You are on page 1of 11

Analytical solution of the depth-averaged flow velocity in case

of submerged rigid cylindrical vegetation

F. Huthoff1,2 , D.C.M Augustijn1 & S.J.M.H. Hulscher1


1
Water Engineering and Management, University of Twente, The Netherlands
2
HKV Consultants, Lelystad, the Netherlands

ABSTRACT: A new model for the depth-averaged velocity for flow in presence of submerged vegetation
is developed. The model is based on a two-layer approach, where flow above and through the vegetation
layer is described separately. Vegetation is treated as a homogeneous field of identical cylindrical stems
and the flow field is considered stationary and uniform. It is demonstrated that scaling considerations
of the bulk flow field can be used to avoid complications associated with smaller scale flow processes,
and that still the behavior of depth-averaged flow over vegetation is described accurately. The derived
scaling expression of the average flow field is simple in form, it follows fundamental laws of fluid flow
and it shows very good agreement with laboratory flume experiments. The new model can be used for
quick evaluation of a rivers hydraulic response in cases where vegetated floodplains are inundated.

Subject headings: Vegetation, Hydraulic roughness, Flow resistance, Hydraulic models.

1 Introduction if the irregularities, or resistance elements are of


the same order of magnitude as the flow depth,
available methods to predict flow resistance are no
The presence of vegetation in floodplains may
longer valid (e.g. Smart et al. 2002, Stone and Shen
have significant influence on the overall discharge
2002). If, for example, a flow field is penetrated by
capacity of a river (e.g. Darby 1999). In par-
vegetation, turbulent vortices are created in the
ticular if floodplains are relatively wide com-
wakes downstream of the protruding stems (e.g.
pared to the main channel, realistic predictions of
Akilli and Rockwell 2002). The associated energy
stage-discharge relations rely strongly on accurate
losses of the mean flow field cause the flow to slow
knowledge of floodplain flow. Therefore, it is cru-
down. These drag effects can even become more
cial to understand the processes that contribute
important than energy losses due to friction at
to floodplain resistance, and the hydraulic impacts
the channel bed (e.g. James et al. 2004). There-
these processes may have.
fore, conventional resistance equations (based on
The interplay between walls or solid objects and wall roughness) are no longer appropriate when
the flow field causes vortices and swirling motions describing flow through vegetation. In such cases,
at various length scales. Trying to represent these both bed resistance and drag effects have to be
detailed flow characteristics is an immense task, taken into account in order to arrive at a represen-
which even with the most modern equipment is tation of hydraulic resistance that covers a wide
impossible for spatial and temporal scales that are range of conditions realistically.
relevant in river engineering studies. Consequently,
depth-averaged quantities and scaling considera- Several previous investigations have focussed on
tions are often used to arrive at bulk flow descrip- effects of vegetative resistance in river flows. Ex-
tions. Several such relations exist that relate the perimental campaigns have been set up to mea-
average flow velocity to surface characteristics, for sure flow resistance in natural vegetated fields
example by means of a roughness height (Strickler (see e.g. Green 2006). Experiments in labora-
1923, Nikuradse 1933, Keulegan 1938). However, tory flumes have been carried out, with some re-

1
cent studies by Armanini et al. (2005), Jarvela
(2002), Wilson et al. (2003), and Stephan and gravitation
Gutknecht (2002)). Also, detailed numerical simu- h-k Us surface layer
lations of flow through vegetation were performed interface shear
(e.g. Shimizu and Tsujimoto 1994, Erduran and
Kutija 2003, Neary 2003, Choi and Kang 2004). As
a result of such studies, several relations were pro- gravitation
posed that describe flow resistance caused by veg-
etation. Some of these relations are empirical (Ree
k Ur drag resistance layer
and Crow 1977, Kouwen and Fathi-Moghadam
2000), others have stronger theoretical founda- bed
resistance
tions (Petryk and Bosmajian 1975, Stone and Shen
2002). Also, there are methods that are based on Figure 1. In the two-layer approach for flow with
modified logarithmic-velocity profiles (e.g. Kouwen submerged vegetation, the height of the vegetation k
and Unny 1973, Stephan and Gutknecht 2002), marks the interface between the resistance layer and
which are a well-established experimental charac- the surface layer (h is the total flow depth). The av-
teristic of turbulent wall-bounded flows and find erage velocity in the resistance layer Ur is determined
theoretical justification in similarity arguments by a balance between the streamwise component of
(e.g. Schlichting 1979). Klopstra et al. (1997) com- the gravitional force, bed resistance, form drag and
bined methodologies by treating flow over vegeta- shear stress due to flow over the vegetation. Flow in
tion in a two-layer approach, where the flow in the surface layer (Us ) is decribed with an equivalent
two layers is described separately. They combined bed roughness model (see also Figure 3).
a modified logarithmic velocity profile in the sur-
face layer with a newly derived velocity profile that In section 2 the concept of the two-layer ap-
is present in between the vegetation. proach is explained. Next, in section 3 and 4 the
Among the existing vegetation-resistance rela- proposed velocity descriptions for the two flow
tions, the empirical ones have the advantage of be- layers are derived. A comparison with selected
ing simple in form. On the other hand, empirical data from flume experiments is made in section
relations have the drawback that their applicabil- 5, which subsequently yields the new calibrated
ity is limited to the range of conditions for which model for the depth-averaged velocity of flow with
they were derived. Theoretical descriptions often submerged vegetation.
are complex. Besides, they may require poorly un-
derstood closure parameters and sometimes pose
practical difficulties when gathering required input 2 The two-layer approach
data.
In the current work a new scaling expression In a flow field over a solid boundary the force bal-
of the average flow field is proposed that com- ance between gravitation and flow resistance leads
bines advantages of existing flow equations: (i) it to an equilibrium flow state, which can be asso-
is simple in form, (ii ) it is based on fundamental ciated with a characteristic average flow velocity.
(flow) principles, (iii) it depends on readily mea- For steady uniform flow this condition is expressed
surable quantities and, most importantly, shows as
excellent agreement with experimental data. The ghi = f U 2 (1)
| {z } | {z }
method is based on a two-layer approach, simi- gravitation bed resistance
lar as done by Klopstra et al. (1997), where flow
in and above the vegetation is treated separately. where the streamwise component of the gravita-
However, here we only consider bulk characteris- tional force (per unit length and width) depends
tics of the flow field by which we avoid difficulties on the density of water , the gravitational ac-
associated with depth-averaging of flow velocities. celeration g, the depth of flow h and the channel
Vegetation is replaced by cylindrical stems that slope i. Dimensional analysis requires that for a
are homogeneously distributed and have identical dimensionless bed-friction function f the average
geometrical and material properties. Furthermore, flow velocity enters the bed resistance term as U 2 .
the vegetated channel is considered to be infinitely For flow over a solid boundary several formulations
wide, such that side-wall effects can be neglected. for f have been proposed to reflect energy losses
An improved understanding of processes in these due to wall friction (see Yen 2002 for an overview).
idealized conditions will eventually lead to a better Strickler (1923) proposed a power-law dependence
description of flow through natural vegetation. on the relative roughness height kS /h, where kS

