You are on page 1of 16

Applied Mathematical Modelling 31 (2007) 2487–2502

www.elsevier.com/locate/apm

Influence of the pile shape on wind erosion CFD


emission simulation
J.A. Toraño, R. Rodriguez, I. Diego, J.M. Rivas, A. Pelegry *

Mining and Civil Works Research Group, Department of Exploitation and Exploration of Mines, School of Mines,
Oviedo University, Asturias, Spain

Received 1 November 2005; received in revised form 1 August 2006; accepted 9 October 2006
Available online 21 December 2006

Abstract

When designing an open storage system for bulk material like coal or iron ore, operational and investment parameters
are to be considered. Those inputs are always managed by engineers and prolific literature can be found. But entering into
environmental parameters like wind erosion, few methodologies are available, US EPA being one of the most extensive.
This source is focused on certain conditions of wind direction, pile shape, etc. what sometimes are not matching the most
interesting way of storing the material: area restrictions, stacking means, etc. Based on state of the art computational fluid
dynamics (CFD) software, in this case we have used ANSYS CFX 5.7., modelling system was developed and compared to
EPA results for conical and oval with flattop piles. After adjusting the calculation process and selecting the most effective
variables, the semicircular shape was studied finding that for the same amount of material stored, lower emissions and
wind erosions are to be expected; however depending on wind direction higher values of friction per surface unit are found
but due to its higher volume of storage per square metre, the balance is positive to the semicircular pile.
 2006 Elsevier Inc. All rights reserved.

Keywords: Airborne dust; Pile shape; Wind speed; Emission rate

1. Introduction

Open areas are common practice for storing huge aggregate bulky cargoes, therefore dust lift must be con-
sidered as a significant environmental and operational problem [1]. Several factors influence the quantity of
dust generated [2,3] and include:

• Wind characteristics like direction, speed, gusts, etc.


• Wind accessibility.
• Elevation of the surface exposed to the wind.
• Particle parameters: diameter, density, shape, moisture, etc.

*
Corresponding author.
E-mail address: angel.pelegry@durofelguera.com (A. Pelegry).

0307-904X/$ - see front matter  2006 Elsevier Inc. All rights reserved.
doi:10.1016/j.apm.2006.10.012
2488 J.A. Toraño et al. / Applied Mathematical Modelling 31 (2007) 2487–2502

• Presence of nonerodible elements and crust formation.


• Disturbance of the surface.

The study of this phenomenon is highly complicate due to this great number of variables to be considered,
so any theoretical – numerical simulation approach would probably result on a forced assumption of several
simplifications [4] but on the other hand an empirical valuation would only cover a limited range of working
conditions, even more, for a given location it is known that conditions are really fluctuant [5,6].
When trying to establish a level of airborne dust generated from an open pile, one of the most extended
methodologies is the USEPA [7], which covers not only that but from unpaved roads up to explosives deto-
nation giving emission factors and procedures to estimate total emissions, control methods, etc.
For piles, USEPA gives several parameters and inputs for cone and flat top oval configurations [8]. How-
ever when a local company placed at the North of Spain found it necessity to storing aggregates (mainly coal)
at an open yard, it considered that from the operational point of view, as well as to minimize the investment,
the most competitive system was to revamp an existing transfer tower. This meant that it would built a slewing
stacking boom of 45 m length supported at such transfer tower, which would stack a 110 semicircular pile
(Fig. 1a).
It was not possible to develop experimental measures to evaluate the level of exposure and dust lift off for
this configuration compared to USEPA standards. Other ways of determining the dust loss from the non-stan-
dard pile shape were considered. It was decided to use CFD software to predict the environmental impact of

(3)

(1)

Figure 1.a.- Option a


(1) Boom Stacker
(2) Feeding Conveyor
(2) (3) Pile

(3)
(3)
(1) (1)

Figure 1.b.- Option b


(2) (1) Reversible Conveyor Stacker
(2) Feeding Conveyor
(3) Cone Piles

Fig. 1. Semicircular and two cone piles.


