Yeh 2010 CFD Shelter Shadow

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

J. Wind Eng. Ind. Aerodyn.

98 (2010) 520–532

Contents lists available at ScienceDirect

Journal of Wind Engineering


and Industrial Aerodynamics
journal homepage: www.elsevier.com/locate/jweia

An investigation into the sheltering performance of porous windbreaks under


various wind directions
Cheng-Peng Yeh a, Chien-Hsiung Tsai b, Ruey-Jen Yang a,n
a
Department of Engineering Science, National Cheng Kung University, Tainan, Taiwan
b
Department of Vehicle Engineering, National Pingtong University of Science and Technology, Pingtong, Taiwan

a r t i c l e in f o a b s t r a c t

Article history: This study performs a series of simulations utilizing the Navier–Stokes equations and the RNG k–e
Received 2 December 2008 turbulence model to investigate the efficacy of porous windbreaks in preventing the wind erosion of
Received in revised form stockpiles in an open storage yard. The simulations focus specifically on the effects of the fence porosity,
19 April 2010
the geometric configuration of the wind fence, and the direction of the incident wind. The basic validity
Accepted 23 April 2010
of the simulation model is confirmed by performing scale-model wind-tunnel experiments. In general,
Available online 15 May 2010
the results show that the dust control efficiency of the windbreak is fundamentally dependent on the
Keywords: direction of the incident wind. It addition, it is shown that a rectangular wind fence provides a poor
Windbreak sheltering effect for wind incident with an angle of 451, but is relatively more effective for winds
Wind erosion
incident in a normal direction. By contrast, an octagonal wind fence yields higher dust control efficiency
CFD
for oblique incident angles, but is less effective for normally incident winds. Finally, it is shown that the
RNG k–e turbulence model
shelter effect can be improved, via the deployment of additional wind fences within the storage yard or
at either end of the yard.
& 2010 Elsevier Ltd. All rights reserved.

1. Introduction fence with a porosity of ep ¼30% reduced the pressure fluctuation


in the downstream region of the fence by around 50% compared to
The wind erosion of stockpiles in an open storage yard is an a solid fence. However, neither study addressed the effect of the
important concern in the wind engineering domain. However, the resulting flow patterns on the sheltered stockpiles, and thus their
mechanisms responsible for wind erosion are highly complex and results were of little direct benefit in determining how best to
vary in accordance with differences in the local geographical, configure the windbreaks around a real-world storage yard. In an
meteorological, and stockpile surface conditions. Loredo-Souza and attempt to address this limitation, Lee et al. (Lee and Kim, 1999;
Schettini (2005) showed that the erosion of coal piles in an open Lee and Park, 2000, 2001; Lee et al., 2002; Park and Lee, 2003)
storage yard was the result of three specific mechanisms, namely performed a series of wind-tunnel experiments to examine the
creep, saltation, and suspension. Park and Lee (2002) reported that mutual interaction between the geometry of the stockpile, the
the dust emissions from open coal piles were directly related to the properties and position of the windbreak, and the conditions of
pressure fluctuations and shear stress acting on the surface of the the atmospheric boundary layer in determining the wind erosion
pile, and could be significantly reduced by arranging porous wind characteristics of open coal and sand piles. Santiago et al. (2007)
fences in such a way that the turbulence intensity of the wind was verified the performance of various turbulence models by
reduced before passing over the pile. performing wind tunnel experiments and showed that a wind
Many previous studies have shown that the intensity of the fence with a porosity of ep ¼30–40% yielded an effective reduction
turbulent flow behind a wind fence is dependent upon the fence in the pressure fluctuations acting on the surface of a stockpile
porosity, ep, i.e. the ratio of the aggregated area of the holes in the with a prism-like geometry. Lee and Lim (2001) utilized the RNG
fence to the total surface area of the fence. Castro (1971) (renormalization group) k–e turbulence model to examine the
performed two-dimensional simulations to investigate the wake flow characteristics around a 2-D triangular-shaped prism located
characteristics of perforated plates positioned normal to an air- behind a porous wind fence as a function of the porosity, height
stream and showed that the recirculation zone behind the plates and position of the fence. Their results showed that a porosity of
disappeared when the porosity exceeded a value of ep ¼30%. around 30–50% resulted in the maximum attenuation of the mean
Similarly, Yaragal et al. (1997) showed experimentally that a wind pressure acting on the prism surface.
However, doubts remains over the ability of small-scale
simulation models to accurately reproduce the wind erosion
n
Corresponding author. Tel.: + 886 6 2002724; fax: + 886 6 2766549. mechanisms observed in the real world (Sakamoto et al., 2001;
E-mail address: rjyang@mail.ncku.edu.tw (R.-J. Yang). Obasaju and Robins, 1998). Consequently, Meroney (2006) and

0167-6105/$ - see front matter & 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.jweia.2010.04.002
C.-P. Yeh et al. / J. Wind Eng. Ind. Aerodyn. 98 (2010) 520–532 521

