You are on page 1of 21

Journal of Arid Environments (2001) 49: 485}505

doi:10.1006/jare.2001.0807, available online at http://www.idealibrary.com on

Drag coefficients, roughness length and zero-plane


displacement height as disturbed by artificial
standing vegetation

Zhibao Dong*-, Shangyu Gao* & Donald W. Fryrear?


*Institute of Desert Research, Chinese Academy of Sciences, No. 260,
West Donggang Road, Lanzhou, Gansu 730000, P.R.China
? USDA-ARS, Big Spring, TX-79720-0909, U.S.A.

( Received 1 September 1999, accepted 20 February 2001, published electronically 24 September 2001)

Standing vegetation, for its disturbance to the near-surface airflow, is widely


used in the arid and some semi-arid lands over the world to control wind
erosion. Knowledge of drag coefficient (CD), roughness length (Z0) and
displacement height (D) is essential to completely define the state of wind and
the protective role of standing vegetation in wind erosion. Using standing sticks
as model standing vegetation, detailed wind velocity distributions were mea-
sured above the vegetated surface in wind tunnel, a mass conservation method
to estimate the zero-plane displacement height was tried, and the drag coef-
ficients and roughness length were derived by a curve-fit method. Due to the
disturbance of standing vegetation, the velocity distributions deviated from the
logarithmic profile. The deviation increased with increasing vegetation density
and height. The wind velocity profile disturbed by standing vegetation can be
divided into three sections, and each section can be expressed by a logarithmic
function. Displacement height is a significant parameter for tall and dense
vegetation; choice of the value for displacement height greatly influences other
parameters such as roughness length and drag coefficient, but its accurate
value is not available yet. In the authors’ philosophy, drag coefficient and
roughness length should be recommended. However, caution should also be
taken in selecting the wind velocity measurements used to derive drag coef-
ficient and roughness length. Both the height and density of standing vegeta-
tion influenced drag coefficient and roughness length, but their relative
importance was different. A new parameter, effective lateral cover
(Lec) was introduced to characterize the structure of standing vegetation. It was
found that effective lateral cover and height/spacing ratio were better
structural parameters than the others when the effects of standing vegeta-
tion on drag coefficient and roughness length were assessed. Good correla-
tion existed between the derived roughness length and drag coefficient,
implying that obtaining drag coefficient and roughness length of vegetated
surface by curve fit method is reliable so long as caution is taken in selecting
the appropriate measurements. Of the three parameters roughness length is
the most sensitive to characterize the effects of roughness length on near-
surface airflow.
 2001 Academic Press
Keywords: drag coefficient; roughness length; displacement height; velo-
city profile; structural parameters of standing vegetation

- Corresponding author.
0140}1963/01/110485#21 $35.00/0  2001 Academic Press
486 Z. DONG ET AL.

Introduction

The protective roles of vegetation in reducing or controlling wind erosion have long been
recognized by agriculturists. Vegetation, in the form of cover crops, wind strips or
post-harvest residues is typically used as a means of managing agricultural soils prone to
wind erosion around the world. Similarly, trees and shrubs are used as windbreaks to
inhibit the transport of snow or sand across roads and fields. The amount and form of
vegetative cover is also a key factor required to estimate soil loss by wind (Woodruff
& Siddoway, 1965; Hagen, 1991; Fryrear et al., 1994). Agricultural engineers, micro-
meteorologists and fluid mechanists have done much of the work concerned with the
role of vegetation in wind erosion (Lee, 1991; Wolf & Nickling, 1993). However, no
systematic design procedures for maximizing the effectiveness of vegetative
measures in controlling wind erosion have been developed except for the design of
vegetative belts (Hagen & Skidmore, 1971; Hagen, 1976; Schwartz et al., 1995, 1997)
due to the lack of an insight into the mechanism of the disturbance of vegetation to the
near-surface air stream. A number of researchers have measured the influence of
vegetation on wind erosion rates (Englehorn et al., 1952; Lyles & Allison, 1981; Fryrear,
1985; Wasson & Nanninga, 1986; Buckley, 1987; van den Ven et al., 1989). Qualitative
descriptions and quantitative formula developed show that the protection afforded
by vegetation against wind erosion is obvious (Fryrear, 1985). For example, Wasson
& Nanninga (1986) measured little erosion when the vegetation coverage reached
35}40%. van den Ven et al. (1989) illustrated that even sparse vegetation can signifi-
cantly reduce soil loss by wind. Micrometeorologists have attempted to define the wind
regime that determines the magnitude and direction of wind force on the erodible soil
surface within vegetation canopies (Lee & Soliman, 1977; Bacche, 1986; Arya, 1988;
Stockton & Gillette, 1990; Musick & Gillette, 1991). Such parameters that include
roughness length or displacement height derived from the observed wind profiles have
been used to assess the effects of vegetation on wind erosion (Abtew et al., 1989;
Lee, 1991). Arrays of artificial standing residues were modeled in a wind tunnel and
the wind drag measured to derive drag coefficients that were evaluated in relation to
the protection offered by sparse vegetation against wind erosion in arid regions
(Marshall, 1970, 1971).
Essentially, vegetation controls wind erosion by protecting the covered soil from wind,
by absorbing wind momentum so that less shear is exerted on the erodible surface, and
by trapping eroded material. Although crop residues have proven effective in
controlling wind erosion, standing vegetation has been found to be several times more
effective in reducing wind erosion losses than vegetation lying flat on soil surface
(Siddoway et al., 1965; Unger, 1994). The direct effect of standing vegetation is the
increased roughness that results in more absorption of the wind momentum from the air
stream. The absorption of momentum leads to drag. The roughness length and drag
produced can be used to assess the protection offered by standing vegetation
against wind erosion. The primary concern of this study focuses on the roughness and
drag properties of the surface with standing vegetation by means of modeling tests in a
wind tunnel. The specific objectives were: to (1) analyse the adaptation of wind profiles to
different configurations of modeled standing vegetation; (2) to define the relationship
between roughness length, drag coefficient and the structural parameters of standing
vegetation; and (3) to determine which parameter is the most reliable for characterizing
the effectiveness of standing vegetation to provide protection from wind erosion.