J.A. Toraño et al. / Applied Mathematical Modelling 31 (2007) 2487–2502 2489

this pile compared to a couple of cone piles with same global capacity located around the transfer tower
(Fig. 1b) which could be considered as another feasible solution from a technical point of view.

2. USEPA explanation

Emissions from storage pile activities may be estimated using the information from sections 13.2.4 and
13.2.5 of AP-42 Compilation of Air Pollutant Emission Factors [16].
The methodology applied by this section 13.2.5 Industrial Wind Erosion requires first to establish wind
parameters being the fastest mile used to convert the values obtained from reference anemometers to equiv-
alent friction velocity:
u ¼ 0:053uþ
10 ; ð1Þ

where u* is the equivalent friction velocity, uþ


10 is the fastest mile of wind at 10 m height.
This equation only applies to flat piles or those with little penetration into surface wind layer. The use of u*
means the assumption that a steady upwind fetch appears.
The emission factor for surface airborne dust subject to disturbances may be obtained in units of grams per
square metre (g/m2) per year as follows:
X
N
Emission Factor ¼ K P i; ð2Þ
i¼1

where K is particle size multiplier, N is number of disturbances per year, Pi is the erosion potential correspond-
ing to the observed fastest mile of wind for the ith period between disturbances, calculated by Eq. (3).
2
P ¼ 58ðu  ut Þ þ 25ðu  ut Þ;
ð3Þ
P ¼0 for u 6 ut ;


where u* is the friction velocity (m/s), ut is the threshold friction velocity (m/s).
Therefore the erosion is directly affected by particle size distribution as well as by the frequency of distur-
bance [9]; because of the nonlinear form of the erosion, each disturbance must be treated separately.
But Eq. (1) assumes a typical roughness height of 0.5 cm and height to base ratio not exceeding 0.2; if the
pile exceeds this value significantly, it would be necessary to divide the pile area into subareas of different
degrees of exposure to wind. When talking about higher height ratios, it would be necessary to use:
u ¼ 0:10uþ S; ð4Þ
u s

S ¼ uþ : ð5Þ
ur 10
EPA Section 13.2.5 gives, for representative cone and oval top flat pile shapes the ratios of surface wind speed
(us) to approach wind (ur) derived from wind tunnel studies. Hence the total surface of the pile is subdivided
into areas of constant u* where Eqs. (3)–(5) can be used.

3. CFD simulation

USEPA procedure for a semicircular shaped pile requires surface sub-division and friction rates for each
sub-division. The commercial CFD software Ansys CFX 5.7 was selected to develop a numerical simulation
of the semicircular pile which could then be used to obtaining the areas of constant u* where Eq. (3) could be
calculated. Before applying CFD calculation to the semicircular pile, validation of the process was done by
comparing CFD model results for cone and flat top oval piles with the USEPA experimental reference study.
As in USEPA Section 13.2.5, input parameters for terrain roughness and wind profile were fixed at 0.5 cm and
by logarithmic profile up to 40 km/h, respectively.
Several options exist for the specification of turbulence quantities at Inlets. Selected one was the Medium
option, which fixes the turbulence intensity at 0.05 and the turbulence viscosity ratio at 10.
2490 J.A. Toraño et al. / Applied Mathematical Modelling 31 (2007) 2487–2502

Ansys CFX 5.7 uses a finite volume method to solve the Reynolds averaged Navier–Stokes equations for
mass and momentum on an unstructured grid. Coupling between pressure and velocity is handled implicitly by
the coupled solver. Ansys CFX has a wide variety of turbulence models available. Initial simulations were per-
formed using the SST and k–e turbulence models with logarithmic wall functions and assuming a smooth wall.
Both models were rejected due to poor results (when compared to US EPA wind tunnel measurements) in
favour of the k–e turbulence with a wall roughness value of 0.5. For this work convergence is considered
acceptable when root mean square normalized value of the residuals is under 105. According to CFX-5.7 Sol-
ver Manual values at this level indicate tight convergence. For all models between 100 and 200 iterations were
needed to achieve this convergence level.
For all models developed, the close-to-wall area (boundary layer) of up to approximately 3 m above the
ground and pile was meshed by using prisms while the rest of the surrounding air was divided into tetrahedral
elements. Modelling with hybrid tetrahedral grids with near surface prism layers resulted in a smaller analysis
model with better convergence of the solution and better analysis results.