2
reflects the height of irregularities at the bed: D s
!1/3
1 kS
fS = 64
. (2)
h
A theoretical justification for equation (2) is given gravitation
h Ur0
by Gioia and Bombardelli (2002), who have shown drag
that the given power-law dependence is valid for
turbulent flow over a hydraulically rough bed. This
bed
condition is usually met in natural rivers. resistance
However, in the situation of flow over vegetation
not only the bed friction slows down the flow but Figure 2. The depth-averaged velocity Ur0 is the re-
also the drag caused by the vegetation. Therefore, sult of the balance between the gravitational force, the
an additional resistance term v is added to the drag force and the bed resistance. Due to the inclina-
force balance from equation (1), which results in tion of the bed (with angle i, not shown here), the
driving force acts to the right. The individual cylinder
ghi = f U 2 + v . (3) stems have
a diameter D [m] and a separation s [m]
(s = 1/ m, m = surface density [m2 ]).
Equation (3) describes the force balance in a flow
layer that is penetrated by vegetation.
If the vegetation is submerged in a flow field,
3 The resistance layer
then flow above the vegetation is not direcly ob-
The resistance layer is defined as the flow layer that
structed by the individual plants. The flow in this
is penetrated by the vegetation elements (see Fig-
particular layer experiences resistance due to a
ure 1). In this section a description of the depth-
slower flowing region in between the vegetation.
averaged velocity in this layer is derived. For that
The resistance to flow in the surface layer can
purpose first the flow through emergent vegetation
therefore be described by an equivalent bed shear
is considered (section 3.1), where hydraulic effects
stress. Hence, the depth-averaged velocity in this
due to bed friction and vegetation drag are taken
surface layer determined by a force balance equiv-
into account. Next, in section 3.2 a model for the
alent to equation (1).
depth-averaged velocity in a submerged resistance
In summary, assuming that all vegetation has
layer is proposed.
the same height, for overflown vegetation two dis-
tinct flow layers are defined that have different
characteristic flow velocities (see also Figure 1): 3.1 Emergent vegetation
(i ) the flow layer that is penetrated by the vege- To derive the average flow velocity in the resistance
tation (i.e. the resistance layer ) with average layer, we first consider the situation where the veg-
velocity Ur and etation is not completely overflown (see Figure 2).
(ii ) the flow layer above the vegetation (i.e. the By replacing the vegetation with a homogeneous
surface layer ) with average velocity Us . field of identical cylindrical stems, the additional
resistance v in equation (3) can be replaced by
New depth-averaged velocity models for the resis- a standard drag force term (stem drag, see also
tance layer and the surface layer are derived in Schlichting 1979) to yield:
section 3 and section 4, respectively.
The average flow velocity of the entire flow depth ghi = f Ur0 2 + 12 CD mDhUr0 2 . (5)
(UT ) is found by proportionally adding the average
flow velocities of the resistance layer Ur and the The velocity for flow through emergent vegeta-
surface layer Us : tion is now denoted by Ur0 . Individual stems have
k hk a diameter D [m] and bed surface density m [m2 ]
UT = Ur + Us , (4) and are characterized by a dimensionless drag co-
h h
efficient CD .
where the height of the vegetation k only partly Note that the product CD mD in equation (5)
penetrates the total depth of flow h (see also Fig- carries the dimension [1/m]. For further simplifi-
ure 1). In section 5.3 the two proposed flow models cation we therefore introduce a new length param-
for the surface and the resistance layer are com- eter (the drag length)
bined in equation (4) to yield the new model for the
depth-averaged flow velocity for the total depth of 1
flow. b= . (6)
CD mD

