11.07.2014 Views

FIELD TESTING AND EVALUATION OF DUST DEPOSITION AND ...

FIELD TESTING AND EVALUATION OF DUST DEPOSITION AND ...

FIELD TESTING AND EVALUATION OF DUST DEPOSITION AND ...

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>FIELD</strong> <strong>TESTING</strong> <strong>AND</strong> <strong>EVALUATION</strong> <strong>OF</strong> <strong>DUST</strong> <strong>DEPOSITION</strong> <strong>AND</strong><br />

REMOVAL MECHANISMS: FINAL REPORT<br />

Vic Etyemezian<br />

Jack Gillies,<br />

Hampden Kuhns,<br />

Djordje Nikolic<br />

and<br />

John Watson<br />

DRI<br />

755 E. Flamingo rd<br />

Las Vegas, NV<br />

(702) 895-0569<br />

vic@dri.edu<br />

John Veranth,<br />

Raed Laban,<br />

and<br />

Gauri Seshadri<br />

University of Utah<br />

SLC, UT<br />

and<br />

Dale Gillette<br />

NOAA<br />

RTP, NC<br />

Report Prepared for:<br />

The WESTAR Council<br />

9 Monroe Pkwy, Suite 250<br />

Lake Oswego, OR 97035<br />

01/01/03


ACKNOWLEDGEMENTS<br />

The support and assistance of Bob Lebens (WESTAR), the project manager for<br />

this study, were essential for the completion of this work and greatly appreciated. The<br />

prompt reviews of the draft report by Duane Ono (Great Basin Unified APCD) and Mark<br />

Scruggs (National Park Service) were very helpful and we are indebted for the<br />

contribution of their time. The authors additionally thank Christopher Biloft, Michael<br />

Brown (Los Alamos National Laboratory), M. Nelson (University of Utah), Dragon Zajic<br />

(Arizona State University), Sean Ahonen, Mark Green, Judith Chow (Desert Research<br />

Institute), Marc Pitchford (NOAA), Geary Schwimmer and Dave Miller (NASA), Clyde<br />

Durham (FT. Bliss), and Thompson Pace (EPA) for contributing to the completion of this<br />

work through equipment loans, comments on the manuscript, and assistance in data<br />

analysis.<br />

This work was funded primarily by the WESTAR Council. Experiments at the Ft.<br />

Bliss military facility and Dugway Proving Ground were funded by DoD SERDP<br />

contracts CP1190 and CP1191. Field testing was made possible through the cooperation<br />

of the United States Army at the Dugway Proving Ground and at Ft. Bliss, the DoE<br />

NNSA Chemical and Biological National Security Program, and the Mock Urban Setting<br />

test participants. USEPA Southwest Center for Environmental Research and Policy<br />

projects A-01-7 and A-02-7 provided data for supplemental analysis.<br />

iii


EXECUTIVE SUMMARY<br />

The Western States Air Resources Council (WESTAR) contracted the Desert<br />

Research Institute (DRI) and the University of Utah (U of U), through subcontract, to<br />

perform an investigation of a model that has been proposed by Dale Gillette of NOAA<br />

(Countess, 2001) to account for near-source deposition of dust emissions from unpaved<br />

roads. This work was motivated by the well-documented disagreement between estimates<br />

of road and other geological dust obtained by emissions inventory methods and the actual<br />

amount of inorganic minerals observed on filter samples at ambient monitoring sites<br />

(Watson and Chow, 2000). This WESTAR project leveraged two existing DoD-funded<br />

efforts (CP 1190, CP 1991) to increase the utility of data collection and analysis in the<br />

context of meeting the objectives of this study.<br />

The work presented in this report drew on two field studies with robust datasets,<br />

three modeling approaches, and a review of the mechanisms of deposition and dispersion<br />

and their parameterization in air quality models. The first field test took place at the Ft.<br />

Bliss military facility near El Paso, TX in April, 2002. The setting was representative of<br />

the desert Southwestern United States. Atmospheric conditions ranged from near-neutral<br />

to moderately unstable. The second field test was conducted at the Dugway Proving<br />