Torano et al. (2006) argued that while small-scale simulations Einstein summation convention as follows:
provide a useful means of validating the experimental results  
@ðrkÞ @ðrkÞ @ mt @k
obtained using a small scale model, obtaining accurate insights þui ¼ þGre ð3Þ
@t @xi @xi sk @xi
into the real-world wind erosion phenomenon requires the use of
full-scale simulation models. In addition to the scaling problem,  
the majority of the windbreak studies presented in the literature @ðreÞ @ðreÞ @ mt @e e e2
þ ui ¼ þ C1e GC2e r ð4Þ
consider the simplified case in which the wind blows in one @t @xi @xi se @xi k k
direction only. Bourdin and Wilson (2007) presented computa-
where k is the turbulence kinetic energy, e is the kinetic energy
tional fluid dynamics (CFD) simulations which showed that the
dissipation rate, sk is the Prandtl number of the turbulence kinetic
nature of the flow structures formed behind windbreaks is
energy, and se is the Prandtl number of the turbulence kinetic
fundamentally dependent on the direction of the wind. However,
energy dissipation rate. The remaining variables in Eqs. (3) and (4)
the simulations did not consider the stockpiles themselves, and
are defined as follows:
therefore definitive conclusions regarding the optimal configura-
tion of the windbreaks relative to the stockpiles could not be G ¼ 2mt Sij Sij , ð5Þ
obtained.
In an attempt to address the various limitations described C2e ¼ C2e þ C20 e , ð6Þ
above, this study conducts full-scale CFD simulations using the
Reynolds averaged Navier–Stokes equations and the RNG k–e
Cm rZ3 ð1Z=Z0 Þ
turbulence model to investigate the efficacy of porous windbreaks C02e ¼ , ð7Þ
1 þ bZ3
in suppressing the wind erosion of stockpiles in an open storage
yard. The simulations commenced by investigating the correlation
between the porosity of the windbreak and the wind velocity at k2
mt ¼ rCm , ð8Þ
which particles are first eroded from the stockpile. The validity of e
the simulation model is confirmed by comparing the numerical
results with the observations obtained from scale-model wind k
Z¼S , ð9Þ
tunnel experiments. A series of simulations are then performed to e
investigate the shelter effect provided by wind fences with qffiffiffiffiffiffiffiffiffiffiffiffi
different geometrical configurations under various normal and S¼ 2Sij Sij , ð10Þ
oblique incident wind directions.
where Sij is the shearing-rate tensor. The coefficients of the
model are the default values proposed by Fluent (2006), i.e.,
Cu ¼0.085, se ¼0.719, sk ¼0.719, C1e ¼ 1.42, C2e ¼1.68, b ¼0.012,
2. Model description and Z0 ¼4.38.

2.1. Governing equation


2.2. Porous media model
2.1.1. 3-D averaged Navier–Stokes equations
In modeling the flow field characteristics within the storage In the simulations, the flow of the wind through the windbreak
yard, the present simulations make two fundamental assump- is modeled in accordance with Darcy’s law. That is, an additional
tions: (1) the existence of a neutral atmospheric boundary layer momentum source term is added to the standard fluid flow
and (2) the fluid flow satisfies the continuity equations and equations. The source term comprises an inertial loss term
3-D Reynolds averaged Navier–Stokes momentum equations. (Idel’chik, 1986).
In addition, the model is simplified by assuming that a non-slip 1
boundary condition and the fluid flows in an incompressible Dp ¼ C2 r9uj 9uj ð11Þ
2
manner. The governing equations can therefore be written as
follows: where C2 (E z/l) is the inertial resistance factor and can be
Continuity equation: computed from empirical equations as follows:
 
@r @ ro1 2 l 1
þ ðrui Þ ¼ 0, ð1Þ z1KB ¼ Dp= ¼ z0 þ l , ð12Þ
@t @xi 2 dr ep 2

where r is the density of the fluid and ui denotes the fluid velocity
ro21 1
in the ith direction. z ¼ Dp ¼ zj þ e0Re z1KB , ð13Þ
2 e2p
Momentum equation:
  where z0 is the contrast constant (Table 1) (Idel’chik, 1986),
@ @ @p @ @u @  
ðrui Þ þ ðrui uj Þ ¼  þ m i þ ru0i u0j , ð2Þ l is the thickness, dr is the hydraulic diameter, l is the friction
@t @xj @xi @xj @xj @xj
coefficient ( ¼0.3164/Re0.25), ep is the porosity, and z is the correct
where p is the static pressure and ru0i u0j is the Reynolds stress value by the Reynolds number (Table 2) (Idel’chik, 1986).
caused by turbulence.
Table 1
Corresponding value by different porosity.

2.1.2. Turbulence model ep 0 0.1 0.2 0.3 0.4 0.5


In the present simulations, the turbulent flow is modeled using z0 1.71 1.67 1.63 1.59 1.55 1.50
the RNG k–e turbulence model because of its good prediction of
ep 0.6 0.7 0.8 0.9 1.0
complex flows (Choudhury, 1993). The complete formulation z0 1.45 1.39 1.32 1.22 1.00
of the RNG k–e turbulence model (Fluent Inc., 2006) is given in
522 C.-P. Yeh et al. / J. Wind Eng. Ind. Aerodyn. 98 (2010) 520–532

Table 2
Corresponding value by different Reynolds number.