4. Mathematical background

All simulations in the present investigation use the commercial general purpose CFD software Ansys CFX
5.7. This software is a second order, pressure/velocity coupled, finite element based control volume method
that uses an unstructured grid and a coupled algebraic multi-grid solver. A detailed description of the relevant
theory can be found in the CFX 5.7. User’s Guide [15]. Main assumptions and hypothesis taken into account
in the development of this paper will be described further on.
The set of equations which describe the processes of momentum, heat and mass transfer in a moving fluid
are known as the Navier–Stokes equations. These partial differential equations were derived in the early nine-
teenth century and have no known general analytical solution but can be discretised and solved numerically.
These equations are three:
The continuity equation:
oq
þ r  ðqU Þ ¼ 0: ð6Þ
ot
The momentum equations:
oðqU Þ
þ r  ðqU  U Þ ¼ rp þ r  s þ S M ; ð7Þ
ot
where the stress tensor, s, is related to the strain rate by
 
T 2
s ¼ l rU þ ðrU Þ  dr  U : ð8Þ
3
The total energy equation:
oðqhtot Þ op
 þ r  ðqUhtot Þ ¼ r  ðkrT Þ þ r  ðU  sÞ þ U  S M þ S E ; ð9Þ
ot ot
where htot is the total enthalpy, related to the static enthalpy h(T, p) by
1
htot ¼ h þ U 2 : ð10Þ
2
The term $ • (U • s) represents the work due to viscous stresses and is called the viscous work term. The term
U • SM represents the work due to external momentum sources and is currently neglected.
Notation used refers to:

q density
$ gradient
$• divergence operator
U vector of velocity
 tensor product
J.A. Toraño et al. / Applied Mathematical Modelling 31 (2007) 2487–2502 2491

p pressure
s stress tensor
SM momentum source
l molecular (dynamic) viscosity
d Kronecker delta function (identity matrix)
htot total enthalpy
k thermal conductivity
T temperature
SE energy source
H static enthalpy

In our case several simplifications can be done, as we will consider an incompressible and Newtonian fluid
and small temperature differences.
In principle, the Navier–Stokes equations describe both laminar and turbulent flows without the need for
additional information. However, turbulent flows at realistic Reynolds numbers span a large range of turbu-
lent length and time scales, and would generally involve length scales much smaller than the smallest finite
volume mesh, which can be practically used in a numerical analysis. The Direct Numerical Simulation
(DNS) of these flows would require computing power which is many orders of magnitude higher than avail-
able in the foreseeable future.
To predict the effects of turbulence, a large amount of CFD research has been concentrated on methods
which make use of turbulence models. Those models have been specifically developed to account for the effects
of turbulence without recourse to a prohibitively fine mesh and Direct Numerical Simulation.
Several solving methods can be used, and most of them have been included in CFX, from quite simple
‘‘Zero Equation Models’’ to complex LES or DES developments.
RANS methods (Reynolds Averaged Navier–Stokes) show a good compromise between accuracy and
calculation effort, being widely used in most of the engineering applications. These are based on describing
the unsteady eddies that form the turbulence by their mean effects on the flow, through the Reynolds
stresses.
Medium complexity turbulence models were selected in order to obtain affordable resolution times in single
processor machines [10,11]. These models were: k–epsilon (with and without surface roughness) and k–w based
Shear–Stress–Transport (SST), combined with different logarithmic and semi-logarithmic wind profiles.
Results were compared against the experimental data included within EPA. The best fit was obtained through
a roughness k–epsilon model using a logarithmic wind profile.
Meshes used in these studies are unstructured, composed by tetrahedral and prismatic elements. Three dif-
ferent meshes (fine, medium and coarse) have been checked and compared their results against the experimen-
tal data contained in EPA. A medium resolution mesh was selected in order to get agreement with the
experimental data but without compromising future models of complex geometry where high density meshes
will make essential to use multiprocessor computers and parallel processing in order to manage the huge
amounts of memory needed to calculate over extensive domains.