3
Based on equation (5) the average velocity for flow After inserting Stricklers bed resistance function
through emergent vegetation Ur0 is written as fS as given in equation (2) into equation (10) the
v scaling velocity Ur0 is written as:
u
u 2bgi v
Ur0 = t 2b , for h k. (7) u
u 2bgi
1+ h
f Ur0 = u , for h k. (12)
t kS 1/3
b
1+ 32k h
The average flow velocity Ur0 can now be deter-
mined if the bed resistance function f is known, The average velocity in the resistance layer Ur
for example by using Stricklers relation as given logically reduces to Ur0 (equation (7)) if no free
by equation (2). flowing layer above the cylindrical stems is present
(i.e. if h = k).
3.2 Submerged vegetation A square root law as in equation (11) has also
been suggested in earlier works. Smart et al. (2002)
When the cylindrical elements become submerged, point out that with increasing relative roughness,
the flow in the surface layer will have a higher av- a conceptual drag coefficient model would lead to a
erage velocity as in this layer no drag due to the square-root law of the flow resistance. Baptist et al.
stems is experienced. The energy losses in the sur- (2006) used dimensionally aware genetic program-
face layer will be entirely due to a shear stress near ming (Keijzer and Babovic 2000) and also found a
the top of the resistance layer, which balances the square root dependence with respect to the relative
streamwise component of the gravitational force flow depth h/k.
that drives the flow. The shear stress between the Considering that the roughness height kS is usu-
surface layer and the resistance layer (i.e. the in- ally much smaller than the flow depth h, the con-
terface shear k , at height k) will also cause the tribution of bed resistance is often negligible. From
flow in the resistance layer to speed up (Figure 1). equation (12) it follows that the average velocity
If we introduce the extra shear stress component in the resistance layer Ur0 is practically equal to
k due to flow in the surface layer, and realize that
q
the height of the resistance layer is now given by Ur0 2bgi, if kS << h. (13)
k, the force balance from equation (5) modifies to
The average flow velocity in the resistance layer
k + gki = f Ur 2 + 12 CD mDkUr 2 . (8) can be estimated from equation (13) if there is no
The velocity in the resistance layer for submerged information available about the properties of the
vegetation is now denoted by Ur . Furthermore, the bed in between the vegetation. Such an estimate is
shear stress at the top of the resistance layer k reasonable for dense or tall emergent vegetation in
balances the gravitational force that acts on the relatively deep flows (small drag length b, and large
water volume in the surface layer, given as vegetation height k and flow depth h). For sparse
vegetation distributions (large drag lengths), bed
k = g(h k)i. (9) resistance may become the dominant source of flow
resistance. This effect is also included in equation
Inserting equation (9) into equation (8) yields the (11), which for extremely large drag length b ulti-
force balance for flow in a submerged resistance mately transforms to Mannings well-known rela-
layer: tion for flow over a rough bed.

ghi = f Ur 2 + 12 CD mDkUr 2 . (10)


4 The Surface Layer
The contribution on the left-hand-side of equation
(10) is due to the gravitational force that acts on In this section a new expression for the average
the surface and the resistance layer. The drag force flow velocity in the surface layer is derived. First,
now acts over a depth k, as opposed to the total a bulk force balance in the surface layer dictates
flow depth h in the force balance that describes the magnitude of the shear stress that is present at
flow through emergent vegetation (equation (5)). the top of the vegetation layer. In section 4.1 this
Rearranging terms in equation (10) gives a new de- shear stress is linked to characteristic velocities
scription for the average velocity in the resistance in the surface layer through scaling assumptions
layer for submerged conditions: of the Reynolds stress (from the Reynolds Aver-
s aged Navier-Stokes equations). Next, energy con-
Ur h siderations in fully developed turbulent flow (Kol-
= , for h k. (11) mogorov scaling) allow us to associate the average
Ur0 k

