. (2010). Erosion and sedimentation. ed. 2. Cambridge Water Resour. Res. 34(9), 2341–2352.
University Press, Cambridge, UK. Lopez, F., Garcia, M. (2001). Mean flow and turbulence struc-
Kaimal, J., Finnigan, J. (1994). Atmospheric boundary layer ture of open-channel flow through non-emergent vegetation.
flows: Their structure and measurement. Oxford University J. Hydraulic Res. 127, 392–402.
Press, Oxford, UK. Lowe, R., Koseff, J., Monismith, S. (2005). Oscillatory flow
van Katwijk, M., Bos, A., Hermus, D., Suykerbuyk, W. (2010). through submerged canopies: 1. Velocity structure. J. Geo-
Sediment modification by seagrass beds: Muddification and phys. Res. 110(C10016). doi:10.1029/2004JC002788.
sandification induced by plant cover and environmental con- Luhar, M., Rominger, J., Nepf, H. (2008). Interaction between
ditions. Estuar. Coast. Shelf Sci, doi:10.1016/j.ecss.2010. flow, transport and vegetation spatial structure. Environ. Fluid
06.008. Mech. 8(5–6), 423–439.
Kays, W., Crawford, M. (1993). Convective heat and mass Luhar, M., Nepf. H. (2011). Flow induced reconfiguration of
transfer. ed. 3. McGraw-Hill, New York. buoyant and flexible aquatic vegetation. Limnol. Oceanogr.
Kemp, J., Harper, D., Crosa, G. (2000). The habitat-scale 56(6), 2003–2017.
ecohydraulics of rivers. Ecol. Eng. 16, 17–29. Luhar, M., Nepf, H. (2012). From the blade scale to the reach
Kobayashi, N., Raichle, A., Asano, T. (1993). Wave attenuation scale: A characterization of aquatic vegetative drag. Adv.
by vegetation. J. Waterw. Port Coast. Ocean Eng. 119, 30–48. Water Resour. (In Press).
Koch, E. (1994). Hydrodynamics, diffusion-boundary layers and Mars, M., Kuruvilla, M., Goen, H. (1999). The role of submer-
photosynthesis of the seagrasses, Thalassia testudinum and gent macrophyte triglochin huegelii in domestic greywater
Cymodocea nodosa. Mar. Biol. 118, 767–776. treatment. Ecol. Eng. 12, 57–66.
278 H.M. Nepf Journal of Hydraulic Research Vol. 50, No. 3 (2012)

Mertes, L. (2002). Remote sensing of riverine landscapes. Implications for sediment transport. J. Fluid Mech. 326,
Freshwater Biol. 47, 799–816. 285–319.
Moore, KA. (2004). Influence of seagrasses on water quality in Oki, T., Kanae, S. (2006). Global hydrological cycles and world
shallow regions of the lower Chesapeake Bay. J. Coast. Res. water resources. Science 313, 1068–1072.
20(Special Issue), 162–178. Plew, D., Cooper, G., Callaghan, F. (2008). Turbulence induced
Murphy, E., Ghisalberti, M., Nepf, H. (2007). Model and forces in a freshwater macrophyte canopy. Water Resour. Res.
laboratory study of dispersion in flows with submerged 44, W02414, doi:10.1029/2007WR006064.
vegetation. Water Resour. Res. 43, W05438, doi:10.1029/ Poggi, D., Katul, G., Albertson, J. (2004). A note on the con-
2006WR005229. tribution of dispersive fluxes to momentum transfer within
National Research Council (2002). Riparian areas: Functions canopies. Bound.-Lay. Meteorol. 111, 615–621.
and strategies for management. National Academy Press, Poggi, D., Krug, C., Katul, G. (2009). Hydraulic resis-
Washington, DC. tance of submerged rigid vegetation derived from first-
Nellemann, C., Corcoran, E., Duarte, C., Valdés, L., De Young, order closure models. Water Resour. Res. 45, W10442,
C., Fonseca, L., Grimsditch, G. (eds.) (2009). Blue carbon. doi:10.1029/2008WR007373.
A rapid response assessment. United Nations Environment Pollen, N., Simon, A. (2005). Estimating the mechanical
Programme, Arendal, Norway. effects of riparian vegetation on stream bank stability using
Nepf, H. (1999). Drag, turbulence, and diffusion in flow through a fiber bundle model. Water Resour. Res. 41, W07025,
emergent vegetation. Water Resour. Res. 35, 479–489. doi:10.1029/2004WR003801.