Re
ep
25 40 60 102 2  102 4  102 103
e0Re

0.34 0.36 0.37 0.40 0.42 0.46 0.53


zj

0 1.94 1.38 1.14 0.89 0.69 0.64 0.39


0.2 1.78 1.36 1.05 0.85 0.67 0.57 0.36
0.3 1.57 1.16 0.88 0.75 0.57 0.43 0.30
0.4 1.35 0.99 0.79 0.57 0.40 0.28 0.19
0.5 1.10 0.75 0.55 0.34 0.19 0.12 0.07
0.6 0.85 0.56 0.30 0.19 0.10 0.06 0.03
0.7 0.58 0.37 0.23 0.11 0.06 0.03 0.02
0.8 0.40 0.24 0.13 0.06 0.03 0.02 0.01
0.9 0.20 0.13 0.08 0.03 0.01 0 0
0.95 0.03 0.03 0.02 0 0 0 0

Re
ep
2  103 4  103 104 2  104 105 2  105 106
e0Re

0.59 0.64 0.74 0.81 0.94 0.96 0.98


zj

0 0.30 0.22 0.15 0.11 0.04 0.01 0


0.2 0.26 0.20 0.13 0.09 0.03 0.01 0
0.3 0.22 0.17 0.10 0.07 0.03 0.01 0
0.4 0.14 0.10 0.06 0.04 0.02 0.01 0
0.5 0.05 0.03 0.02 0.01 0.01 0.01 0
0.6 0.02 0.01 0.01 0 0 0 0
0.7 0.01 0 0 0 0 0 0
0.8 0 0 0 0 0 0 0
0.9 0 0 0 0 0 0 0
0.95 0 0 0 0 0 0 0

2.3. Numerical method

The simulations were performed using FLUENT CFD software.


In the solution procedure, the differential equations were
discretized using the finite volume method and the pressure
and velocity fields were decoupled using the semi-implicit
method for pressure-linked equation (SIMPLE) (Patankar, 1980).

2.4. Boundary conditions

2.4.1. Atmospheric boundary layer


Generally speaking, the atmospheric boundary layer within a
height of 200 m of the Earth’s surface is called the atmospheric
surface layer and has a virtually constant shear stress. Thus, in the
present simulations, the input profile of the velocity and
turbulence (k and e) conform to the following neutral atmospheric
boundary layer distribution (Richards and Hoxey, 1993):
 
u z þz0
UðzÞ ¼ ln , ð14Þ
K z0
 n Fig. 1. Computational domain of 3-D storage yard model.
z
UðzÞ ¼ Uref , ð15Þ
zref

ðu Þ2
kðzÞ ¼ pffiffiffiffiffiffi , ð16Þ length ( ¼0.01 m, which is equivalent to a roughness height of
Cm
approximately 0.3 m of the suburb), Cm is a model constant of the
standard k–e model ( ¼0.09), k is the turbulent kinetic energy, and
ðu Þ3
eðzÞ ¼ , ð17Þ e is the turbulent dissipation rate. Note that Eq. (14) is equivalent
Kðz þ z0 Þ
to the power law profile governing the velocity in the atmospheric
where U is the velocity, un is the friction velocity, K denotes the boundary layer (shown in Eq. (15)), for which n has a value of
von-Karman constant ( ¼0.41), z is the height, z0 is the roughness approximately 0.14.
C.-P. Yeh et al. / J. Wind Eng. Ind. Aerodyn. 98 (2010) 520–532 523