Definitions and relations

It is well known that for a neutral turbulent boundary layer where the temperature
gradient is small and of uniform density, the wind profiles follow a log law:
DRAG COEFFICIENTS DISTURBED BY ARTIFICIAL STANDING VEGETATION 487

U"(U /K ) ln((Z!D)/Z0) (1)


*
and U is defined as:
*
U "(q/o)1/2 (2)
*
where U is the wind velocity at height Z, U is drag velocity, Z0 is roughness length
*
dependent on the surface texture, D is the displacement height, K is Karman’s constant,
q is surface shear stress and o is air density.
Equation (1) defines a group of straight lines or velocity rays known as wind profiles
when the velocities are plotted against a log-scale of height. It bears three important
parameters, U , Z0 and D. Z0 fixes the point at which the wind speed approximately
*
becomes zero at distance above the surface, and U defines the gradients of the rays. The
*
displacement length (D) is an empirical parameter introduced to preserve the ‘log-law’
above tall vegetation. Both Z0 and D can be used to evaluate the disturbance of
vegetation to the near-surface airflow (Abtew et al., 1989). For rough surface that
has greater roughness length compared with the ideal smooth bed, we can either
think that Z0 in Eqn (1) is increased with respect to the same reference plane
(zero-plane), or the reference plane characterized by D in Eqn (1) is elevated higher.
Although the value chosen for D can significantly affect the analysis of other
micrometeorological parameters, especially flux-profile relationship and stability
functions, the interpretation of the zero-plane displacement and its physical significance
becomes controversial (Molion & Moore, 1983). To date, no accurate method can be
applied to determine D, which still remains a matter of crude estimation in practice.
To compare Z0 a specific value should be ascribed to D so that the analysis has the
same reference plane. Mathematically, in Eqn (1) much information concerned with
D can be included in Z0 when D in Eqn (1) is considered to be a constant such as zero.
Then
U"(U /K ) ln(Z/Z0) (3)
*
So a knowledge of U and Z0 completely defines the state of wind. The effects of
*
standing vegetation on wind can be assessed by examining changes in U and Z0 as
*
a function of the configuration of the vegetation. Equation (2) illustrates that U is also
*
a dimension of the drag exerted on the surface because the air density (o) can be
considered as a constant.
For bodies immersed in fluid flow the drag, or the force component parallel to the
fluid motion (FD) is customarily written as
FD"12 CD AoV 2 (4)
where CD is drag coefficient, A is a characteristic area, o is fluid density and V is
a uniform fluid velocity (Pao, 1967). Except for a few simple cases, CD is determined
experimentally.
For airflow, Taylor (1916) introduced a factor known as the coefficient of skin
friction and developed velocity square law to relate the drag and wind velocity. Sutc-
liffe (1936) later referred to this as ‘the surface resistance to atmospheric flow’, and
argued that knowledge of the coefficient would allow the rate at which horizontal
momentum is transferred from atmosphere to surface be calculated. In the succeeding
years comprehensive observations and considerable advances made in turbulence the-
ories and the advent of similarity theories for both atmospheric surface and boundary
layers, allow the drag coefficient (analogous to Taylor’s coefficient) to be defined
explicitly in terms of aerodynamic roughness, atmospheric thermal stability and other
relevant physical parameters. So knowledge of the drag coefficient of a vegetated
surface is essential to understanding its resistance to the near-surface air or the
protection it offers against wind erosion.
488 Z. DONG ET AL.

In the constant flux layer of the atmosphere adjacent to a horizontally homogeneous


surface the local surface stress or drag may be written in terms of the wind velocity (U )
at height Z (Taylor, 1916) as:
q"oCDU 2 (5)
where CD may be related to the aerodynamic roughness length (Z0) and stability
parameter (f) through the Monin}Oboukhov theory, viz.,

DN t(f)]
CD"CDN /[1!K\1C1/2 2
(6a)
CDN"K 2/(ln Z/Z0)2 (6b)
Knowledge of the empirical form of the non-dimensional wind gradient allows
CD /CDN to be computed as a function of f (or a bulk Richardson number) and Z/Z0
(Garratt, 1977). Frank & Kocurek (1994) also pointed out that atmospheric conditions
effect the wind profile and sand transport, but for the airflow in this study, the
temperature gradient is very small throughout the layer and the effects of atmo-
spheric conditions on wind profiles can be neglected, then
CD"CDN"K 2/(ln Z/Z0)2"(U /U )2 (7)
*
So U in Eqn. (1) can be replaced by CD .
*

Materials and methods

Wind tunnel tests

This study was carried out in a wind tunnel at the USDA-ARS Agricultural Research
Station, Big Spring, Texas. The suction type non-circulating wind tunnel has a cross-
sectional area of 0·5;1·0 m and an experimental section 2 m long. Screens of dif-
ferent mesh were installed at different locations of the upwind entrance to diminish
large-scale eddying and establish logarithmic wind profiles of the experimental section
for the smooth floor surface, which functioned similarly to modifying the tunnel surface
roughness upwind of the test tray. The screens were kept unchanged throughout the
experiment. A removable smooth flat tray was set in the experimental section and its
surface leveled to the tunnel floor.
Standing vegetation in most agricultural fields is usually distributed randomly except
for crops, but those spaced uniformly can provide maximum sheltering capacity in wind.
The general design of artificial standing vegetation and their distribution in this study
was not intended to provide a highly realistic simulation of the field situation. The
experiment modeled standing vegetation without leaves with specific stalk diameter, but
different height and density using wooden cylindrical dowels. This design was
chosen because it minimized the geometrical complexity of the standing vegetation, and
thereby introduced a minimum of variables to be examined. Cylindrical dowels, 0·3 cm
in diameter, of different density and height were distributed in staggered array on
the removable tray (0·5;2·0 m). Wind profiles for nine free-stream velocities (5, 6, 7, 8,
9, 10, 11, 12, 13 m s\1) for each configuration (Table 1) were measured immediately
downwind of the tray to ensure the wind profiles are representative of the roughness of
the test section before they start adjusting to the smoother downwind surface. Wind
speed profiles were measured using nine Pitot-static probes at 0·5, 1, 2, 3·75, 7, 14, 27,
52 and 85 cm above the floor using differential U-tube alcohol manometers.
The manometer readings were corrected for the effects of atmospheric pressure
and temperature. Five repetitions of the nine wind speeds were run for each
configuration.
DRAG COEFFICIENTS DISTURBED BY ARTIFICIAL STANDING VEGETATION 489