Ground in September, 2001 as part of the Mock Urban Experiment. The setting was<br />

representative of areas with large roughness elements, similar to suburban tract homes.<br />

Atmospheric conditions were stable. Common to both field tests were towers that were<br />

instrumented with highly time-resolved particle concentration monitors at multiple<br />

heights. This experimental setup allowed for calculation of horizontal dust flux<br />

approximately 100 m downwind of the unpaved road test strip on a per vehicle basis.<br />

The three models evaluated were the Gillette Box Model, a Gaussian plume<br />

model similar to that used in EPA’s ISC, and a numerical solution to a simplified version<br />

of the Atmospheric Diffusion Equation (ADE). Predictions from all three models were<br />

compared with the data from the field studies.<br />

There were six major conclusions from the work documented in this report.<br />

First, the hypothesis that fugitive dust particles emitted from unpaved roads can<br />

deposit to an appreciable extent within several hundred meters downwind of the road<br />

(Watson and Chow , 2000; Countess, 2001) is well-founded, though the fraction of<br />

particles removed varies greatly based on setting and atmospheric stability. In arid<br />

regions such as the Southwestern United States, and for daytime emissions, the<br />

transportable fraction of unpaved road dust emissions is likely to be near unity.<br />

Second, for the purposes of modeling the transport and removal of dust particles<br />

near an unpaved road source, the Gaussian-style models, such as EPA’s ISC3, provide an<br />

imperfect, but reasonable preliminary approach. Comparison of modeled and measured<br />

concentration profiles for desert, daytime conditions indicated that the ISC approximately<br />

simulates the shape of the dust plume from an unpaved road. Similar comparison for<br />

fugitive road dust emissions collected under stable conditions, amidst large roughness<br />

elements, indicated that the ISC model may under estimate the deposition of dust in some<br />

cases. The deposition velocity used by the model is difficult to verify with<br />

measurements, and may be incorrect by a factor of two or more. This means that any<br />

model used, especially in the absence of additional field data, would provide a rough<br />

v


approximation at best. Solution of the Atmospheric Diffusion Equation did not provide<br />

useful information for the first 1,000 or so meters downwind of the road. This was<br />

attributed to the violation of some basic assumptions in the underlying mixing length<br />

theory.<br />

Third, the Gillette Box Model approach is elegant in its simplicity, but the<br />

assumptions behind the model are probably, in general, not valid for unpaved road dust<br />

emissions. Specifically, the assumptions required for the accurate parameterization of<br />

dispersion are not met. Thus, the Box Model may be used as a “back of the envelope”<br />

test for the importance of particle removal downwind of an unpaved road source, but it is<br />

not likely to give reliable quantitative data in all cases.<br />

Fourth, the height of the wake generated behind a moving vehicle has an effect on<br />

the fraction of particles removed as the plume travels downwind, though this effect is<br />

only significant for stable conditions. The turbulent wake of the vehicle increases in<br />

vertical extent approximately linearly with the vehicle height. Considering that the main<br />

processes, dispersion and deposition, are not well-quantified, inclusion of the turbulent<br />

wake height in current modeling attempts appears to be of secondary importance<br />

according to preliminary tests.<br />

Fifth, measurement of emission factors for seven different vehicles with gross<br />

weights varying from 1,622 kg to 23,636 kg were at odds with the silt-based emission<br />

factors suggested in AP-42 (USEPA). The Ft. Bliss measurements indicated that<br />

emission factors were highly dependent on vehicle speed. Linear regression of emission<br />

factor vs. speed for each of the seven vehicles resulted in R 2 values ranging from 0.37 to<br />

0.95, with six of the seven vehicles giving values greater than 0.70. In contrast, the AP-<br />