4
gravitation
(2002) assume that streamwise velocity fluctua-
tions vx scale with the average flow velocity Us ,
and that vertical fluctuations vz are determined
Us surface layer
by eddies between the roughness elements. It is
equiv. bed h-k assumed that a characteristic spatial scale r and
resistance
characteristic velocity ur can be associated with
k these eddies. Following Gioia and Bombardelli, the
ur , r
resistance layer
interface shear stress at the artificial bed (i.e near
the top of the resistance layer, see Figure 3) scales
Figure 3. The depth-averaged velocity in the surface as
layer Us . An equivalent bed resistance at the artificial k Us ur . (15)
bed (i.e. at the top of the resistance layer) balances
the gravitational force that acts on the surface layer.
Together with the expression for the interface
The length parameter r reflects the roughness height shear stress as given in equation (9) this yields
of the artificial rough bed.
g(h k)i Us ur . (16)
flow velocity in the surface layer to the depth of
To find a relation for the average velocity in the
the surface layer (section 4.2). Finally, in section
surface layer Us , an independent expression for ur
4.3 the velocity model is completed by subjecting it
needs to be found. For that purpose, a methodol-
to two limiting situations: (i) continuity with flow
ogy similar to the one demonstrated by Gioia and
in the resistance layer when the surface layer van-
Bombardelli (2002) is used, which is based on the
ishes and (ii ) transition to Mannings flow equa-
condition of a constant turbulent energy dissipa-
tion if the surface layer is much larger than the
tion rate from large to smaller flow scales (Kol-
vegetation height.
mogorov scaling).
The resulting velocity model carries two un-
known paramaters: (i) a scaling length that re-
flects the size of turbulent eddies at the top of the 4.2 Turbulent energy and Kol-
resistance layer and (ii) an incomplete similarity mogorov scaling
exponent that describes the transition to the ve-
locity in the resistance layer. Properties of these In the Kolmogorov view on turbulent flow, tur-
two parameters are determined empirically in sec- bulent energy is created through external forcing
tion 5 where a comparison is made with laboratory at the largest scale of the system (energy contain-
flume experiments. ing range) and is dissipated to successively smaller
scales until eventually viscosity damps the small-
est flow patterns (e.g. Pope 2000). The rate of pro-
4.1 Shear stress in the surface layer duction of turbulent kinetic energy per unit mass
is denoted by [m2 /s3 ] and, at large scales, is inde-
In steady flow over a rough bed, the shear stress pendent from the viscosity (Kolmogorovs second
at the bed surface balances the gravitational force similarity hypothesis). For the energy input at the
on the water body. Similarly, for flow over vegeta- largest scale, a scaling expression for is obtained
tion, resistance to flow in the surface layer is due that is composed of representative geometrical pa-
to a shear stress that originates at the top of the rameters and the representative flow velocity. For
resistance layer (see Figure 3). An expression for surface layer flow as depicted in Figure 3, it seems
this shear stress is already given in equation (9). natural to choose the depth of the surface layer
From the Reynolds Averaged Navier-Stokes equa- (h k) for the geometrical spatial parameter, and
tion (RANS) it follows that the (Reynolds) shear the average surface velocity Us for the representa-
stress at any location in the flow is determined by tive velocity. Dimensional analysis yields
the magnitude of fluctuations in the velocity field
(e.g. Pope 2000) Us3
. (17)
hk
xz = vx vz (14)
Furthermore, the concept of an energy cascade, in
where vx and vz denote turbulent velocity fluctua- which turbulent energy is transferred from large
tions in streamwise and vertical direction (i.e over to smaller spatial scales, requires that energy dis-
depth), respectively. sipated at small spatial scales equals production
In the theoretical derivation of Mannings law at the largest scale of the system (e.g. Pope 2000).
for rough channel flow, Gioia and Bombardelli In our particular case, if it is assumed that eddies

5
of size r and velocity ur dominate the flow field the following power law asymptotic is defined:
near the artificial rough bed (i.e. at the top of the !
vegetation), the dissipation rate scales as Us hk
. (21)
Ur0 `
u3r
. (18)
r An unknown scaling length ` is introduced in equa-
tion (21), which is determined from comparison
Now is related to the characteristic velocity ur with experimental data (see section 5).
and a length scale r, which represents the extent
If in equation (20) r indeed reflects an equivalent
of local eddies. The characteristic velocity in the
roughness height, at large depths the power law
resistance layer can be related to the average ve-
exponent in equation (21) must have the value =
locity in the surface layer as (using eqs. (17) and
2/3:
(18))
1/3
r !2/3
ur U s . (19) Us hk
hk , for h >> k. (22)
Ur0 `
Next, equation (19) is inserted into equation (16),
which yields a scaling expression for the average However, a property of equation (22) is that Us re-
velocity in the surface layer: duces to zero when the surface layer vanishes (i.e.
!1/6 q when h approaches k). This behavior is not realis-
hk tic because the average flow velocity in the surface
Us g(h k)i. (20)
r layer should always be larger than the flow veloc-
ity in the resistance layer (Ur0 ), even if the surface
Note that equation (20) reduces to the Man- layer is extremely shallow. Essentially, the power
ning/Strickler equation (Manning 1889, Strickler law in equation (22) does not give realistic results
1923) if the spacing hydraulic radius r is in- when the depth of the surface layer, hk, becomes
dependent of flow depth. It is well-established close to or smaller than the scaling length `. This
that for turbulent flow over rough-walls the re- can be solved if we assume incomplete similarity in
lation between the average flow velocity and the surface layer depth h k, and no kind of sim-
wall-characteristics is represented by Mannings ilarity in the relative depth h/k (see Barenblatt
formula (Manning 1889, Gioia and Bombardelli 2003 for theory on complete and incomplete sim-
2002). This will be used as a limiting condition, in ilarity). Assuming incomplete similarity in h k,
case the depth of the surface layer becomes much we allow the power exponent in equation (21) to
larger than the roughness elements. be variable (depending on the relative flow depth
h/k): !
4.3 Similarity considerations Us h k (h/k)
. (23)
Ur0 `
In the previous section it was shown that energy
considerations of the turbulent flow field lead to an We have already argued that for large depths the
average flow description that is consistent with the power exponent should approach 2/3, in corre-
equation proposed by Manning and Strickler. The spondence with Mannings resistance law. How-
expression derived in section 4.2, equation (20), ever, as h approaches k the velocity in the surface
is valid for deep surface layer flows (for h >> k) layer Us should approach Ur0 , which requires that
and involves an unknown scaling length r. In the vanishes. In summary:
current section we will use similarity arguments
in combination with known asymptotic behaviour if h/k >> 1 then 2/3,
of the flow, to find a flow description of in surface
if h/k 1 then 0.
layer that also describes shallow surface layer flows
realistically. These conditions are met when is defined as
The velocity in the surface layer can be treated
!
as a dimensionless quantity by scaling Us to the
2 h
characteristic velocity in the resistance layer Ur0 . = 1 , for > 0. (24)
The same scaling methodology was followed to ar- 3 k
rive at equation (11) for the dimensionless velocity
in the resistance layer. Assuming that the dimen- Equation (24) describes the transition from a
sionless velocity in the surface layer shows similar- power exponent = 2/3 to = 0 as the depth of
ity with respect to the surface layer depth (h k), the surface layer decreases (i.e. as h approaches k).