Table 1. Structural parameters of artificial standing vegetation

Serial No. Ds (n m\2) h (cm) AR p Lc h/Sp C

01 25 2·5 8·33 0·094 0·0019 0·11 0·00018


02 25 5·0 16·67 0·047 0·0038 0·22 0·00018
03 25 7·5 25·00 0·031 0·0057 0·33 0·00018
04 25 10·0 33·33 0·024 0·0075 0·44 0·00018
05 25 12·5 41·67 0·019 0·0094 0·55 0·00018
06 25 15·0 50·00 0·016 0·0100 0·66 0·00018
07 100 2·5 8·33 0·094 0·0075 0·22 0·00071
08 100 5·0 16·67 0·047 0·0150 0·44 0·00071
09 100 7·5 25·00 0·031 0·0230 0·66 0·00071
10 100 10·0 33·33 0·024 0·0300 0·89 0·00071
11 100 12·5 41·67 0·019 0·0375 1·11 0·00071
12 100 15·0 50·00 0·016 0·0450 1·33 0·00071
13 200 2·5 8·33 0·094 0·0150 0·32 0·0014
14 200 5·0 16·67 0·047 0·0300 0·64 0·0014
15 200 7·5 25·00 0·031 0·0450 0·95 0·0014
16 200 10·0 33·33 0·024 0·0600 1·27 0·0014
17 200 12·5 41·67 0·019 0·0750 1·59 0·0014
18 200 15·0 50·00 0·016 0·0900 1·91 0·0014
19 400 2·5 8·33 0·094 0·03 0·46 0·0028
20 400 5·0 16·67 0·047 0·06 0·91 0·0028
21 400 7·5 25·00 0·031 0·09 1·37 0·0028
22 400 10·0 33·33 0·024 0·12 1·83 0·0028
23 400 12·5 41·67 0·019 0·15 2·28 0·0028
24 400 15·0 50·00 0·016 0·18 2·74 0·0028
25 800 2·5 8·33 0·094 0·06 0·66 0·0057
26 800 5·0 16·67 0·047 0·12 1·32 0·0057
27 800 7·5 25·00 0·031 0·18 1·98 0·0057
28 800 10·0 33·33 0·024 0·24 2·64 0·0057
29 800 12·5 41·67 0·019 0·30 3·30 0·0057
30 800 15·0 50·00 0·016 0·36 3·97 0·0057
31 1300 15 50·00 0·016 0·59 5·17 0·0092
32 1600 2·5 8·33 0·094 0·12 0·97 0·011
33 1600 5·0 16·67 0·047 0·24 1·93 0·011
34 1600 7·5 25·00 0·031 0·36 2·90 0·011
35 1600 10·0 33·33 0·024 0·48 3·87 0·011
36 1600 12·5 41·67 0·019 0·60 4·83 0·011
37 1600 15·0 50·00 0·016 0·72 5·80 0·011
38 2400 15·0 50·00 0·016 1·08 7·29 0·017

The diameter of dowels"0·3 cm, Ds"Density, h"Height, AR"Aspect ratio, p"The ratio of basal to
frontal silhouette area, Lc"Lateral cover, C"Coverage.

Scaling

For small-scale experimental models to adequately simulate field phenomena, a number


of dimensionless variables should not differ significantly from full-scale values
(Kind, 1976; Iversen, 1980). Physical constraints that are functions of wind tunnel
dimension rarely allow these similitude requirements to be met simultaneously, and
some compromise must be adopted (Musick et al., 1996). The selection of the
490 Z. DONG ET AL.

dimension of artificial soil clod and wind velocity was designed to permit the approximate
adequacy of scaling so that the results from this test will be significant for field situations.
The main aspect of scaling considered was Reynolds Number similarity with the field
situation. Reynolds Number (Re ) was defined as:
Re"UL/l (8)
where U is wind velocity, L is characteristic length, l is the kinematic viscosity,
(l+0·15 cm2 s\1 at normal temperature and air pressure). Residues left standing for
erosion control, such as sunflower, sorghum and kenaf, are typically 40}120 cm high.
For the field situation using Eqn (8) gives Re of using vegetation with height (h) and
wind velocity of 5 m s\1 at the typical wind sensor height of meteorological station is
1·33;105 to 4·0;105. In this experiment, at the maximum wind velocity used,
13 m s\1, and the maximum model vegetation height of 15 cm the Reynolds’ Number at
the centerline height was about 1·3;105. The wind tunnel experiments were conducted
at wind velocities and model vegetation heights to provide drag coefficient and
roughness length data over the Reynolds’ Number range 8·33;103 to 1·3;105.