5. Cone

The first case simulated was for a cone pile with the geometry shown in the sketch in Fig. 2. In order to
simplify the numerical calculations, a 3D model was constructed in Solidworks to simulate just one half of
the case. The model was meshed by approximately 105,000 elements and symmetry plane applied.
A revolution control surface 25 cm over the cone was defined in order to measure the ratio us/ur named
Variable 1 following the equivalent procedure used by Stunder and Arya [8]. Approximately 13,500 control
locations were identified on such surface. Fig. 3 shows the predicted distribution of Variable 1 for a cone pile
affected by logarithmic profile wind from left to right side.
Values of Variable 1 predicted by CFX were extracted to an Excel calculation data sheet. This was used to
calculate the percentage of total cone surface for each step of values, and then grouped according to those
subareas used by USEPA (Table 1).
2492 J.A. Toraño et al. / Applied Mathematical Modelling 31 (2007) 2487–2502

Fig. 2. Cone pile model.

Fig. 3. Variable 1 (us/ur).

Table 1 shows a comparison of the results predicted by CFD with the EPA values from wind tunnel. Good
agreement between the two is observed.
A better and more objective classification of the agreement between the CFD results and EPA data is the
root mean square values of deviation given by the following:
vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
u
u1 X N
RMSE ¼ t 
2
ðy  ^y i Þ ; ð11Þ
N i¼1 i
J.A. Toraño et al. / Applied Mathematical Modelling 31 (2007) 2487–2502 2493

Table 1
Comparison of subareas by USEPA and CFD – cone
EPA Pile A CFX
0.2a 5% 11.80%
0.2b 35% 27.83%
0.2c 0% 0.01%
0.6a 48% 47.55%
0.6b 0% 0.01%
0.9 12% 12.79
1.1 0% 0.01%
Total 100.00% 100.00%

where N is the number of measurements (in this case the number of subareas), yi is the obtained value (EPA
Pile A column in Table 1), ^y i is the CFD predicted value.
For the cone pile only 3.75% deviation was found, indicating the simulation is quite accurate and valid.

6. Flat top oval

An additional test of the model for an oval flat top pile with different wind directions was considered. Fig. 4
shows the dimension of the computational domain of pile.
A USEPA Pile B1 means oval type with wind 0, Pile B2 is oval one with 20 wind and Pile B3 is when wind
has a 40 angle with respect to the perpendicular of the longitudinal axle of the pile.
For this model approximately 569,000 elements were required to obtain smooth convergence with residuals
of 105.
In order to create the control surface at 25 cm distance, perpendicular to pile shape, several sub surfaces
were generated by MATLAB as per the sketch in Fig. 5. Within each subarea 2000 control points were placed
to obtain values of Variable 1.
Again by using this control surface predicted values of Variable 1 on it were obtained and are shown in
Fig. 6 for different wind directions. Predicted values are compared to USEPA values in Fig. 7 and good agree-
ment is found.
The accuracy of the model is proven by the low values of RSME which are 4.61% for 0, 3.50% for 20 and
3.09% for 40.

Fig. 4. Flat top oval pile model.


2494 J.A. Toraño et al. / Applied Mathematical Modelling 31 (2007) 2487–2502

Fig. 5. Flat top oval pile control surface structure.

Once again it could be considered that CFX computational model has the same level of accuracy as the
USEPA standard when determining values of friction velocities.