6
2
h = 1.81 m
1.8
2/3
1.6
U = 0.54 m/s
1.4 s

1.2
[]

z [m]
1 k = 0.9 m
1/3
0.8
2
0.6 m = 256 m
D = 0.008 m
=2 Ur = 0.12 m/s
0.4 CD = 0.99
=5
= 10 0.2
s = 0.055 m
0 i = 0.0009
1 2 3 4 5 6
h/k [] 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
u(z) [m/s]
Figure 4. Possible functions of the exponent in
equation (23), when assuming incomplete similarity Figure 5. Measured velocity profile (+) for one of
between the dimensionless flow velocity in the surface the 48 flow experiments with rigid cylindrical vege-
layer (Us /Ur0 ) and the depth of the surface layer hk. tation (Meijer and van Velzen 1998). Arrows depict
Larger values for result in faster transitions from the magnitude of the depth-averaged velocities in the
= 0 to = 2/3 (see equation (24)). surface layer Us and the resistance layer Ur .

The exponent determines how quickly makes of experiments was carried out with a surface den-
this transition (see Figure 4). sity m of 256 cylindrical stems per square meter,
Inserting equation (24) into the scaling equation the other half with a surface density of m = 64
for the average velocity in the surface layer, equa- m2 .
tion (23), yields For each of the experiments, flow velocities were
measured at depths 0.10 m apart with an Acous-

! 2 1( h ) tic Doppler Velocimeter (ADV). Using linear inter-
Us hk 3 k
polation, the average flow velocity for the surface
. (25)
Ur0 ` layer Us and the resistance layer Ur were estimated
separately (see Figure 5). These will be used for
Thus, we have derived a new model for the average comparison with the velocity models given in eqs.
velocity in the surface layer. For large flow depths, (11), (12) and (25). From the measured average ve-
equation (25) transforms to Mannings law. For locities in the resistance layer, local Reynolds num-
shallow surface layer flows it approaches the drag- bers were calculated (Re 3 104 4 104 ) and
dominated velocity in the resistance layer. Specifi- subsequently, the drag coefficients CD for the cir-
cation of the unknown scaling length ` in terms of cular cylinders. Standard works on fluid mechanics
measurable geometric parameters, and a determi- (e.g. Schlichting 1979) report that in such a flow
nation of the exponent , follows from comparison regime the drag coefficient remains fairly constant
with laboratory experiments. with a value of nearly 1. Based on four flow exper-
iments without artificial vegetation, the roughness
height kS of the flume bed was determined to be
5 Comparison to data from kS = 2.3 0.6 mm.
flume experiments
5.2 Average velocities in the two
5.1 The flume experiments flow layers
For model calibration we use the experimental Figure 6 shows the measured average velocities in
data from Meijer and van Velzen (1998) (for data the resistance layer as compared to the ones pre-
see also Baptist 2005). The available experimental dicted by equation (11). Although for slow flows
data comprise 48 flow experiments with homoge- the velocities are slightly overestimated, the over-
neously distributed cylindrical stems that all have all agreement between equation (11) and measured
the same stem diameter (D = 8 mm). The cylin- values is very good (R2 = 0.94).
der height k varied between three distinct values The scaling expression for the velocity in the sur-
(k = 0.45, 0.9 or 1.5 m), and flow depths h were face layer, equation (25), consists of two unknown
varied such that relative depths were in the range parameters; the scaling length ` and the transi-
from h/k = 1.3 to h/k = 5.5. Half of the total set tion exponent . Both parameters will now be de-

7
Resistance layer Surface layer
3 15
2 2
m = 64 [m ] m = 64 [m ]
2 2
2.5 m = 256 [m ] m = 256 [m ]

Predicted Us/Ur0 []
Predicted Ur/Ur0 []

2 10

1.5

1 5

0.5 2 2
R = 0.94 R = 0.98

0 0
0 0.5 1 1.5 2 2.5 3 0 5 10 15
Measured Ur/Ur0 [] Measured Us/Ur0 []