Parameters to characterize the structure of standing vegetation

Review of the literature indicates that various dimensionless parameters are used to
characterize the structure of vegetation when trying to assess its roles in controlling wind
erosion (Wolf & Nickling, 1993). Parameters defining the geometry of vegetation
include aspect ratio (AR), height/spacing ratio (Sh ), the ratio of basal to frontal silhouette
area (p), coverage (C), lateral cover (Lc ), and element porosity (Pf ). Various relation-
ships have been developed to relate roughness length or drag coefficient to the listed
parameters (Abtew et al., 1989).
The lateral cover or roughness concentration was first defined by Schlichting (1936).
It is defined as the ratio of the sum of the upstream projected area of the roughness
elements, such as standing vegetation, to the floor area
Lc"hdN/A (9)
where Lc is the lateral cover, h, d, N, are the height, diameter and the number of the
roughness elements, A is the area of the floor.
For cylindrical elements, the aspect ratio is defined as the ratio of cylinder height to
diameter:
AR"h/d (10)
where AR is aspect ratio, h and d are the height and stalk diameter of the vegetation.
For a cylinder, AR is related to the ratio of basal to frontal-silhouette area as:
p"n/4AR (11)
where p is the ratio of basal to frontal area.
The height/spacing ratio is defined as:
Sh"h/Sp (12)
where Sh is the height/spacing ratio, h is the height of vegetation and Sp is the distance
between two nearby elements.
Porosity is defined as the ratio of the overall frontal-silhouette area to the total stem
frontal-silhouette area.
Pf"DH/hdN (13)
where Pf is the porosity, D, H are the width (or diameter) and height of the vegetated
area, h, d and N are the height, stalk diameter and number of plant bodies.
The structural parameters of the modeled standing vegetation are listed in Table 1.
DRAG COEFFICIENTS DISTURBED BY ARTIFICIAL STANDING VEGETATION 491

Derivation of displacement height, roughness length and drag coefficient

Displacement height in Eqn (1) is estimated by a mass conservation method proposed


by Molion & Moore (1983) that proves more advantageous than graphical and various
analytical least-square error methods. D is calculated by

   
N
Zf!Z0s 1
D" Zf! ! (Ui#Ui#1)(Zi!Zi!1)#U1 Z1 (14)
ln(Zf /Z0s ) 2U i"1

where Ui is the wind velocity at height Zi , N is the number of height levels, U is the
free-stream wind velocity at free-stream height which is the centerline height in wind
tunnel, Z0s is the roughness length of the smooth tunnel floor.
Once displacement height is determined, roughness length and drag coefficient in
Eqn (1) are derived by the least-square curve-fit method. The measured wind profiles
were fit through the least squares regression to Eqn (15) to derive roughness length (Z0)
and drag coefficient (CD) for each configuration of dowels.
U"A#B ln(Z!D) (15)
where U is the wind velocity at height Z, A and B are regression constants.
In Eqn (15), Z0 can be obtained when U"0, therefore,
Z0"exp(!A/B) (16)
Suppose U1, U2 are the wind velocities at height Z1 and Z2 , then
By Eqn (1),
U1!U2"(U /K )(ln(Z1!D)!ln(Z2!D)) (17)
*
By Eqn (15)
U1!U2"B(ln(Z1!D)ln(Z2!D)) (18)
So,
U "KB (19)
*
Substitute Eqn (19) to Eqn (7), then
CD"(KB/U )2 (20)
If the centerline height is selected as the reference height, then
CD"(KB/Uc)2 (21)
where Uc is the centerline wind velocity. Therefore, Z0 and CD could be obtained by
Eqns (16) and (21) so long as the fitting function, Eqn (15) was established.
The fetch effect in the wind tunnel was negligible because the test section was
only 60 cm to the screens at the inlet section. The wind profiler used to monitor the wind
profile was close enough to the test section to ensure the measured wind profiles are
representative of the roughness of the modeled standing vegetation. Only those
measurements from probes determined to lie within the logarithmic region of the rough
sub-boundary layer were used to determine Z0 and CD . No fewer than four probes were
used in any case. For taller and denser standing vegetation, the lowest probe used in the
fit was typically just below the plane of vegetation tops. Upper probes were excluded
from the fit when they were estimated to deviate significantly from the logarithmic law,
as determined by Cole’s wake profile, quoted by Wolf & Nickling (1996). Those results
were eliminated if the logarithmic curve fit gave a value of R2(0·90.
492 Z. DONG ET AL.

Results and discussion

Adaptation of wind velocity profiles

Detailed wind velocity profiles above different character of standing vegetation


(height series: 0, 2·5, 5·0, 7·5, 10·0, 12·5 and 15 cm; density series: 0, 25, 100, 200, 400,
800, 1300, 1600, 2400 n m\2) were monitored, and some of them are illustrated in
Fig. 1. It can be seen that on the semi-log plots the wind profiles over smooth floor were
almost linear, the wind velocities at all nine heights fitted Eqn (3) reasonably well
(R 2'0·97).
Above a specific character of standing vegetation the shape of velocity profiles at
different wind speeds were very similar. As vegetation density increased over
100 n m\2 or when the dowels were over 7·5 cm high the wind profiles deviated from
the log law. The wind profile could be divided into three sections, or sub-layers each with
different wind gradients. The lowest or inner sub-layer with the least wind gradient
ranged from 1 to 8 cm thick and its thickness increased with vegetation density and
height. For standing vegetation 15 cm high at a density over 1600 n m\2, little air motion
occurred in the inner sub-layer. This implied that no wind force was exerted on the
intervening surface and the ground surface would be totally protected. The top or outer
sub-layer had almost the same wind gradient as the inner sub-layer and its lowest point
ranged from 7 to 27 cm high above the surface, increasing with vegetation density and
height. The middle or rough sub-layer had the greatest wind gradient. The gradient
increased with vegetation height and density. Assessment of all the wind tunnel results
indicates that velocity profile within each sub-layer was logarithmic. In summary, the
velocity profiles over vegetated surface can be described as discontinuous functions. The
velocity curves versus height showed three different logarithmic sections, each with
different intercepts and slopes.