2.4.2. Computational domain domain. A depth of 8H will lead to the blockage ratio r4%, and is
Fig. 1 presents a schematic illustration of the computational consistent with that suggested by Franke et al. (2004).
domain for the 3-D full-scale storage yard model used in the
present simulations. As shown, the storage yard is assumed to 2.5. Grid system
contain five parallel rows of prism-like stockpiles interrupted at
their mid-length positions by a central aisle. Appropriate The computational domain was meshed using an orthogonal
conditions are specified at boundary cells to correctly reflect grid system in order to accurately reproduce the boundary layer
the physical phenomena of the flow field. At the upstream section, phenomenon near the ground. Furthermore, as shown in Fig. 3,
the neutral atmospheric boundary layer velocity profile is the mesh was constructed using a non-uniform grid size in order
imposed. A reference pressure for the outlet at the downstream to achieve a satisfactory trade-off between the computational cost
section of the computational domain is specified. On the and the need to capture the dramatic variations in the flow field in
boundary at the top and the lateral of the domain, a zero-gra- the vicinity of the windbreaks and stockpile surface. The
dient in the direction normal to the boundary for pressure is computational grids keep expansion ratio g r1.5 in regions of
assumed. For solid boundaries, a non-slip condition is used. In high velocity gradients, and the maximum aspect ratio of the
addition, the wall boundaries are assigned to the terrain surface, grids are less than 50. The height of first cell adjacent to the wall
which is applied through the standard wall functions imple- (bottom of domain) is 2.5 times of the roughness height (Franke
mented in Fluent 6.3.26 (2006). The wall function definition plays et al., 2004). In total, the grid comprised around two million
an important role that should be compatible with inlet velocity, individual cells.
turbulent kinetic energy, and turbulence dissipation rate.
Otherwise, the velocity profile will be immediately affected by
2.6. Numerical validation
the incorrect wall boundary. Recently, this compatibility problem
has been addressed by Blocken et al. (2007). As suggested,
To our knowledge, this paper maybe the first study to simulate
alleviating the requirement yp 4ks can partially solve this
the flow around a porous wind fence and flow interacting with
problem. Here yp is a distance from the center point P of the
stockpiles embedded in an atmospheric layer. Hence, in order to
wall-adjacent cell to the wall (bottom of domain), and ks is the
validate the numerical method and the possible applicability of
height of physical roughness. Fig. 2 indicates the principal real-
the turbulence models used later for simulating these cases in this
world dimensions of interest in the model, namely the height and
study, we choose the Bradley and Mulhearn (1983) data as a
width of each stockpile, the separation distance between
benchmark. They measured field data of velocity and stress
neighboring stockpiles, the distance between the final stockpile
perturbations behind a shelter fence (height¼1.2 m) in a neutral
and the fence, and the height of the fence, respectively. Note that
atmospheric boundary layer. In the comparison, streamwise
for full-scale storage yard model, the depth of the computational
velocity component is normalized with the average wind speed
domain is 8H (¼ 0.15 L). To consider both accuracy and efficiency
for our numerical calculations, several numerical experiments
have been carried out to ensure the erosion potential calculated in
this paper do not depend on the choice of the computational

Fig. 2. Schematic illustration showing critical dimensions in full-scale simulation Fig. 4. Variations for horizontal profiles of streamwise mean velocity (U/U04) at z/
model. h¼0.38, 1.88.

Fig. 3. Body-fit grid system used in present simulations.


524 C.-P. Yeh et al. / J. Wind Eng. Ind. Aerodyn. 98 (2010) 520–532

of oncoming upstream flow at height z ¼4 m (U04). Normalized As shown in Eq. (18), the emission factor varies as a function of
velocities at fixed heights of z/h¼0.38 and 1.88 are shown in a constant m and the wind erosion potential Pi. In practice, Pi
Fig. 4. Computed normalized velocity for the two turbulence represents the rate of emission of particles from the stockpile. For
models are qualitative in good agreement with the experimental instance, to reduce Pi by 10% means to decrease particle emission
data. Overall, the numerical results of RNG k–e model (present rate by 10%, in other words, the dust control efficiency improves
model) are more accurate than the standard k–e model in terms of by 10%. Thus, in the current simulations the aim is to determine
the variations of the flow field. the windbreak porosity and windbreak configuration, which
minimize the value of Pi. Note that in this study, the dust control
efficiency of the various windbreaks is evaluated via a comparison
2.7. Evaluation of windbreak efficacy with the scenario in which no fence is deployed.

In the present simulations, the efficacy of the various wind


fences was evaluated in terms of two performance indicators, 2.8. Experimental apparatus and methods
namely the wind erosion potential and the dust control efficiency.
According to Section 13.2.5 of the AP-42 Compilation of Air The accuracy of the simulation model in predicting the wind
Pollutant Emission Factors published by the U.S. Environmental erosion effect was validated by performing an initial set of
Protection Agency (EPA), the wind erosion emissions from a experimental trials in a closed-return type wind-tunnel with a
stockpile can be estimated by the following emission factor: test section of 0.55  0.55 m2 (width  height). As shown in Fig. 5,
the experimental apparatus included a green laser with an output
X
N
power of 100 mW (GM-XP01-100), a single-lens reflex camera
Emission factor ¼ m Pi , ð18Þ
i¼1
(Canon, 350D), a DV video recorder (SONY, DCR-PC115), a VCR
and a PC. In the experiments, a 1/300 scale model of a sand
where m is a particle size multiplier, N is the number of pile was constructed within the wind tunnel and the critical
disturbances per year, and Pi is the erosion potential (g/m2). threshold velocity (Vc), i.e. the velocity at which the particles
The wind erosion potential, P, is defined as first begin to drift, was identified by gradually increasing the
velocity whilst continuously observing the image of the sand pile
P ¼ 58ðu ut Þ2 þ 25ðu ut Þ, ð19Þ
captured by the video recorder. In order to establish the cor-
relation between the windbreak porosity and the critical
P¼0 for u r ut ,
threshold velocity, experiments were performed for three
where un is the friction velocity (m/s) on the surface of the storage different sheltering conditions, namely (1) no windbreak (i.e.
pile and ut is the threshold friction velocity (m/s) and is ep ¼100%), (2) a windbreak with a porosity of ep ¼30%, and (3) a
determined experimentally. The friction velocity is given by solid windbreak (ep ¼0%).
pffiffiffiffiffiffiffiffiffiffiffi
u ¼ tw =r ð20Þ
In evaluating the efficacy of the windbreaks considered in this 3. Results and discussion
study, the simulations adopt reference value of ut ¼0 m/s for the
most conservative estimation, and denote the corresponding 3.1. Wind-tunnel experiments and numerical validations for the
values of the wind erosion potential as P0. Note that while some 1/300th geometry-scale
researchers have used a shear stress performance indicator to
evaluate the windbreak efficiency, this approach is similar to the The qualitative analysis of the simulation model in predicting
use of P0 in the present simulations since P0 ¼58 un2 +25 un. the wind erosion effect is validated by comparing the numerical