42 emission factor formulation does not consider vehicle speed. According to the<br />

measurements at Ft. Bliss, for passenger-sized vehicles, the AP-42 appears to<br />

overestimate emissions for vehicle speeds under 35 mph. Emission inventories often<br />

assume an average unpaved road travel speed of 20 or 25 mph. Under those conditions,<br />

use of AP-42 would result in an overestimate of PM 10 emissions by 50% to 100% based<br />

on the field study results.<br />

Sixth, prior to applying a correction to existing emissions inventories, additional<br />

field studies are recommended. In view of the good agreement between the ISC and data<br />

collected in desert, daytime conditions but the poor agreement between the ISC and<br />

measurements under stable conditions with large roughness heights, some selected field<br />

studies that span a range of land cover and atmospheric stabilities are recommended. In<br />

addition, there are a number of other possible sources of error that may result in inflated<br />

emission inventories (Watson and Chow, 2000; Countess, 2001) that should be<br />

considered as a source of the discrepancy between PM 10 dust emission inventories and<br />

ambient concentrations. One finding of the present study is that the AP-42 tends to over<br />

predict emissions of PM 10 dust for smaller, passenger-sized vehicles, especially when the<br />

vehicles are traveling at speeds less than 35 mph.<br />

Watson and Chow (2000) documented the discrepancy between emission<br />

inventories for PM 10 fugitive dust and the source attribution of ambient filter samples.<br />

Their analysis indicated that the amount of geologically-derived PM 10 found in the air is<br />

much smaller than would be expected based on the emission inventories and dispersion<br />

vi


models. Countess (2001) summarized eleven shortcomings in the current treatment of<br />

fugitive dust emissions. Though not all eleven are directly relevant to the work in this<br />

report, we address several of them below.<br />

The removal of dust in the area near a source has been hypothesized to be<br />

substantial because concentrations of particles downwind of the source rapidly attenuate<br />

to the background. Watson and Chow (2000) and Countess (2001) both cite an earlier<br />

study (Watson et al., 1996) where the concentration of PM 10 dust was found to decrease<br />

by 90% only 50 m downwind from the source. The modeling effort in this study showed<br />

that much of the decrease in concentration - indeed for the conditions at Ft. Bliss, most of<br />

the decrease - is due to dispersion in the vertical direction. Therefore, using the<br />

magnitude of the PM 10 concentration downwind of a dust source to estimate the removal<br />

of particles will suggest a much higher deposition rate than is achieved in reality. This<br />

does not invalidate the correct assumption that some of the PM 10 emitted will be removed<br />

as the dust plume travels downwind of the source. However, the removal occurs to a<br />

smaller degree and over larger distances than previously speculated, especially for the<br />

open, sparsely vegetated terrain that much of the American southwest is comprised of.<br />

The results of this study support the first finding put forth by Countess (2001),<br />

“Not all suspendable particles are transported long distances”. However, the Ft. Bliss and<br />

Dugway studies show a large range of near-field removal rates. In the rangeland setting<br />

of Ft. Bliss, little removal of suspended dust was observed at a downwind distances of<br />

100 m. This result was supported with a modeling effort that used a Gaussian approach<br />

similar to the ISC model (USEPA, 1995). Nighttime experiments at the Mock Urban<br />

Setting at the Dugway Proving Grounds suggested that under stable atmospheric<br />

conditions and large roughness elements (cargo containers) 85% of the PM 10 may be<br />

removed by deposition 95 m downwind of the unpaved road. This enormous range of<br />

removal rates emphasizes that it is not appropriate to apply a single correction factor to<br />

all fugitive dust emissions as a means of accounting for near-field particle removal.<br />

Though not documented, the community of scientists and professionals have, in the last<br />

several years, been circulating the idea that if fugitive dust emissions were divided by a<br />

factor of four, then the discrepancy between emissions and ambient measurements of<br />

geological PM 10 would disappear. While it is possible that this is true on an average<br />

basis (i.e. over large spatial domains), it is unlikely that this factor of four is applicable to<br />

every combination of air shed, land use distribution, and atmospheric conditions. Each<br />

combination of setting and meteorological conditions should be considered separately in<br />

a modeling framework that makes use of the known physics of particle dispersion and<br />

deposition.<br />

Related to the last point, the removal of PM 10 dust particles in the “impact zone”<br />

immediately after emission, at the junction of the road edge and the first row of<br />

vegetation, should be examined more closely. The work presented in this report only<br />

considers deposition of particles from “above” the vegetative canopy. It does not<br />

consider that particles may be removed to an appreciable extent by a process similar to<br />

filtration when the dust plume is forced through vegetative elements.<br />

As one of the recommendations from his second finding, “Regional-scale vertical<br />

flux is smaller than the local-scale fugitive dust flux”, Countess (2001) suggests testing<br />

the Gillette box model and modifying the model parameters if appropriate. In this study,<br />

vii


the box model was examined, from a theoretical and practical standpoint. The analysis<br />

indicated that the model is not practical for wide spread use because the parameterization<br />

of dispersion is reliant on assumptions that cannot always be met in reality.<br />