Nepf, H., Vivoni, E. (2000). Flow structure in depth-limited, Pollen-Bankhead, N., Simon, A. (2010). Hydrologic and
vegetated flow. J. Geophys. Res. 105(28), 547–557. hydraulic effects of riparian root networks on streambank
Nepf, H., Ghisalberti, M., White, B., Murphy, E. (2007). stability: Is mechanical root-reinforcement the whole story?
Retention time and dispersion associated with sub- Geomorphology 116(3–4), 353–362.
merged aquatic canopies. Water Resour. Res. 43, W04422, Raupach, M. (1992). Drag and drag partition on rough surfaces.
doi:10.1029/2006WR005362. Bound.-Lay. Meteorol. 60(4), 375–395.
Nepf, H. (2012a). Flow over and through biota. In Treatise on Raupach, M., Finnigan, J., Brunet, Y. (1996). Coherent eddies
estuarine and coastal science. E. Wolanski, D. McLusky, eds. and turbulence in vegetation canopies: The mixing-layer
Vol. 2, Elsevier Inc., San Diego, CA, pp. 267–288. analogy. Bound.-Lay. Meteorol. 60, 375–395.
Nepf, H. (2012b). Flow and transport in regions with aquatic Ree, W.O. (1949). Hydraulic characteristics of vegetation for
vegetation. Ann. Rev. Fluid Mech. 44, 123–142. vegetated waterways. Agric. Eng. 30, 184–189.
Nezu, I., Rodi, W. (1986). Open-channel flow measurements with Righetti, M. (2008). Flow analysis in a channel with flexible
a laser Doppler anemometer. J. Hydraulic Eng. 112(5), 335– vegetation using double-averaging method. Acta Geophysica
355. 56(3), 801–823.
Nicolle, A., Eames, I. (2011). Numerical study of flow through Rominger, J., Lightbody, A., Nepf, H. (2010). Effects of added
and around a circular array of cylinders. J. Fluid Mech. 679, vegetation on sand bar stability and stream hydrodynamics. J.
1–31. Hydraulic Eng. 136(12), 994–1002.
Nikora, N., Nikora, V. (2007). A viscous drag concept for flow Rominger, J., Nepf, H. (2011). Flow adjustment and interior
resistance in vegetated channels [CD-ROM]. Proc. 32nd IAHR flow associated with a rectangular porous obstruction. J. Fluid
Congress, Venice, 1–6 July. Mech. 680, 636–659.
Nikora, V., McEwan, I., McLean, S., Coleman, S., Pokrajac, D., Sand-Jensen, K., Madsen, T. (1992). Patch dynamics of the
Walters, R. (2007). Double-averaging concept for rough-bed stream macrophyte, Callitriche cophocarpa. Freshwater Biol.
open-channel and overland flows: Theoretical background. J. 27, 277–282.
Hydraulic Eng. 133(8), 873–883. Sand-Jensen, K. (2003). Drag and reconfiguration of freshwater
Nikora, V., Larned, S., Nikora, N., Debnath, K., Cooper, G., Reid, macrophytes. Freshwater Biol. 48, 271–283.
M. (2008). Hydraulic resistance due to aquatic vegetation in Shimizu, Y., Tsujimoto, T. (1994). Numerical analysis of tur-
small streams: Field study. J. Hydraulic Eng. 134(9), 1326– bulent open-channel flow over a vegetation layer using a k-e
1332. turbulence model. J. Hydrosci. Hydraulic Eng. 11, 57–67.
Nikora, V. (2010). Hydrodynamics of aquatic ecosystems: An Siniscalchi, F., Nikora, V., Aberle, J. (2012). Plant patch hydro-
interface between ecology, biomechanics and environmental dynamics in streams: Mean flow, turbulence, and drag forces.
fluid mechanics. River Res. Applic. 26, 367–384. Water Resour. Res. 48, W01512.
Nilsson, C., Reidy, C., Dynesius, M., Revenga, C. (2005). Frag- Stapleton, K., Huntley, D. (1995). Seabed stress determination
mentation and flow regulation of the world’s large river using the inertial dissipation method and turbulent kinetic
systems. Science 308, 405–408. energy method. Earth Surf. Process. Landforms 20, 807–815.
Nino, Y., Garcia, M. (1996). Experiments on particle-turbulence Statzner, B., Lamouroux, N., Nikora, V., Sagnes, P. (2006).
interactions in the near-wall region of an open channel flow: The debate about drag and reconfiguration of freshwater
Journal of Hydraulic Research Vol. 50, No. 3 (2012) Vegetated channels 279

macrophytes: Comparing results obtained by three recently channel. Water Resour. Res. 44(1), W01412, doi:10.1029/
discussed approaches. Freshwater Biol. 51, 2173–2183. 2006WR005651.