Fig. 5. Schematic illustration showing experimental setup for particle motion visualization and threshold velocity measurement.
C.-P. Yeh et al. / J. Wind Eng. Ind. Aerodyn. 98 (2010) 520–532 525

results with the same scale-model used in wind-tunnel experi- results for the case in which the stockpile is sheltered behind a solid
ment. The scale-model is a 1/300 scaled-down of the original coal fence (i.e. ep ¼0%). In this case, it can be seen that particle saltation
pile model. The computational domain for this case adopts the occurs principally on the windward surface of the pile due to the
test section size of the wind-tunnel in experiment. For solid formation of a vortex between the fence and the sand pile. In
boundaries, a non-slip condition is used. At the inlet section, the addition, the experimental results show that the introduction of the
uniform flow with a constant velocity (U¼2.4, 3.6, and 7.1 m/s, for solid windbreak increases the threshold velocity to a value of around
three simulating cases) is imposed. 3.6 m/s. Fig. 8 presents the simulation and experimental results
Fig. 6 presents the simulation and experimental results obtained obtained when the stockpile is shielded using a windbreak with a
for the test-tunnel experiment performed without using a windbreak porosity of ep ¼30%. Both sets of results show that particle saltation is
(i.e. ep ¼100%). The simulation results presented in Fig. 6a show the therefore confined primarily to the leeward surface of the stockpile.
streamline patterns around the stockpile and the distribution of the Finally, the wind-tunnel results indicate that the introduction of the
wind erosion potential on the surface of the stockpile. The results porous windbreak increases the critical velocity of the particles to
clearly show that the majority of the airborne particles originate from almost 7.1 m/s. Overall, the experimental images presented in
the top of the stockpile. This observation is readily confirmed in the Figs. 6–8 show that the use of a windbreak with a porosity of
experimental image presented in Fig. 6b. According to the wind- ep ¼30% yields a higher critical particle velocity than either a solid
tunnel results, the corresponding critical threshold velocity has a windbreak or no windbreak at all, and is therefore more effective in
value of 2.4 m/s. Fig. 7 presents the simulation and experimental suppressing the wind erosion of the stockpile.

Fig. 6. Motion of sand particles without wind fence: (a) numerical simulation results (U ¼ 2.4 m/s) and (b) wind-tunnel visualization results (Vc ¼2.4 m/s).
526 C.-P. Yeh et al. / J. Wind Eng. Ind. Aerodyn. 98 (2010) 520–532

Fig. 7. Motion of sand particles behind solid wind fence (ep ¼0%): (a) numerical simulation results (U ¼3.6 m/s) and (b) wind-tunnel visualization results (Vc ¼ 3.6 m/s).

Figs. 6–8 demonstrate that an excellent qualitative agreement which maximizes the shelter afforded to the stockpiles, it is not
exists between the numerical and experimental results. Further- enough to consider one wind direction alone. A further series of
more, it is evident that the regions of particle saltation shown in simulations for the full-scale model were performed in which the
the experimental images coincide with the regions of higher wind stockyard was sheltered using wind fences with various geometrical
erosion potential shown in the simulation results. Thus, the configurations and the wind was assumed to approach the yard from
overall validity of the simulation model is confirmed. five different directions, namely normally (i.e. y ¼01 and 901) and
obliquely (i.e. y ¼ 22.51, 451, and 67.51). (Note that y is measured with
respect to the longitudinal axis of the stockpiles of the windward side
3.2. Efficacy of wind fences in suppressing wind erosion of stockpile of the shelterbelt.)
under different wind directions
3.2.1. Efficacy of rectangular wind fence
The results presented above have considered the wind to be The simulations commenced by considering the shelter effect
incident on the stockpile in one direction only for the wind-tunnel provided by a conventional rectangular wind fence around the
test model. However, in practice, the stockpiles within an open perimeter of the storage yard. Fig. 9 presents the distribution of
storage yard are exposed to wind from a variety of directions the wind erosion potential on the stockpile surface for each of the
(depending on the local environmental conditions, the time of year, five considered wind directions. Note that in performing the
and so on). Accordingly, in searching for the wind fence configuration simulations, the fence was assumed to have a height of x¼20 m
C.-P. Yeh et al. / J. Wind Eng. Ind. Aerodyn. 98 (2010) 520–532 527

Fig. 8. Motion of sand particles behind wind fence with porosity ep ¼30%: (a) numerical simulation results (U¼ 7.1 m/s) and (b) wind-tunnel visualization results
(Vc ¼ 7.1 m/s).