Figure 6. Comparison between the proposed model Figure 7. Measured and predicted values (equation
for the depth-averaged velocity Ur in the resistance (27)) of the depth-averaged velocity Us in the surface
layer (equation (11), on vertical axis) and results from layer.
laboratory flume experiments (m = number of cylin-
drical stems per m2 on the bed surface).

termined from comparison with the experimental Table 1. Model results to predict the average velocity
data. Table 1 shows some model results using dif- in the surface layer using Equation (25). For different
ferent geometrical quantitites for the scaling length geometric quantities to describe the scaling length `,
`, and the corresponding values for that fit the the optimal value for exponent is determined (in
measurements best. It turns out that for ` = Ks case ` = Ks three values for are shown).
the proposed scaling law gives best agreement with ` [m] K [-] [-] R2
laboratory data. If the transition exponent is cho- Ks 1.04 4 0.976
sen as = 5, the coefficient of determination is Ks 1.00 5 0.981
R2 = 0.98 and the coefficient of proportionality K Ks 0.99 6 0.978
is practically equal to unity (K = 1.00). Therefore, Kb 4.69 2 0.88
we propose the following expression for the scaling KD 0.22 7 0.60
length ` in the surface layer: KCD D 0.22 7 0.59
Kk 4.29 1 0.50
` = s. (26)

In combination with a scaling length equal to the of 2/3, when the depth of the surface layer is large.
cylinder spacing (equation (26)), a value for the By comparison with experimental data (with rela-
transition exponent of = 5 in Eequation (24) tive depths in the range h/k = 1.3 5.5) we found
gives the best results. Consequently, the new scal- that shallow flows are also well described by intro-
ing relation for the flow velocity in the surface layer ducing a transition exponent of = 5 in equation
becomes (24). This value for implies that the exponent
only deviates significantly from 2/3 for relative

! 2 1( h )5 flow depths h/k < 2.5 (see Figure 4). Therefore,
Us hk 3 k
even though the scaling relation given in equation
= . (27)
Ur0 s (22) was only expected to give realistic results for
deep surface layer flows, it also performs well for
Figure 7 shows how the predictions of the cali- flows in relatively shallow surface layers when us-
brated scaling expression, equation (27), compare ing the scaling length ` = s (equation (26)). If we
to the measured flow velocities. Surface layer veloc- would assume complete similarity between flow ve-
ities are predicted accurately for the entire dataset locity and depth of the surface layer (i.e. by using a
of 48 experiments. constant exponent = 2/3 in equation (25)), small
As shown in section 4.2, theory predicts a depth- values of Us /Ur0 are predicted less accurately, but
independent scaling length `, and a constant power still a coefficient of R2 = 0.94 is found.

8
Total depth 6 Discussion
12
2
m = 64 [m ]
2
The scaling length `, reflecting the flow resistance
10 m = 256 [m ] that the surface layer experiences due to the cylin-
drical elements, is well-represented by the spac-
ing between the cylindrical stems s. Based on this
Predicted UT/Ur0 []

8
result, it is expected that the separation between
cylinders is no longer an appropriate choice for the
6 scaling length in cases that the height of the cylin-
ders is smaller than the distance between them.
4
Therefore, it seems natural to assume that ` may
never exceed the height of the cylinders k. Further
investigations with different geometrical properties
2 2
R = 0.99 of the cylinders, and different spatial distributions,
should be carried out to study these situations.
0 Complications arise when we want to represent
0 2 4 6 8
Measured UT/Ur0 []
10 12 hydraulic resistance of natural vegetation with the
newly proposed two-layer flow model. In the pro-
Figure 8. The predicted depth-averaged velocity over posed flow model the plant height k is considered
the entire flow depth (UT as in equation (28), on ver- fixed and identical for all indivual plants. A fixed
tical axis) vs. measured average velocities. value for the plant height k allowed us to define
two distinct flow layers. Is this still possible when
5.3 A two-layer scaling model for the plant height is spatially variable, or when in
the entire flow depth case of flexible vegetation bending effects decrease
the depth of the resistance layer? An average plant
height is possibly still appropriate for defining two
By using the newly proposed flow velocity mod- flow layers if plant height variability is concen-
els for the resistance layer (equation (11)) and the trated around a well-pronounced mean value. If,
surface layer (equation (27)), the expression for the however, the distribution of different heights shows
average velocity of the entire flow depth (equation distinct peaks then more flow layers should be de-
(4)) becomes: fined, each associated with an average flow veloc-
ity.
Also, the spacing parameter s, that reflects the
s ! 2 1( h )5
UT k hk hk 3 k spatial density of the vegetation, is easily deter-
= + (28) mined for a homogeneous field of identical stems.
Ur0 h h s However, which value should be used for a rep-
resentative plant separation if the spatial distri-
bution of individual plants is not homogeneous,
In Figure 8, the predictions of the new two-layer or when variability between plants exists and side
velocity scaling model for the entire flow depth branches with leaves also obstruct the flow? For
(equation (28)) are compared to the measured av- flow through emergent vegetation this problem is
erage flow velocities. The combined performance more easily solved because the drag coefficient CD
of the velocity models in the resistance and sur- can be adjusted to match flow measurements for
face layer (eqs. (11) and (27)) is even better than a specific choice of plant separation (or similarly,
their separate predictions, yielding a coefficient of a choice of frontal plant area, see e.g. Fischenich
determination R2 = 0.99. and Dudley 2000). Whether it is justified to use
In section 5.2 it was adressed that the variable this same plant separation value to determine the
power exponent in the surface layer only affects flow velocity in the surface layer when the vegeta-
predicted flow velocities for relatively shallow sur- tion becomes overflown is questionable.
face layers (if h/k < 2.5). For the average velocity The difficulties of choosing representative geo-
over the entire depth (equation (28)), this effect is metrical parameters for natural vegetation are not
even further surpressed because for shallow surface unique to the newly found description, but are rel-
layers the overall flow field is largely detrmined by evant to all descriptions that link flow resistance to
flow in the resistance layer. Therefore, predictions geometrical plant parameters. It is recommended
of depth-averaged flow velocities are not very sen- that future work on the flow resistance of natural
sitive to the choice of parameter in equation (25). vegetation adresses these issues.