Related to the third finding from Countess (2001), “Fugitive dust emission factors<br />

need to be appropriate”, the emission factors measured at Ft Bliss with high timeresolution<br />

instruments suggest that AP-42 may overestimate PM 10 emissions from<br />

unpaved roads. This overestimate was due to the lack of accounting for vehicle speed<br />

and to the use of the average weight of all vehicles in calculating the emission factors.<br />

All in all, this report concludes that there is no single, “silver bullet” solution to<br />

the problem of reconciling fugitive dust emission inventories and ambient source<br />

contribution estimates. Progress has been achieved in the quantification of the near-field<br />

removal of dust particles and work in that area should continue. There are still<br />

uncertainties about the AP-42 emission factors for unpaved road dust that have resurfaced<br />

in this study and must be resolved.<br />

viii


TABLE <strong>OF</strong> CONTENTS<br />

1. INTRODUCTION 1-1<br />

1.1 Background 1-1<br />

1.2 Objectives 1-1<br />

1.3 Approach 1-1<br />

1.4 Report Organization 1-2<br />

2. BACKGROUND 2-1<br />

2.1 Dry Deposition 2-2<br />

2.1.1 The Model of Slinn (1982) 2-4<br />

2.1.2 Dry Deposition in the ISC3 2-12<br />

2.1.3 The Model of Raupach (1999) 2-13<br />

2.2 Dispersion in the atmosphere 2-14<br />

3. MEASUREMENT <strong>OF</strong> DEPOSTION VELOCITY 3-1<br />

3.1 Methods 3-1<br />

3.1.1 Deposition on Flat Substrates 3-1<br />

3.1.2 Deposition on Artificial Vegetation 3-4<br />

3.1.3 Flat Substrate Method Testing 3-5<br />

3.1.4 Vegetative Method Testing 3-5<br />

3.2 Results: Deposition velocities measured by SEM and gravimetric techniques<br />

3-6<br />

3.3 Discussion 3-11<br />

3.3.1 Comparison of results with the model of Slinn (1982) 3-11<br />

3.3.2 Results in the context of the resistance model for deposition 3-13<br />

3.4 Conclusions 3-14<br />

4. MODELS USED TO STUDY PARTICLE REMOVAL BY DRY <strong>DEPOSITION</strong><br />

DOWNWIND <strong>OF</strong> AN UNPAVED ROAD 4-1<br />

4.1 Presentation of Models 4-1<br />

4.1.1 The one-dimensional Atmospheric Diffusion Equation (ADE) 4-1<br />

4.1.2 The Integrated Source Complex Model 3 (ISC3) 4-4<br />

4.1.3 The Gillette Box Model This section of the text was provided by Dr. Dale<br />

Gillette (NOAA) with some minor adjustments by DRI 4-6<br />

ix


4.2 Inter-comparison of the ADE, ISC3, and the Gillette Box Model 4-10<br />

4.2.1 Analysis of the Gillette Box Model 4-11<br />

4.2.2 Comparison of results for particle removal between predictions from ADE,<br />

ISC3, and the Gillette Box Model 4-14<br />

4.3 Summary 4-20<br />

5. DETERMINATION <strong>OF</strong> PARTICLE REMOVAL RATES, EMISSION<br />

FACTORS, <strong>AND</strong> INJECTION HEIGHTS ON AN UNPAVED ROAD IN AN ARID<br />