and a porosity of ep ¼30%. Note also that the pink numerals in the 3.2.1.1. Flow-field phenomena associated with corner-flow
table indicate the dust control efficiency, while the black (y ¼451). The results presented in Fig. 9 indicate that the rec-
numerals indicate the average wind erosion potential, i.e. tangular wind fence yields a particularly poor shelter effect when
R the wind is incident with an angle of y ¼451. To clarify this result,
P0 dA
P0 ¼ A , ð21Þ two further simulations were performed to examine the char-
A
acteristics of the flow field induced in the storage yard under the
where A denotes the surface area of the stockpile. effects of a corner flow. Fig. 10 illustrates the velocity vector
Overall, the results show that the rectangular wind fence contours at a height of 10 m above the ground and shows that the
provides a relatively poor sheltering effect. For example, when the wind fence essentially protects only those stockpiles immediately
wind is incident with an oblique angle of y ¼22.51 or 451, the behind the fence. As a result, no more than a minor improvement
average wind erosion potential has a value of 64.8 and 76.6 g/m2, in the dust control efficiency is obtained (i.e. 17%). Fig. 11 presents
respectively. Nonetheless, the pink numerals in the table indicate the corresponding surface streamline patterns at various positions
that the rectangular fence provides at least some improvement in of the storage yard. Since the wind fence has a porosity of ep ¼30%,
the dust control efficiency relative to that achieved when no fence flow vortices are induced behind the fence for a section of
is used, e.g. an improvement of 56% for the case of wind incident distances. The vortex structures extend and enlarge gradually
with an angle of y ¼901, 36% for winds incident with an angle of along the length of the windbreak. After the corner-wind passes
y ¼01 or 67.51, and so on. over the fence, it crosses the vortex behind the fence and
528 C.-P. Yeh et al. / J. Wind Eng. Ind. Aerodyn. 98 (2010) 520–532

Fig. 9. Distribution of P0 on stockpile surface under different wind directions for rectangular wind fence (U10 ¼8 m/s, xfence ¼ 20 m, ep ¼ 30%).

However, its performance is notably poorer than that of the


rectangular wind fence for normally incident wind. The wind
erosion for winds incident with an angle of y ¼ 01 is serious
because the length of the lateral fence of the octagonal
configuration is much shorter than the rectangular form. Thus,
further simulations were performed in which additional fences
were added to the octagonal fence in either the perpendicular or
parallel direction of the storage yard.

3.2.3. Efficacy of octagonal wind fence with additional mid-fences in


perpendicular direction
Fig. 13 illustrates the distribution of the wind erosion potential
on the surface of the stockpiles under various wind directions
when the storage yard is sheltered using an octagonal fence with
two additional fences installed in the perpendicular direction in the
middle of the yard. From inspection of the data presented in the
table, it is seen that the addition of the two perpendicular fences
yields a reduction in the average wind erosion potential for all
Fig. 10. Velocity vector fields at height 10m above ground for wind direction of directions of the incident wind. In particular, it is observed that the
y ¼ 451 and rectangular wind fence (U10 ¼ 8 m/s, xfence ¼20m, ep ¼30%). additional fences greatly improve the shelter effect when the wind
is incident with an angle of y ¼01. As a result of the lower wind
erosion potential, the dust control efficiency is improved to 49%
congregates at the middle of the stockpile. Therefore, high wind
(y ¼901), 48% (y ¼67.51), 43% (y ¼451), 48% (y ¼ 22.51), and 42%
erosion potential is produced on the surface of the stockpile, and
(y ¼01), respectively, which is better than that achieved by the
thus poor dust control efficiency is obtained.
conventional rectangular wind fence for all wind directions other
than y ¼901. Thus, the additional fences yield a significant
improvement in the protection afforded to the stockpiles within
3.2.2. Efficacy of octagonal wind fence
the open yard. However, coal transportation vehicles generally
A further series of simulations were performed in which the
operate in a direction parallel to that of the stockpiles in CSC (China
storage yard was sheltered using an octagonal wind fence, as
Steel Corporation), and thus the use of perpendicular fences is
shown in Fig. 12. The table at the foot of the figure compares the
somewhat impractical from an operational perspective. Therefore,
wind erosion potential and dust control efficiency characteristics
another two configurations of material yard are proposed to satisfy
of the rectangular wind fence (upper row) with those of the
the requirement of CSC. The efficiency of new proposed
octagonal fence (lower row). Comparing the two sets of results, it
configurations is described in the next two sections.
is evident that for all wind incident angles other than y ¼ 01 or 901,
the octagonal wind fence results in both a lower wind erosion
potential and a higher dust control efficiency. In other words, the 3.2.4. Efficacy of octagonal wind fence with additional mid-fences in
octagonal wind fence is more effective than the conventional parallel direction
rectangular fence in sheltering the stockpiles within the storage Fig. 14 illustrates the wind erosion potential distribution on the
yard from the erosive effects of an obliquely incident wind. stockpile surface for the case where the storage yard is protected by
C.-P. Yeh et al. / J. Wind Eng. Ind. Aerodyn. 98 (2010) 520–532 529

Fig. 11. Surface streamline patterns at different positions of storage yard for wind direction of y ¼ 451 and rectangular wind fence (U10 ¼8 m/s, xfence ¼ 20 m, ep ¼ 30%).