7. Semicircular

Based on the previous two cases the CFD model is able to reproduce EPA profiles. The semicircular pile
shape is analysed using the model to assess its potential contribution to dust lift off. Fig. 8 shows the compu-
tational domain used for this case.
Close to 300,000 elements were used to represent the pile and assure a high resolution for the calculations.
Fig. 9 shows the surface mesh used in the semicircular pile calculation.
Once again to get input values of friction velocity to use in the USEPA (AP-42) calculation, a control sur-
face was inserted to calculate Variable 1 around the pile at 25 cm height from pile surface (Fig. 9). Close to
8000 locations for calculation of Variable 1 on the control surface were used.
Due to the asymmetric characterization of the pile, different wind directions were studied: 0, 45, 90, 135
and 180 (Fig. 10a–e).
For each model subareas of equal value of us/ur were plotted and by dividing the control surface into
approximately 8000 locations, the share of the total surface was distributed between these subareas (Table 2).

8. Results and case discussion

USEPA and Axetell and Cowherd [12] provide a methodology for a hypothetical comparison of two dif-
ferent piles located at the same place and disturbed by equal number of incidents under the same wind con-
ditions [13]. This methodology says that the only possible deviation from one to another would be the subarea
distribution for regimes of us/ur. So by comparing the results obtained by a cone pile and the semicircular one,
we would be able to predict which solution from Option a or Option b (Fig. 1a and b, respectively) could be
considered the most conservative or less pollutanting [5].
When analysing the group of two cone shaped piles, it was assumed that wind accessibility is the same for
both of them due to their symmetry. Possible interferences of the transfer tower or conveyors were not con-
sidered because their dimension is not significant and distance to piles longer. This was also adopted when
studying the semicircular pile.
Analysis of the semicircular and cone values (Table 2) shows that the global friction value (absolute value
of Range multiplied by Share of Total Surface) is higher for the semicircular shape when wind has 135 or 180
J.A. Toraño et al. / Applied Mathematical Modelling 31 (2007) 2487–2502 2495

Fig. 6. Variable 1 at 25 cm from pile surface: (a) wind 0 from left, (b) wind 20 from left, (c) wind 40 from left.
2496 J.A. Toraño et al. / Applied Mathematical Modelling 31 (2007) 2487–2502

35%

30%

25%

20% EPA Pile B1


CFX
15%

10%

5%

0%
0.2a 0.2b 0.2c 0.6a 0.6b 0.9 1.1

35%

30%

25%

20% EPA Pile B2


CFX
15%

10%

5%

0%
0.2a 0.2b 0.2c 0.6a 0.6b 0.9 1.1

30%

25%

20% EPA Pile B3


CFX
15%

10%

5%

0%
0.2a 0.2b 0.2c 0.6a 0.6b 0.9 1.1
Fig. 7. Comparison of subareas by USEPA and CFD: (a) flat top oval 0, (b) flat top oval 20, (c) flat top oval 40.

direction. Explanation can be found in the fact that wind faces an aggressive concave geometry which forces it
to accelerate and create a chimney to escape from this tramp; it could be compared with a Venturi effect (see
Fig. 11).
On the other hand, when the incident flow comes from 0, the convex shape changes the flow stream lines
smoothly. Therefore the wind is not accelerated as on the 135 or 180 wind direction cases.
J.A. Toraño et al. / Applied Mathematical Modelling 31 (2007) 2487–2502 2497

Fig. 8. Semicircular pile model.

Fig. 9. Surface mesh around semicircular pile.

The cone pile case and the semicircular pile with 0, 45 and 90 wind direction cases present a convex shape
against the wind flow, but lower values of erosion are predicted for the semicircular case. A possible explana-
tion of this behaviour could be the fact that the reverse flow downwind the cone pile creates a subarea of high
friction over the surface playing determinant role on total emission factor.
Applying the share of the subareas over the total surface area, a global friction value is obtained and shown
in Table 3. The minimum value of the threshold friction velocity was not considered. This assumption could
affect the results in certain cases, because we would take out from the erodible areas lower values of Variable 1.
This ‘‘global friction value’’ is the result of multiplying each range of Variable 1 by its share of the total area,
this giving a quantitative valuation of the fluctuation of the friction velocities affecting a surface. Bigger value
2498 J.A. Toraño et al. / Applied Mathematical Modelling 31 (2007) 2487–2502