Zero-plane displacement height of vegetated surface

Figure 2 shows the estimated displacement height of the standing vegetation. Displace-
ment height increases with both height and density of standing vegetation. Very short
and sparse vegetation (height (5·0 cm and density (800 n m\2) has no displacement
height. The maximum displacement height of dense standing vegetation (2400 n m\2)
reaches 0·87 vegetation height. These results imply that displacement height estimated
by Eqn (14) is a good parameter to characterize the influence of tall vegetation on
near-surface airflow. However, the corresponding roughness length (Fig. 3) taking into
account of displacement height is very scattered. This is because roughness length is
very sensitive to displacement height and to derive meaningful roughness length in Eqn
(1) accurate values for displacement height are needed. The anomalies in Fig. 3 and
those recognized by previous researchers (Thom et al., 1975; Raupach, 1979; Garratt,
1980, Abtew et al., 1989) imply that the interpretation, physical significance of zero-
plane displacement height and its calculation method themselves are issues in need of
further study. In the authors’ philosophy, roughness length has clear physical signifi-
cance and sound theoretical basis and is a better parameter to characterize the interac-
tions between roughness elements and near-surface airflow. To evaluate the variation of
roughness length with standing vegetation a fixed value, zero value is ascribed to D in
Eqn (1) so that the question is simplified. In fact the introduction of D in Eqn (1) is to
preserve the ‘log-law’ that can also be ensured by the higher correlation coefficient
required in deriving drag coefficients and roughness length by curve-fit method
when D is zero. The following discussion on drag coefficients and roughness length
is based on zero value for D.
DRAG COEFFICIENTS DISTURBED BY ARTIFICIAL STANDING VEGETATION 493
Figure 1. The wind profiles over the vegetated surface.
494
Z. DONG ET AL.
Figure 1. (Continued).
DRAG COEFFICIENTS DISTURBED BY ARTIFICIAL STANDING VEGETATION 495
Figure 1. (Continued).
496 Z. DONG ET AL.

Figure 2. The estimated displacement height of standing vegetation. Ds"25 ( ); Ds"100


( ); Ds"200 ( ); Ds"400 ( ); Ds"800 ( ), Ds"1600 ( ).

Drag coefficients of vegetated surface

The drag coefficients (CD) of vegetated surface were obtained by Eqn (21). Above
analysis of the velocity profiles over vegetated surface indicates that the wind profiles
disturbed by vegetation were very complicated, so caution should be taken in using
Eqn (15) to derive CD . The main concern here is how to determine coefficient B,
which is greatly dependent on which measurements were used in the curve fitting
procedure. The selection of the measurements used to fit was based on the velocity
profile curves shown in Fig. 1. Drag coefficients of the vegetated surface with
different height and density at different centerline wind velocities were ob-
tained. Although CD should decrease with wind velocity, the wind tunnel results indicate
that in this study the change of CD with wind velocity was complex. Of the 39 groups
(each group had nine wind velocities), 18 showed decrease, 16 irregular change, three
increase and two almost no change of CD with wind velocity. So the primary concern
here focuses on comparing the influence of vegetation structure on CD . The final drag
coefficients were the mean values of CD at different velocities. The results are
shown in Fig. 4.

Figure 3. Roughness length of standing vegetation with displacement height illustrated in Fig. 2.
Ds"25 ( ); Ds"100 ( ); Ds"200 ( ); Ds"400 ( ); Ds"800 ( ),
Ds"1600 ( ).
DRAG COEFFICIENTS DISTURBED BY ARTIFICIAL STANDING VEGETATION 497

Figure 4. Wind tunnel results of the drag coefficients of the surface with standing vegetation.
Ds"25 ( ); Ds"100 ( ); Ds"200 ( ); Ds"400 ( ); Ds"800 ( ),
Ds"1600 ( ).

The CD of the tested bare surface free of any disturbance of standing vegetation was
about 0·00109, a little less than that derived by Wolf & Nickling (1996) from wind
tunnel tests on bare surfaces. The difference might be largely due to the reference
height adopted. The drag coefficients of the bare surface determined by Wolf and
Nickling from the wind speeds at a height of 15 cm ranged from 0·00381 to 0·00503.
While the U in Eqn (21) here referred to the wind velocity at the height of 52 cm, its
values should have been a little grater than those at 15 cm, leading to lower CD values.
CD increased 1·2}24·8-fold with the addition of standing vegetation. For example, the
CD for the 15 cm high and 2400 n m\2 in density was 0·0269. CD increased with both
the density and height of the standing vegetation. The relationships between drag
coefficient and vegetation height, density (Tables 2 & 3) were obtained by statistical
method. Very good correlation existed between CD and vegetation density and height.
Tables 2 and 3 reveal that the relationship between CD and h and that between CD and
Ds could be expressed as power functions (Eqn (22) and (23)).

CD"a1#b1hc (22)

CD"a2#b2D ks (23)

Table 2. The regression analysis results of the relationship between drag coef-
ficient and the height of standing residues
Significance
Ds (n m\2) a1 b1 (;10\4) c R2 level
25 0·00109 0·921 0·7730 0·96 0·001
100 0·00114 0·365 1·5550 1·00 0·001
200 0·00105 1·31 1·4335 1·00 0·001
400 0·00130 1·02 1·7440 1·00 0·001
800 0·00101 1·67 1·6600 1·00 0·001
1600 0·00115 4·18 1·3904 1·00 0·001
Fitting function: CD"a1#b1h c.
498 Z. DONG ET AL.

Table 3. The regression analysis results of the relationship between drag coef-
ficient and the density of standing residues
Significance
h (cm) a2 b2 (;10\4) k R2 level
2·5 0·00120 0·035 0·8479 0·95 0·001
5·0 0·00106 0·422 0·5977 0·98 0·001
7·5 0·000861 1·88 0·4909 0·97 0·001
10·0 0·000663 2·67 0·5100 0·97 0·001
12·5 0·000490 3·00 0·5348 0·98 0·001
15·0 0·000010 5·71 0·4947 0·96 0·001
Fitting function: CD"a2#b2D ks.