Fig. 12. Distribution of P0 on stockpile surface under different wind directions for octagonal wind fence (U10 ¼8 m/s, xfence ¼ 20 m, ep ¼ 30%).

an octagonal wind fence around its perimeter and two additional 3.2.5. Efficacy of octagonal wind fence with additional mid-fences
fences deployed in the parallel direction between the 3rd and 4th and end-fences
stockpile rows. The results indicate that the introduction of the two In an attempt to improve the relatively poor sheltering effect
fences yields an effective improvement in the dust control efficiency provided by the octagonal windbreak configuration with parallel
of the windbreak configuration for wind with an incident angle of mid-fences in the case of wind with an incident angle of 01, a final
y ¼22.5–901 compared to that achieved using the octagonal fence configuration was considered in which additional fences were
alone. However, the configuration provides a poor sheltering effect added not only in the parallel direction within the storage yard, but
when the wind is incident on the stockyard with an angle of 01 (i.e. a also in the perpendicular direction at the two ends of the yard (see
dust control efficiency of just 20%). Under this case, the main reason Fig. 15). The results presented in the lower row of the table indicate
for the high erosion potential is that the mid-fences are parallel to the that the additional end fences improve the dust control efficiency
wind direction. Furthermore, comparing Figs. 13 and 14, it is evident from 20% to 35% for wind with an incident angle of 01, whilst
that while the configuration with parallel mid-fences represents an retaining a high degree of protection against the erosive effects of
improvement over that with perpendicular fences from an wind with an incident angle of 901. However, it is also observed
operational sense. Finally, in a separate series of simulations, it was that the additional fences prompt a notable degradation in the dust
determined that the dust control efficiency values obtained when the control efficiency for wind incident with an angle of y ¼ 22.51.
parallel fences were positioned between the second and third
stockpiles rows were identical to those obtained when deploying
the fences between the third and fourth rows. Thus, it can be inferred 4. Conclusions
that the wind fence configuration shown in Fig. 14 is equally effective
for the case in which the wind reverses and blows in the opposite The effectiveness of porous wind fences in preventing wind
direction to that shown in the figure. erosion has been investigated numerically utilizing CFD simulations
530 C.-P. Yeh et al. / J. Wind Eng. Ind. Aerodyn. 98 (2010) 520–532

Fig. 13. Distribution of P0 on stockpile surface under different wind directions for octagonal wind fence with additional vertical fences in center of storage yard (U10 ¼8
m/s, xfence ¼20 m, ep ¼30%).

Fig. 14. Distribution of P0 on stockpile surface under different wind directions for octagonal wind fence with additional parallel fences between 3rd and 4th stockpile rows
(U10 ¼8 m/s, xfence ¼ 20 m, ep ¼ 30%).
C.-P. Yeh et al. / J. Wind Eng. Ind. Aerodyn. 98 (2010) 520–532 531

Fig. 15. Distribution of P0 on stockpile surface under different wind directions for octagonal wind fence with parallel fences between 3rd and 4th stockpile rows and
additional vertical end fences (U10 ¼ 8 m/s, xfence ¼20m, ep ¼30%).