Fig. 10. Variable 1 at 25 cm from pile surface: (a) wind 0 from right, (b) wind 45 from right, (c) wind 90 from right, (d) wind 135 from
right, (e) wind 180 from right.

means higher dust lift off when considering the same wind speeds. Note that in Table 2, values for cone pile
were redistributed on more subareas than for USEPA comparison in order to make it equivalent to semicir-
cular pile analysis.
For the semicircular shape analysis it is found (Table 2) that the worst case is considering wind at 135 or
180. This case would be used for comparison with cone piles (Table 3).
Results in Table 3 show that the cone pile has values similar to semicircular pile when the angle of the wind
is 135 or 180. It would be necessary to apply these results to a particular case to see the impact of other vari-
ables like threshold friction velocities, fastest mile, etc.
Table 2

J.A. Toraño et al. / Applied Mathematical Modelling 31 (2007) 2487–2502


Variable 1 at 25 cm
Range Variable 1 (us/ur) Cone Semicircular pile
Share of total Share of total Share of total Share of total surface Share of total surface Share of total surface
surface surface 0 surface 45 90 135 180
0.5 <0.5 1.51 0.00 0.00 0.00 0.00 0.00
0.4 P0.4 and <0.3 2.56 0.13 0.00 1.00 0.00 0.35
0.3 P0.3 and <0.2 2.53 12.35 0.00 5.29 0.00 3.59
0.2 P0.2 and <0.1 2.38 17.09 0.00 3.13 0.00 4.15
0.1 P0.1 and <0 0.91 3.25 0.00 0.64 0.00 0.33
0 P0 and <0.1 0.99 0.41 4.50 0.48 0.57 0.14
0.1 P0.1 and <0.2 2.24 1.62 12.60 2.20 4.49 1.34
0.2 P0.2 and <0.3 3.90 3.50 14.35 4.90 4.57 1.57
0.3 P0.3 and <0.4 4.47 2.12 10.74 7.39 1.63 1.49
0.4 P0.4 and <0.5 5.36 2.88 2.65 6.79 2.80 1.38
0.5 P0.5 and <0.6 6.37 3.83 3.86 5.90 8.01 9.95
0.6 P0.6 and <0.7 7.71 4.69 4.94 6.42 9.79 10.40
0.7 P0.7 and <0.8 9.00 6.08 6.48 10.57 12.37 10.98
0.8 P0.8 and <0.9 9.92 8.29 7.48 9.99 15.02 11.00
0.9 P0.9 and <1 11.16 12.80 11.33 11.25 20.13 13.00
1 P1 and <1.1 13.11 13.59 19.32 14.87 20.16 14.74
1.1 >1.1 15.88 7.37 1.75 9.18 0.46 15.59

Total 100.00% 100.00% 100.00% 100.00% 100.00% 100.00%


Semicircular pile compared to cone shape.

2499
2500 J.A. Toraño et al. / Applied Mathematical Modelling 31 (2007) 2487–2502

Fig. 11. Stream lines and concentration of flow vectors for wind from 180.

Table 3
Global friction value for semicircular pile
Cone Wind at 0 Wind at 45 Wind at 90 Wind at 135 Wind at 180
Global friction factor 70.26 58.98 55.27 60.09 72.31 74.81