9
7 Conclusions of Applied Physics at the University of Twente for
his critical views and comments.
In the present study, flow over large-scale rough-
ness elements is described by an average-velocity
model where distinct flow characteristics are at- Notation
tributed to two separate flow layers. By describing
flow by its bulk behavior, we avoided the neces- The following symbols are used in this paper:
sity of integration over depth and the associated
complications of depth-dependent turbulence in-
tensities. Predictions of the new expression for the b [m] Drag length.
depth-averaged velocity for flow with submerged CD [-] Drag coefficient.
rigid vegetation gives excellent agreement with D [m] Diameter of cylindrical
measured flow velocities (coefficient of determina- resistance elements.
tion R2 = 0.99). f [-] Bed resistance function.
The new proposed models that describe the fS [-] Stricklers bed resistance
depth-averged flow velocity in the two flow lay- function.
ers seperately, also each show very good agree- g [m/s2 ] Gravitational acceleration.
ment with measured flow velocities. It was shown h [m] Flow depth.
that depth-averaged flow in the resistance layer is i [-] Bed slope of channel.
adequately described by means of a simple bulk- k [m] Height of resistance elements.
force balance. Apart from the drag coefficient that kS [m] Stricklers roughness height.
is particular to the shape of the resistance ele- K [-] Coefficient of proportionality.
ments, no further calibration parameter is needed ` [m] Unknown scaling length.
for the resistance layer model. For the surface m [m2 ] Bed surface density of
layer, a comparison with results from flume exper- resistance elements.
iments shows that the scaling length `, required r [m] Spacing hydraulic radius of
in the depth-averaged velocity relation, is well- resistance layer.
represented by the spacing between the cylindrical s [m] Separation between individual
stems. A depth-dependent power law exponent is resistance elements.
used to force the scaling law to a realistic limiting R2 [-] Coefficient of determination.
value for relatively shallow surface layer flows. This Ur0 [m/s] Depth-averaged flow velocity
improves predicted velocities in the surface layer in the resistance layer
if the vegetation penetrates about half of the to- for emergent resistance elements.
tal water column, until vegetation becomes emer- ur [m/s] Characteristic eddy-velocity near
gent (and the surface layer vanishes). For relatively top of resistance elements.
larger depths of the surface layer, average flow ve- Ur [m/s] Depth-averaged flow velocity in
locities follow Mannings resistance law. the resistance layer.
The proposed flow resistance model is based Us [m/s] Depth-averaged flow velocity in
on idealized vegetation characteristics with homo- the surface layer.
geneous geometrical and material properties. Be- UT [m/s] Depth-averaged flow velocity
cause of its accurate predictions for such condi- over entire (Total) flow depth.
tions, it is expected that also the hydraulic re- vx,z [m/s] Velocity fluctuations in x
sponse to natural vegetation can be described with and z direction, respectively.
the model. [-] Transition exponent.
[m2 / s3 ] Energy dissipation rate.
[-] Similarity exponent.
Acknowledgements [kg/m3 ] Water density.

This research is supported by the Technology


Foundation STW, applied science division of NWO
and the technology programme of the Ministry REFERENCES
of Economic Affairs (the Netherlands) and the
German Federal Institute of Hydrology (Bunde- Akilli, H. and D. Rockwell (2002). Vortex formation from
a cylinder in shallow water. Physics of Fluids 14 (9),
sanstallt f
ur Gewasserkunde). We gratefully thank 29572967.
HKV Consultants for granting us permission to Armanini, A., M. Righetti, and P. Grisenti (2005). Direct
use data of their flume experiments, and thank measurement of vegetation resistance in prototype
Francisco Fontanele Araujo from the Department scale. Journal of Hydraulic Research 43 (5), 481487.