Stevens, C., Hurd, C., Isachsen. P. (2003). Modelling of diffusion White, F. (2008). Fluid mechanics. ed. 6. McGraw Hill, Boston,
boundary-layers in subtidal macrogalgal canopies: Response MA.
to waves and currents. Aquat. Sci. 65, 81–91. Widdows, J., Pope, N., Brinsley, M. (2008). Effect of Spartina
Sukhodolov, A. (2005). Comment on drag and reconfiguration anglica stems on near-bed hydrodynamics, sediment erodabil-
of macrophytes. Freshwater Biol. 50, 194–195. ity and morphological changes on an intertidal mudflat. Mar.
Sukhodolov, A., Sukhodolova, T. (2010). Case study: Effect of Ecol. Prog. Ser. 362, 45–57.
submerged aquatic plants on turbulence structure in a lowland Wilcock, R., Champion, P., Nagels, J., Crocker, G. (1999).
river. J. Hydraulic Eng. 136(7): 434–446. The influence of aquatic macrophytes on the hydraulic and
Sukhodolova, T., Sukhodolov, A., Kozerski, H., Köhler, J. physicochemical properties of a New Zealand lowland stream.
(2006). Longitudinal dispersion in a lowland river with Hydrobiologia 416(1), 203–214.
submersed vegetation. In: River flow 2006, R.M.L. Fer- Wilson, R., Shaw, R. (1977). A higher order closure
reira, E.C.T.L. Alves, J.G.A.B. Leal and A.H. Cardoso (eds) model for canopy flow. J. Applic. Met. 116(11), 1197–
Proc. of 3rd Int. Conference on Fluvial Hydraulics. Septem- 1205.
ber 6–8, 2006, Lisbon, Portugal, V1, pp. 631–637, ISBN: Wohl, E., Angermeier, P., Bledsoe, B., Kondolf, G., Mac-
978-0-415-40815-8. Donnell, L., Merritt, D., Palmer, M., Poff, N., Tarboton, D.
Tal, M., Paola, C. (2007). Dynamic single-thread channels main- (2005). River restoration. Water Resour. Res. 41, W10301,
tained by the interaction of flow and vegetation. Geol. Soc. Am. doi:10.1029/2005WR003985.
35, 347–350. Wooding, R., Bradley, E., Marshall, J. (1973). Drag due to regular
Tanino, Y., Nepf, H. (2008). Lateral dispersion in random cylin- arrays of roughness elements. Bound.-Lay. Meteorol. 5, 285–
der arrays at high Reynolds number. J. Fluid Mech. 600, 308.
339–371. Wu, F., Shen, H., Chou, Y. (1999). Variation of roughness
Tsujimoto, T. (1999). Fluvial processes in streams with vegeta- coefficients for unsubmerged and submerged vegetation.
tiona. J. Hydr. Res. 37(6), 789–803. J. Hydraulic Eng. 125(9), 934–942.
US Environmental Protection Agency (2000). Principles for the Zavistoski, R. (1992). Hydrodynamic effects of surface pierc-
ecological restoration of aquatic resources, Report EPA841- ing plants. SM Thesis, Massachusetts Institute of Technology,
F-00- 003, Office of Water (4501F), Washington, DC. Cambridge.
Vogel, S. (1994). Life in moving fluid. ed. 2. Princeton University Zimmerman, R. (2003). A biooptical model of irradiance dis-
Press, Princeton, NJ. tribution and photosynthesis in seagrass canopies. Limnol
Vollmer, S., Kleinhans, G. (2007). Predicting incipient motion, Oceanogr. 48(1), 568–585.
including the effect of turbulent pressure fluctuations in Zong, L., Nepf, H. (2010). Flow and deposition in and around a
the bed. Water Resour. Res. 43, W05410, doi:10.1029/ finite patch of vegetation. Geomorphology 116, 363–372.
2006WR004919. Zong, L., Nepf, H. (2011). Spatial distribution of deposition
White, B., Nepf, H. (2007). Shear instability and coherent struc- within a patch of vegetation. Water Resour. Res. 47, W03516,
tures in a flow adjacent to a porous layer. J. Fluid Mech. 593, doi:10.1029/2010WR009516.
1–32. Zong, L., Nepf, H. (2012). Vortex development behind a
White, B., Nepf, H. (2008). A vortex-based model of veloc- finite porous obstruction in a channel. J. Fluid Mech. 691,
ity and shear stress in a partially vegetated shallow 368–391.

You might also like