One simplification has been considered on the cone case: there are two cones so when wind has 90 direction
from the left, the first pile creates a shadow of wind over the other pile. This influence could be in the range of
reducing speeds up to 50% [10]. Even on this particular case, the values obtained for just one pile are higher
than any of the semicircular cases.
Our case study was based on a total volume of coal of 18,100 m3 (approximately 15,400 t with density
0.85 t/m3) and angle of repose of 32. To store such tonnage, two cone piles of height 15 m were required while
a 110 semicircular pile of 10.45 m height could be used for same amount of coal.
Total surface for two cones and semicircular designs will be approximately 4300 m2 and 1450 m2,
respectively.
The purpose of this pile as emergency storage in case of over production does not allow us to estimate how
many times it would be used per year. In order to develop the present study, it was considered a disturbance
ratio of 12 per year. This assumes that after storing the material it would not suffer any operation for several
days until it is recovered by pay loaders and transported by truck.
Analysing wind statistics from an anemometer located 8 m above ground level, it was found that 12 m/s
could be taken as average of the fastest mile wind in the area, with west and north (180 and 90) the most
significant directions. Considering that emission of dust decrease rapidly (min) if the pile surface is not dis-
turbed and that every pile will be exposed to wind for a long period of time (days) prior to be reclaimed, it
is assumed that taking average values for this comparison does not invalidate the method.
Correcting these wind values to 10 m height by taking surface roughness as 0.005 we have
 
þ þ lnð10=0:005Þ
u10 ¼ u8 ;
lnð8=0:005Þ

10 ¼ 12:36 m=s:

Applying formulas (1) and (3) and a threshold velocity of 1 m/s [14], emission values of PM-30 (K = 1) are
obtained as per Table 4.
It is apparent that there is a great quantitative difference of emissions between the cone and semicircular
piles, even when comparing the cone with the 90 semicircular case which could be considered quite similar
J.A. Toraño et al. / Applied Mathematical Modelling 31 (2007) 2487–2502 2501

Table 4
Calculation of PM-30 emissions per disturbance
Variable 1 us/ur u* P (g/m2) Emission PM-30 (g) per period between disturbances
Cone Semicircular
90 180
0.1 0.12 0.00 0 0 0
0.2 0.25 0.00 0 0 0
0.3 0.37 0.00 0 0 0
0.4 0.49 0.00 0 0 0
0.5 0.62 0.00 0 0 0
0.6 0.74 0.00 0 0 0
0.7 0.87 0.00 0 0 0
0.8 0.99 0.00 0 0 0
0.9 1.11 3.55 1692.18 580.29 670.56
1.0 1.24 9.15 5117.19 1974.48 1957.21
1.1 1.36 16.51 11,190.52 2200.67 3737.30

Total 17,999.90 4755.44 6365.07

Table 5
Calculation of PM-30 emissions
Emission per Emission per Rate emission Rate emission
disturbance (kg) year (kg) per m2 per tonne
Cone 18.00 216.00 0.051 0.014

Semicircular
90 4.76 57.07 0.039 0.004
180 6.37 76.38 0.053 0.005

due to their convex shape exposed to the upwind direction. Emission rates obtained for one year taking the
number of disturbances N = 12 (see Table 5) shows clearly that environmental reasons would recommend the
use of semicircular pile.
If we consider Table 2 values for wind speed close to pile surface, it is an unexpected difference. But it is
only necessary to remember that the surface exposed to wind by two cones is almost three times the area of
the semicircular one, so emission rates per square metre are quite similar. Therefore it is clear that semicir-
cular pile has an advantage because it has a smaller surface to contain the same quantity of material but
not only that, it is able to compensate for the greater value of surface wind, of higher emission per square
metre.

9. Conclusions

Actual accessibility to CFD software by medium and small sized companies, as well as the simplicity of
USEPA standards to obtain empirical relationships between wind speed and quantification of dust lift off from
exposed surfaces enables engineers to develop models to predict with enough accuracy environmental impacts
of future installations.
The interest of this potential accessibility to simple tools to evaluate emissions is clear if we consider that
according to simulations carried out above, for the cones case hundreds of kilos of coal stored during the year
is lost by wind erosion and just by taking 12 disturbances per year when industrial facilities with daily oper-
ation would mean several hundreds of emission episodes.
Thanks to the study presented in this paper, it was obtained a clear understanding of how the shape and
location of the piles can influence the feasibility of the storage systems when talking about emission restric-
tions, and how to combine economical and environmental analysis to obtain the most advantageous design.
2502 J.A. Toraño et al. / Applied Mathematical Modelling 31 (2007) 2487–2502