where h and Ds are vegetation height and density, a1, b1, a2 , b2 , c and k are regression
constants.
However, their exponents were different. The exponent c ranged from 0·773 to
1·744, with an average of 1·43, while the exponent k ranged from 0·4947 to 0·8479, with
an average of 0·58. This implies that the effects of the height and density of
vegetation on drag coefficient are different. The drag coefficient tended to
increase more with increasing height of vegetation, but the increase rate get less with
increasing density. This might be explained by the inter-sheltering effect of the
airflow regime around each plant body. Airflow around a single plant is separated into
six energy regions: the covered area, slight wind area immediately behind the plant, the
wake zone, the zone of accelerated flows on both sides, and the unaffected area
(Wasson & Nanninga, 1988). When vegetation is sparsely distributed, the flow regime is
one of isolated flow, where all six regions can be fully developed. However, as the density
increases the energy regions around the plants tend to interfere with each other, resulting
in wake flow and eventually skimming flow if the density is sufficient (Lee, 1991). In
the wake flow some parts of the downwind plants are in the wake area of the upwind
ones. In skimming flow the downwind plants are almost totally immersed in the wake of
the upwind ones. Therefore, the wind drag or the resistance of a single plant is reduced
by adding plants (at given velocity), resulting more and more stable drag or drag
coefficient.

Figure 5. The relationship between drag coefficient and lateral cover.


DRAG COEFFICIENTS DISTURBED BY ARTIFICIAL STANDING VEGETATION 499

Figure 6. The relationship between drag coefficient and height/spacing ratio.

The relationships between drag coefficient and other parameters to characterize


standing vegetation were also analysed. Of the various parameters the lateral cover (Lc )
and height/spacing ratio (Sh ) were selected. The correlation between drag coeffi-
cient and lateral cover, height/spacing ratio was obtained (Eqn (24), Eqn (25), Fig. 5,
Fig. 6).
CD"0·00096#0·0243L0·7683
c R 2"0·89 (24)
CD"0·00332S1·05
h R 2"0·96 (25)
Therefore, of the two parameters height/spacing better characterizes the effect of
the vegetation’s structure on drag coefficient. The weaker correlation between drag
coefficient and lateral cover may be due to the inter-sheltering effect of plant
bodies as explained earlier. Equation (24) means that the height and density of vegeta-
tion have the similar importance in changing the drag coefficient, as is not the case
(Tables 2 & 3). Tables 2 and 3 (the exponents c and k) shows that with the increasing
lateral cover of vegetation, the height of vegetation is more important than density in

Figure 7. The relationship between drag coefficient and effective lateral cover.
500 Z. DONG ET AL.

Figure 8. Wind tunnel results of the roughness length of the surface with standing vegetation.
Ds"25 ( ); Ds"100 ( ); Ds"200 ( ); Ds"400 ( ); Ds"800 ( ),
Ds"1600 ( ).

increasing drag coefficient. To relate drag coefficient to the lateral cover a re-
duction factor (Rlc ) responsible for the inter-sheltering effect with increasing den-
0·6
sity is needed. Regressive analyses show that the CD is closely related to Lc D\ s , so
0·6
Rlc"D\ s . The dimensionless parameter determined by the product of lateral cover
(Lc ) and the reduction factor (Rlc ) is termed effective lateral cover (Lec ). So
0·6
Lec"Lc Rlc , or Lec"Lc D\
s (26)
Regressing drag coefficients on effective lateral cover reached very good
correlation between them (Eqn (27), Fig. 7).
CD"0·0006619#15·0214L1·3830
ec R 2"0·97 (27)

Roughness length of vegetated surface

Roughness lengths of vegetated surfaces (Z0) were derived by using Eqn (16). Many
field observations show that roughness length of vegetated surface decreased with wind
velocity (Sutton, 1953), an effect which is ascribed to the bending of stems of plants

Table 4. The regression analysis results of the relationship between roughness


length and the height of standing residues
Significance
Ds (n m\2) a3 b3 r R2 level
25 0·000878 1·76e!10 6·9254 0·98 0·001
100 0·00123 7·70e!7 4·4385 1·00 0·001
200 !0·00324 0·000150 3·0000 0·99 0·001
400 !0·0151 0·000671 2·8059 0·99 0·001
800 !0·0212 0·000497 3·1535 1·00 0·001
1600 !0·0795 0·004514 2·4565 0·99 0·001
Fitting function: Z0"a3#b3h r.
DRAG COEFFICIENTS DISTURBED BY ARTIFICIAL STANDING VEGETATION 501

Table 5. The regression analysis results of the relationship between roughness


length and the density of standing residues
Significance
h (cm) a4 b4 d R2 level
2·5 0·000733 3·69e!9 2·1457 0·96 0·001
5·0 0·000973 1·97e!6 1·5000 0·98 0·001
7·5 !0·01216 0·00069 0·8998 0·99 0·001
10·0 !0·04607 0·00303 0·8310 0·98 0·001
12·5 !0·08057 0·00335 0·9035 0·99 0·001
15·0 0·1700 0·01570 0·7385 0·98 0·001
Fitting function: Z0"a4#b4D ds.

in the wind. But the wind tunnel results of this study indicate that, similar to drag
coefficients, the effect on Z0 was highly variable. Seven of the configurations
showed decrease, nine irregular changes, 10 increased and 13 showed almost no change
in Z0 with wind velocity. So the primary concern here focuses on comparing the
influence of vegetation structure on Z0 . The final roughness lengths were the mean
values of Z0 at different velocities. The results are shown in Fig. 8.
Wind tunnel results show that the Z0 of the bare surface free of any disturbance of
standing vegetation was about 0·00022 cm, at the same order as that estimated by
Bagnold (1941) for a smooth natural sandy surface. Z0 increased 3}25773-fold with the
addition of standing vegetation. Compared with drag coefficient, Z0 was much more
sensitive to the disturbance of vegetation. For example, Z0 of the vegetation of 15 cm
high and 2400 n m\2 in density reached 5·67 cm. Z0 also increased with both the density
and height of the standing vegetation. Their relationships (Tables 4 & 5) were obtained
by regressive analysis. Very good correlation existed between Z0 and vegetation density
and height. Tables 4 and 5 indicate that the relationship between Z0 and h and that
between Z0 and Ds could be expressed as power functions (Eqn (28) and (29))

Z0"a3#b3 hr (28)

Z0"a4#b4 Dds (29)

Figure 9. The relationship between roughness length and lateral cover.