SETTING 5-1<br />

5.1 Methods 5-1<br />

5.1.1 Upwind/Downwind Towers 5-1<br />

5.1.2 Sonic Anemometer Tests 5-7<br />

5.2 Results 5-10<br />

5.2.1 Dispersion and deposition downwind of the unpaved road at Ft. Bliss 5-10<br />

5.2.2 Comparison of Measured Emission Factors with AP-42 Emissions 5-17<br />

5.2.3 Injection Height 5-20<br />

5.3 Summary 5-24<br />

6. DETERMINATION <strong>OF</strong> PARTICLE REMOVAL RATES UNDER STABLE<br />

CONDITIONS IN A MOCK URBAN SETTING 6-1<br />

6.1 Experimental Methods 6-1<br />

6.2 Data Analysis Methods 6-2<br />

6.2.1 Wind Speed Models 6-3<br />

6.2.2 Vertical Dust Concentration Models 6-3<br />

6.2.3 Calculation of Horizontal Flux 6-4<br />

6.3 Results 6-4<br />

6.3.1 Wind Profile 6-5<br />

6.3.2 Dust Concentration 6-5<br />

6.3.3 Horizontal Dust Flux 6-6<br />

6.4 Discussion 6-8<br />

6.4.1 Comparison of measurements with model results 6-8<br />

7. SUMMARY <strong>AND</strong> CONCLUSIONS 7-1<br />

7.1 The findings of this study in relation to recent work 7-5<br />

8. REFERENCES 8-1<br />

x


TABLE <strong>OF</strong> FIGURES<br />

Figure 2-1. Development of a dust plume downwind of an unpaved road...................... 2-1<br />

Figure 2-2. Illustration of resistance analogy to deposition. The numbers on the right side<br />

of the figure give an order of magnitude estimate of the physical depth of each layer.<br />

The arrow going through r g indicates that gravity only acts in the downward<br />

direction. The equations for flux in the figure do not include the effect of gravity. 2-3<br />

Figure 2-3. Deposition velocity vs. aerodynamic particle size using Slinn's (1982) model.<br />

See Table 2-1 for parameters used........................................................................... 2-8<br />

Figure 2-4. Contributions of Brownian motion, Interception, and inertial impaction to the<br />

total removal efficiency. See Table 2-1 for parameters used................................... 2-8<br />

Figure 2-5. Variation of Slinn (1982) deposition velocity with estimated parameters. a.<br />

with assumed length scales of vegetation; b. with assumed rebound coefficient b; c.<br />

with assumed viscous to total drag ratio c v /c d . See Table 2-1 for base case.......... 2-10<br />

Figure 2-6. Variation of Slinn (1982) deposition velocity with wind conditions that obey<br />

a logarithmic profile as in Equation 2-9. a. with roughness height z 0 while u * and d<br />

held constant; b. with friction velocity u * while z 0 and d held constant c. with<br />

assumed fraction of canopy height equal to turbulence scale at the top of the canopy<br />

while u * and z 0 held constant. See Table 2-1 for base case ................................... 2-11<br />

Figure 3-1 The experimental setup of the passive deposition on flat surfaces for<br />

examination by electron microscopy. Three-axis (corresponding to the stream-wise,<br />

crosswind, and vertical components of wind direction i.e. +x, +y, +z, -x, -y and –z<br />

directions) passive collection substrates were mounted on a ring stand. Standard<br />

right hand coordinates are used with the positive x-axis downwind perpendicular to<br />

the road. A GRIMM inlet was placed right next to it to monitor the dust<br />

concentration. The substrates were placed 1 or 2m high and 5 or 10m away from the<br />

unpaved road............................................................................................................ 3-2<br />

Figure 3-2. The experimental set up for passive deposition on artificial vegetation (Fir<br />

Christmas garland). The plastic foliage was wrapped around a PVC frame, prewashed<br />

and installed above a clean plastic tub placed 5 m away from the road and 1<br />

m above grade. A GRIMM inlet was placed alongside it to monitor the dust<br />

concentration............................................................................................................ 3-4<br />

Figure 3-3. The deposition velocities calculated for the flat polycarbonate substrates (17,<br />