When designers are able to select position of piles it would be interesting to orientate semicircular piles so
major wind regimes attack the convex side (like in Fig. 10a) in order to reduce friction over the piles. At the
same time, if windbreakers are required, priority will be focused on concave side of the pile.
Even more, Table 2 values are suitable for assessment applications to predict dust uptake from semicircular
shapes by only adjusting on site parameters and getting threshold speeds from empirical study of aggregates or
just from the literature.

Acknowledgement

The authors would like to thank CFX User’s support for their contribution. Also special thanks to the
Ministry of Education and Science of the State of Spain as this work was funded by grant CTM2005-
00187/TECBO through a subcontract with the University of Oviedo: ‘‘Particle Atmospheric Contamination:
Prediction Models and Prevention Systems in Industrial Environment’’.

References

[1] S.J. Page, J.A. Organiscak, Semi-empirical model for predicting surface coal mine drill respirable dust emissions, Int. J. Surf. Mining,
Reclamation Environ. 18 (1) (2004) 42–59.
[2] J. Xuan, Turbulence factors for threshold velocity and emission rate of atmospheric mineral dust, Atmos. Environ. 38 (2004) 1777–
1783.
[3] EPA Technical Assessment Paper, Available Information for Estimating Air Emissions for Stone Mining and Quarrying Operations,
1998.
[4] P.J. Witt, K. Carey, T. Nguyen, Prediction of dust loss from conveyors using CFD modelling, Appl. Math. Modell. 26 (2) (2002) 297–
309.
[5] G.E. Muleski, Review of Surface Coal Mining Emission Factors, EPA-454/R-95-007, US Environmental Protection Agency,
Research Triangle Park, NC, 1991.
[6] Anon, Compilation of Past Practices and Interpretations by EPA Region VIII on Air Quality-Mining, US EPA, Region VIII, Denver,
1979.
[7] U.S. EPA, Revised Draft – User’s Guide for the AMS/EPA Regulatory Model – AERMOD, Office of Air Quality Planning and
Standards, Research Triangle Park, NC, 1998.
[8] B.J.B. Stunder, S.P.S. Arya, Windbreak effectiveness for storage pile fugitive dust control: a wind tunnel study, J. Air Pollut. Control
Assoc. 38 (1988) 135–143.
[9] S.T. Parker, R.P. Kinnersley, A computational and wind tunnel study of particle dry deposition in complex topography, Atmos.
Environ. 38 (2004) 3867–3878.
[10] L. Temmerman, C. Wang, M.A. Leschziner, A comparative study of separation from a three-dimensional Hill using large Eddy
simulation and second-moment-closure Rans modelling, European Congress on Computational Methods in Applied Sciences and
Engineering, ECCOMAS, 2004.
[11] S.A. Silvester, I.S. Lowndes, S.W. Kingman, The ventilation of an underground crushing plant, Mining Technol., Trans. Inst. Min.
Metall., A 113 (2004) 201–214.
[12] K. Axetell, C. Cowherd, Improved Emission Factors for Fugitive Dust from Western Surface Coal Mining Sources, Volume II:
Pre-Publication Copy, US Environmental Protection Agency, Cincinnati, November 1981.
[13] I.P. Castro, W.H. Snyder, R.E. Lawson Jr., Wind direction effects on dispersion from sources downwind of steep hills, Atmos.
Environ. 22 (10) (1988) 2229–2238.
[14] K.S. Hayden, K. Park, J.S. Curtis, Effect of particle characteristics on particle pickup velocity, Powder Technol. 131 (2003) 7–14.
[15] CFX 2004, CFX User’s Guide V. 5.7.
[16] U.S. EPA, AP 42. Fifth Edition, vol. I, Chapter 13, Miscellaneous Sources, 1995.

You might also like