502 Z. DONG ET AL.

Figure 10. The relationship between roughness length and height/spacing ratio.

The exponent r in Eqn (28) ranged from 2·46 to 6·93, averaging 3·80, while the
exponent d in Eqn (29) ranged from 0·74 to 2·15, in most cases (1, averaging 1·12.
This also implies that the effects of the height and density of vegetation on
roughness length were different. The reasons are similar to drag coefficient.
The correlation between drag coefficient and lateral cover, height/spacing ratio was
found to comply with power functions (Eqn (30), Eqn (31), Fig. 9, Fig. 10).

Z0"!0·04748#5·3346L1·2401
c R 2"0·90 (30)

Z0"!0·08737#0·1613Sh1·8117 R 2"0·96 (31)

Similar to drag coefficient, of the two parameters height/spacing is better to


characterize the effect of vegetation’s structure on roughness length.
It was found that the roughness length was more closely related to effective lateral
cover of standing vegetation (Eqn (32), Fig. 11).

Z0"!0·007496#182766·11L2·2545
ec R 2"0·97 (32)

Figure 11. The relationship between roughness length and effective lateral cover.
DRAG COEFFICIENTS DISTURBED BY ARTIFICIAL STANDING VEGETATION 503

Figure 12. The relationship between roughness length and drag coefficient.

Relationship between roughness length and drag coefficient

The relationship between roughness length and drag coefficient was established by
analysing all the wind tunnel data (Eqn (33)) and the correlation proved very good
(Fig. 12).
0·5
ln Z0"4·4414!0·4467C\ d R 2"0·99 (33)
Equation (32) can be rewritten as
CD"K 2/ln(Zc /Z00·9)2 (34)
where Zc"52 cm, is the centerline height of the tunnel. So Eqn (34) is close to Eqn (7).

Conclusions

The displacement height, drag coefficient and roughness length as influenced by the
disturbance of standing vegetation were modeled in wind tunnel by wooden dowels
of varying height and density. The adaptation of wind velocity profiles to standing
vegetation, drag coefficients and roughness lengths of vegetated surface were ana-
lysed. Conclusions reached are:
The wind velocity profile over the standing vegetation could be divided into three
sections, each section can be described by logarithmic function. All the three parameters
can be used to evaluate the effects of standing vegetation on near-surface airflow.
They reflect the same object from different angles. Displacement height is a signifi-
cant parameter for tall and dense vegetation, but its physical interpretation and calcu-
lation method are in need of further study. Although drag coefficient and roughness
length of vegetated surface should decrease with wind velocity, the tests in this study
didn’t show such regular changes. However, averageing the drag coefficient and
roughness length values at different wind velocities indicated good dependence of
these two dimensions on the structural parameters of standing vegetation. Both the
height and density of standing vegetation influenced drag coefficient and roughness
length, but their relative importance were different. Of the comprehensive struc-
tural parameters height/spacing ratio was better than the others were when the ef-
fects of standing vegetation on drag coefficient and roughness length were assessed.
The effective lateral cover, the product of lateral cover and a reduction factor
responsible for the inter-sheltering effect of increasing vegetation density could
express the structure of standing vegetation more precisely than the lateral cover used in
504 Z. DONG ET AL.

the previous literature to explain their effects on drag coefficient and roughness
length. Hence, height/spacing ratio or effective lateral cover is recommended for
assessing the effect of the structure of standing vegetation on wind erosion.
The good correlation between the derived roughness length and drag coefficient,
Eqn (34) and its reasonable agreement with Eqn (7) implies that obtaining drag
coefficient and roughness length of vegetated surfaces by a curve fit method is
reliable so long as caution is taken in selecting the appropriate measurements.
Of the three parameters roughness length is the most sensitive to characterize of the
effects of roughness elements on near-surface airflow because the estimated dis-
placement height becomes significant only for tall and dense vegetation. When the drag
coefficient of vegetated surfaces increases about 1·2}25 times the roughness length
increases almost 3}26000 times.

The authors appreciated the dedication of Charles Yates in preparing and calibrating the
experiment equipmenH t, Ace Berry and Jim Ed Jone helping do the experiment, Rhea Fryar
processing the data. We also extend our thanks to several anonymous reviewers for their critical
comments improving the manuscript. We are also due to Professor Q. Zhang, X.Y. Zou and
L.Y. Liu for their suggestions while revising the manuscript.