10
Baptist, M. J. (2005). Modelling Floodplain Biogeomor- Meijer, D. G. and E. H. van Velzen (1998). Prototype-
phology. Ph. D. thesis, Delft University of Technol- scale flume experiments on hydraulic roughness of
ogy. submerged vegetation. Technical Report PR 121,
Baptist, M. J., V. Babovic, J. Rodriguez Uthurburu, HKV Consultants, Lelystad, The Netherlands.
M. Keijzer, R. E. Uittenbogaard, A. Mynett, and Neary, V. S. (2003). Numerical solution of fully-
A. Verwey (2006). On inducing equations for vegeta- developed flow with vegetative resistance. Journal
tion resistance. Journal of Hydraulic Research. (sub- of Engineering Mechanics 129 (5), 558563.
mitted). Nikuradse, J. (1933). Stromungsgesetze in rauhen
Barenblatt, G. I. (2003). Scaling. Cambridge Texts in rohren. Forschungsheft VDI Verlag, Berlin, Ger-
Applied Mathematics. Cambridge University Press. many.
Choi, S. U. and H. Kang (2004). Reynolds stress model- Petryk, S. and G. Bosmajian (1975). Analysis of flow
ing of vegetated open-channel flows. Journal of Hy- through vegetation. Journal of the Hydraulics Divi-
draulic Research 42 (1), 311. sion 101 (HY7), 871884.
Pope, S. B. (2000). Turbulent Flows (third ed.). Cam-
Darby, S. (1999). Effect of riparian vegetation on flow
bridge University press.
resistance and flood potential. Journal of Hydraulic
Engineering 125 (5), 443454. Ree, W. O. and F. R. Crow (1977). Friction factors for
vegetated waterways of small slope. Technical Re-
Erduran, K. S. and V. Kutija (2003). Quasi-three- port Publication S-151, U.S. Department of Agricul-
dimensional numerical model for flow through flexi- ture, Agricultural Research Service.
ble, rigid, submerged and non-submerged vegetation.
Journal of Hydroinformatics 5 (3), 189202. Schlichting, H. (1979). Boundary Layer Theory (7th ed.).
New York: McGraw-Hill.
Fischenich, C. and S. Dudley (2000). Determining drag
coefficients and area for vegetation. Ecosytem Man- Shimizu, Y. and T. Tsujimoto (1994). Numerical anal-
agement & Restoration Research Program SR 08, ysis of turbulent open-channel flow over vegetation
U.S. Army Corps of Engineers. layer using a k turbulence model. Journal of Hy-
droscience and Hydraulic Engineering 11 (2), 5767.
Gioia, G. and F. A. Bombardelli (2002). Scaling and sim- Smart, G. M., M. J. Duncan, and J. M. Walsh (2002).
ilarity in rough channel flows. Physical Review Let- Relatively rough flow resistance equations. Journal
ters 88 (1), 1450114504. of Hydraulic Engineering 128 (6), 568578.
Green, J. C. (2006). Effect of macrophyte spatial vari- Stephan, U. and D. Gutknecht (2002). Hydraulic resis-
ability on channel resistance. Advances in Water Re- tance of submerged flexible vegetation. Journal of
sources 29, 426438. Hydrology 269, 2743.
James, C. S., L. Birkhead, A. A. Jordanova, and J. J. Stone, B. M. and H. T. Shen (2002). Hydraulic resis-
OSullivan (2004). Flow resistance of emergent veg- tance of flow in channels with cylindrical roughness.
etation. Journal of Hydraulic Research 42 (4), 390 Journal of Hydraulic Engineering 128 (5), 500506.
398. Strickler, A. (1923). Beitrage zur Frage der
Jarvela, J. (2002). Flow resistance of flexible and stiff Geschwindigkeitsformel und der Rauhigkeitszahlen
vegetation: a flume study with natural plants. Jour- f
ur Strome, Kanale und geschlossene Leitungen.
nal of Hydrology 269, 4454. Mitteilungen des Amtes f ur Wasserwirtschaft 16,
Keijzer, M. and V. Babovic (2000). Genetic program- Eidg. Department des Innern, Bern, Switserland.
ming within a framework of computer-aided discov- Wilson, C. A. M. E., T. Stoesser, P. D. Bates, and
ery of scientific knowledge. In D. Whitley, D. Gold- A. Batemann-Pinzen (2003). Open channel flow
berg, E. Cantu-Paz, L. Spector, I. Parmee, and H. G. through different forms of submerged flexible vegeta-
Beye (Eds.), Proceedings of the Genetic and Evo- tion. Journal of Hydraulic Engineering 129, 847853.
lutionary Computation Conference (GECCO-2000), Yen, B. C. (2002). Open channel flow resistance. Journal
Las Vegas, pp. 543550. Morgan Kaufmann. of Hydraulic Engineering 128 (1), 2039.
Keulegan, G. H. (1938). Laws of turbulent flow in open
channels. Journal of Research of the National Bureau
of Standards 21 (Research paper 1151), 707741.
Klopstra, D., H. J. Barneveld, J. M. van Noortwijk, and
E. H. van Velzen (1997). Analytical model for hy-
draulic roughness of submerged vegetation. In Man-
aging Water: Coping with scarcity and abundance,
San Fransisco, pp. 775780. Proceedings of the 27th
IAHR Congress.
Kouwen, N. and M. Fathi-Moghadam (2000). Friction
factors for coniferous trees along rivers. Journal of
Hydraulic Engineering 126 (10), 732740.
Kouwen, N. and T. E. Unny (1973). Flexible roughness
in open channels. Journal of the Hydraulics Divi-
sion 99 (5), 713728.
Manning, R. (1889). On the flow of water in open chan-
nels and pipes. Transactions of the Institution of
Civil Engineers of Ireland 20, 161207.

11

You might also like