based on the Navier–Stokes equations and the RNG k–e turbulence References
model. It has been shown that the simulation results are in good
agreement with those obtained from wind-tunnel experiments. The ANSYS Fluent 6.3.26, Fluent Inc., 2006. Lebanon, NH, USA.
results have shown that a wind fence porosity of ep ¼30% yields an Blocken, B., Stathopoulos, T., Carmeliet, J., 2007. CFD simulation of the atmospheric bou-
effective reduction in the wind erosion potential acting on the ndary layer–wall function problems. Atmospheric Environment 41 (2), 238–252.
Bourdin, P., Wilson, J.D., 2007. Windbreak aerodynamics: is computational fluid
surface of coal piles in an open storage yard. Moreover, it has been dynamics reliable? Boundary-Layer Meteorology 126 (2) 181–208.
shown that the dust control efficiency of porous windbreak Bradley, E.F., Mulhearn, P.J., 1983. Development of velocity and shear stress
structures is significantly dependent upon the direction of the distributions in the wake of a porous shelter fence. Journal of Wind
Engineering and Industrial Aerodynamics 15 (1-3), 145–156.
incident wind. The simulation results have revealed that the Castro, I.P., 1971. Wake characteristics of two-dimensional perforated plates
traditional rectangular wind fence configuration protects stockpiles normal to an air-stream. Journal of Fluid Mechanics 46 (3), 599–609.
from the erosive effects of wind incident with a normal angle, but Choudhury, D., 1993. Introduction to the renormalization group method and
turbulence modeling. Technical Memorandum TM 107. Fluent Inc., Lebanon NH.
has extremely poor dust control efficiency for wind incident with an
Franke, J., Hirsch, C., Jensen, A.G., Krüs, H.W., Miles, S.D., Schatzmann, M.,
angle of 451. By contrast, an octagonal wind fence improves the Westbury, P.S., Wisse, J.A., Wright, N.G., 2004. Recommendations on the use
shelter effect when the wind is incident with an oblique angle. It has of CFD in wind engineering. Urban Wind Engineering and Building Aero-
been shown that the shelter effect of the octagonal wind fence can dynamics, von Karman Institute, Sint-Genesius-Rode, Belgium, May 5–7.
Idel’chik, I.E., 1986. Erwin Fried, Handbook of Hydraulic Resistance 2rd ed.
be improved for all values of the wind incident angle via the addition Springer-Verlag.
of perpendicular or parallel windbreak structures within the Lee, S.J., Kim, H.B., 1999. Laboratory measurements of velocity and turbulence field
confines of the storage yard. However, perpendicular mid-fences behind porous fences. Journal of Wind Engineering and Industrial Aerody-
namics 80 (3), 311–326.
are generally impractical from an operational point of view, while
Lee, S.J., Park, C.W., 2000. The shelter effect of porous wind fences on coal piles in
parallel fences yield only a marginal improvement in the shelter POSCO open storage yard. Journal of Wind Engineering and Industrial
effect afforded to the stockpiles from wind incident with an angle of Aerodynamics 84 (1), 101–118.
01. Thus, a final configuration has been considered in which Lee, S.J., Park, C.W., 2001. The effects of a bottom gap and non-uniform porosity in a
wind fence on the surface pressure of a triangular prism located behind the fence.
perpendicular fences are added to the two ends of the octagonal Journal of Wind Engineering and Industrial Aerodynamics 89 (13), 1137–1154.
wind fence in addition to the two parallel fences within the storage Lee, S.J., Park, K.C., Park, C.W., 2002. Wind tunnel observations about the shelter
yard. The results have shown that this windbreak configuration effect of porous fences on the sand particle movements. Atmospheric
Environment 36 (9), 1453–1463.
provides a high degree of protection against the erosive effects of
Lee, S.J., Lim, H.C., 2001. A numerical study on flow around a triangular
normally incident wind, but is less effective in sheltering the prism located behind a porous fence. Fluid Dynamics Research 28 (3),
stockpiles from wind incident with an angle of y ¼22.51. 209–221.
532 C.-P. Yeh et al. / J. Wind Eng. Ind. Aerodyn. 98 (2010) 520–532

Loredo-Souza, A.M., Schettini, E.B.C., 2005. Wind tunnel studies on the shelter Richards, P.J., Hoxey, R.P., 1993. Appropriate boundary conditions for computa-
effect of porous fences on coal piles models of the CVRD-Vitóia, Brazil. In: tional wind engineering models using the k–e turbulence model. Journal of
Proceedings of the A&WMA’s 98th Annual Conference & Exhibition—Exploring Wind Engineering and Industrial Aerodynamics 46&47, 145–153.
Innovative Solutions, Minneapolis, Minnesota, USA, June 21–24. Sakamoto, H., Moriya, M., Takai, K., Obata, Y., 2001. Development of a new type
Meroney, R.N., 2006. CFD prediction of cooling tower drift. Journal of Wind snow fence with airfoil snow plates to prevent blowing-snow disasters: Part 1,
Engineering and Industrial Aerodynamics 94 (6), 463–490. evaluation of performance by blowing-snow simulation in a wind tunnel.
Obasaju, E.D., Robins, A.G., 1998. Simulation of pollution dispersion using small Journal of Natural Disaster Science 23 (1), 1–11.
scale physical models-an assessment of scaling options. Environmental Santiago, J.L., Martı́n, F., Cuerva, A., Bezdenejnykh, N., Sanz-Andrés, A., 2007.
Monitoring and Assessment 52 (1–2), 239–254. Experimental and numerical study of wind flow behind windbreaks. Atmo-
Park, C.W., Lee, S.J., 2002. Verification of the shelter effect of a windbreak on coal spheric Environment 41 (30), 6406–6420.
piles in the POSCO open storage yards at the Kwang-Yang works. Atmospheric Torano, J., Rodrigues, R., Diego, I., 2006. Surface velocity contour analysis
Environment 36 (13), 2171–2185. n the airborne dust generation due to open storage piles. In: Proceedings
Park, C.W., Lee, S.J., 2003. Experimental study on surface pressure and flow of the European Conference on Computational Fluid Dynamics.
structure around a triangular prism located behind a porous fence. Journal of ECCOMAS CFD.
Wind Engineering and Industrial Aerodynamics 91 (1–2), 165–184. Yaragal, S.C., Ram, H.S.G., Murthy, K.K., 1997. An experimental investigation of
Patankar, S.V., 1980. Numerical heat transfer and fluid flow. Hemisphere, flow fields downstream of solid and porous fences. Journal of Wind
Washington, DC. Engineering and Industrial Aerodynamics 66 (2), 127–140.

You might also like