References

Abtew, W., Gregory, J.M. & Borreli, J. (1989). Wind profile: estimation of displacement height
and aerodynamic roughness. Transactions of the ASAE, 32: 521}527.
Arya, P.S. (1988). An Introduction to Micrometeorology. San Diego: Academic Press, Inc. 368 pp.
Bacche, P.H. (1986). Momentum transfer to plant canopies: influence of structure and variable
drag. Atmospheric Environment, 20: 1369}1378.
Bagnold, R.A. (1941). The Physics of Blown Sand and Desert Dunes. New York: William Morrow
& Company. 265 pp.
Buckley, R. (1987). The effect of sparse vegetation on the transport of dune sand by wind.
Nature, 325: 426}428.
Englehorn, C.L., Zingg, A.W. & Woodruff, N.P. (1952). The effect of plant residue cover
and clod structure on soil losses by wind. Soil Science Society America Proceedings, 16: 29}33.
Frank, A. & Kocurek, G. (1994). Effects of atmospheric conditions on wind profiles and
aeolian sand transport with an example from White Sand National Monument. Earth Surface
Processes and Landforms, 19: 735}745.
Fryrear, D.W. (1985). Soil cover and wind erosion. Transactions of the ASAE, 28: 781}784.
Fryrear, D.W., Saleh, A., Bilbro, J.D., Zobeck, T.M. & Stout, J.E. (1994). Field tested wind
erosion model. In: Buerkert, B., Allison, B.E. & von Oppen, M. (Eds), Proceedings of the
International Symposium ‘Wind Erosion in West Africa: The Problem and Its Control’ pp. 343}355.
Weikersheim, Germany: Margraft Verlag. 397 pp.
Garratt, J.R. (1977). Review of drag coefficients over oceans and continents. Monthly Weather
Review, 105: 915}929.
Garratt, J.R. (1980). Surface influences upon vertical profiles in the atmospheric near-surface
layer. Quarterly Journal of the Royal Meteorological Society, 106: 803}819.
Hagen, L.J. (1976). Windbreak design for optimum erosion control. In: Tinus, R.W. (Ed.),
Shelterbelts on the Great Plains, pp. 31}36. Great Plains Agricultural Publication no. 78.
Manhattan: Kansas. 293 pp.
Hagen, L.J. (1991). A wind erosion prediction system to meet the users’ need. Journal of Soil and
Water Conservation, 46: 106}111.
Hagen, L.J. & Skidmore, E.L. (1971). Windbreak drag as influenced by porosity. Transactions of
the ASAE, 14: 464}465.
Iversen, J.D. (1980). Drifting snow similitude—transport rate and roughness modeling. Journal of
Geology, 26: 393}403.
Kind, R.J. (1976). A critical examination of the requirements of for model simulation of
wind-induced erosion/deposition phenomena such as snow drifting. Atmospheric Environment,
10: 219}227.
DRAG COEFFICIENTS DISTURBED BY ARTIFICIAL STANDING VEGETATION 505

Lee, J.A. (1991). The role of desert shrub size and spacing on wind profile parameters. Physical
Geography, 12: 72}89.
Lee, J.A. (1991). Near surface wind flow around desert shrubs. Physical Geography, 12: 140}146.
Lee, B.E. & Soliman, B.F. (1977). An investigation of the forces on three dimensional bluff
bodies in rough wall turbulent layers. Transactions of the ASME Journal Fluids Engineering, 99:
503}510.
Lyles, L. & Allison, B.E. (1981). Equivalent wind-erosion protection from selected crop residues.
Transactions of the ASAE, 24: 405}408.
Marshall, J.K. (1970). Assessing the protective role of shrub-dominated rangeland vegetation
against soil erosion by wind. Proceedings of the 11th International Grassland Conference. Surfer’s
Paradise, Queensland, Australia. 19}23.
Marshall, J.K. (1971). Drag measurements in roughness arrays of varying density. Agricultural
Meteorology, 8: 269}292.
Molion, L.C.B. & Moore, C.J. (1983). Estimating the zero-plane displacement for tall vegetation
using a mass conservation method. Boundary-Layer Meteorology, 26: 115}125.
Musick, H.B. & Gillette, D.A. (1991). Field evaluation of relationships between vegetation
structure parameter and sheltering against wind erosion. Land Degradation and Rehabilitation, 2:
87}94.
Musick, H.B., Trujillo, S.M. & Truman, C.R. (1996). Wind tunnel modeling of the influence
of vegetation structure on saltation threshold. Earth Surface Processes and Landforms, 21:
589}605.
Pao, R.H.F. (1967). Fluid Mechanics. New York, London: John Wiley and Sons, Inc. 502 pp.
Raupach, M.R. (1979). Anomalies in flux-gradient relationships over forest. Boundary-Layer
Meteorology, 16: 467}486.
Schichting, H. (1936). Experimentelle Untersuchungen zum Rauhigkeitsproblem. Ingenieur
Archiv, 7: 1}34.
Schwartz, R.C., Fryrear, D.W., Harris, B.E., Bilbro, J.D. & R. Juo, A.S. (1995). Mean flow and
shear stress distributions as influenced by vegetative windbreak structure. Agricultural and Forest
Meteorology, 75: 1}22.
Schwartz, R.C., Fryrear, D.W. & Juo, A.S.R. (1997). Simulation of wind forces and erosion in
a field with windbreaks. Soil Science, 162: 372}381.
Siddoway, F.H., Chepil, W.S. & Armbrust, D.V. (1965). Effect of kind, amount, and
placement of residue on wind erosion control. Transactions of the ASAE, 8: 327}331.
Stockton, P.H. & Gillette, D.A. (1990). Field measurement of the sheltering effect of
vegetation on erodible land surface. Land Degradation and Rehabilitation, 2: 77}85.
Sutcliffe, R.C. (1936). Surface resistance in atmospheric flow. Quarterly Journal of Royal
Meteorological Society, 62: 3}14.
Sutton, O.G. (1953). Micrometeorology. New York: McGraw-Hill Book Company. 333 pp.
Taylor, G.I. (1916). Skin friction of the wind on the earth’s surface. Proceedings of the Royal Society
London, 92: 196}199.
Thom, A.S., Stewart, J.B., Oliver, H.R. & Gash, J.H.C. (1975). Comparison of aerodynamic and
energy budget estimates of fluxes over a pine forest. Quarterly Journal of Royal Meteorological
Society, 101: 93}105.
Unger, P.W. (1994). Managing Agricultural Residues. Boca Raton: Lewis Publishers. 289 pp.
Van den Ven, T.A.M., Fryrear, D.W. & Spaan, W.S. (1989). Vegetation characteristics and soil
loss by wind. Journal of Soil and Water Conservation, 44: 347}349.
Wasson, R.J. & Nanninga, P.M. (1986). Estimating wind transport of sand on vegetated surfaces.
Earth Surface Processes and Landforms, 11: 505}514.
Wolf, S.A. & Nickling, W.G. (1993). The protective role of sparse vegetation in wind erosion.
Progress in Physical Geography, 17: 50}68.
Wolf, S.A. & Nickling, W.G. (1996). Shear stress partition in sparsely vegetated desert canopies.
Earth Surface Processes and Landforms, 21: 607}619.
Woodruff, N.P. & Siddoway, F.H. (1965). A wind erosion equation. Proceedings of SSSA, 29:
602}608.

You might also like