18, 19) and the artificial vegetative samples (F002, C002) are compared to the<br />

literature values (Slinn 1982). Slinn values are for wind velocity of 5 m/s and<br />

gamma of 3.5. The deposition velocity predicted by the Slinn model is based on the<br />

horizontal area, but the deposition velocity for the polycarbonate surfaces has been<br />

calculated in terms of projected area. ..................................................................... 3-8<br />

Figure 3-4. Directional variation in deposition velocity normal to a flat surface. Data<br />

given are mean and one standard deviation, normalized by the deposition velocity on<br />

the horizontal surface facing up (PZ) averaged for the range 0.5-7.5 m (physical<br />

diameter) based on four determinations (SEM 12, 17, 18, 19). The vertical surfaces<br />

(NX, NY, PY) show enhanced deposition due to impaction. Deposition on surface<br />

facing down (NZ) and away from the road (PX) is presumed to be due to turbulence.<br />

.................................................................................................................................. 3-9<br />

xi


Figure 3-5. Directional variation in deposition velocity normal to a flat surface as a<br />

function of particle size. Data are the mean of four determinations (SEM 12, 17, 8,<br />

19), normalized by the deposition velocity on the horizontal surface facing up (PZ).<br />

The plot indicates no consistent size dependence for particles less than 10 µm in<br />

diameter.................................................................................................................... 3-9<br />

Figure 3-6. . The grand average deposition velocity (SEM 17, 18, 19) for the SEM<br />

samples versus the particle size. The dashed line represents one standard deviation<br />

of the three determinations. Deposition velocity and standard deviation of replicates<br />

data outside the 0.5 m to 5 m range are not shown due to uncertainty in the<br />

accuracy of the data. For d5 m, the results are uncertain due to inlet line loss.<br />

................................................................................................................................ 3-10<br />

Figure 3-7. Variation in particle concentration between sampling days and between<br />

instruments with different inlet configurations. SEM 12 data were obtained from a<br />

GRIMM with a direct sample inlet and SEM 17, 18, 19 data were obtained from a<br />

GRIMM with a 2.5 m long PVC tubing attached to its sample inlet. For diameters<br />

less than 5 m (physical), both the inlets agree but at particle diameters greater than<br />

5 m, the GRIMM with the long inlet tube exhibited sampling losses. Particle<br />

losses affect deposition velocity calculations. ....................................................... 3-10<br />

Figure 3-8. Deposition velocities calculated from SEM particle count for a dry<br />

polycarbonate membrane and a clear tape surface that were collected simultaneously<br />

(SEM 18). From 5 m to 10 m both the substrates give similar particle counts. The<br />

count is increased (and therefore deposition velocity is increased) for particles d20 m) that fell off the dry substrate (data not shown). .................... 3-11<br />

Figure 3-9. Measured deposition velocities on polycarbonate substrates and deposition<br />

velocities calculated based on the boundary layer resistance r b only vs. particle<br />

aerodynamic diameter. The average u * for the sample periods represented by the<br />

measured data and used for the theoretical prediction was 0.35 m/s..................... 3-14<br />

Figure 4-1. Control Volume for Fugitive Dust Model Depicting Vertical and Horizontal<br />

Fluxes....................................................................................................................... 4-6<br />

Figure 4-2. Representation of the Gillette Box Model in the context of the resistance<br />

model to transport. The dashed horizontal line effectively divides the resistance r a<br />

into two smaller resistances, r a1 and r a2 . The equations for fluxes in the figure do not<br />

include the effect of gravity. .................................................................................. 4-12<br />

Figure 4-3. The total aerodynamic resistance from the ground up to the height indicated<br />

on the x-axis. R a calculated after Byun and Dennis (1995). L is the Monin-Obukhov<br />

length and indicates stability. (100,000 > ⏐L⏐ > 10,000 near neutral, 10,000 > L<br />

> 10 stable, 10 > L > 0 very stable, 0 > L > -100 very unstable, -100 > L ><br />

-10,000 unstable). U * = 0.2 m/s and z 0 =0.01 m in all cases.............................. 4-13<br />

Figure 4-4. Fraction of particles remaining in suspension vs. time since emission in the<br />

near field (0


Figure 4-6. The fraction of particles remaining in suspension and the Concentration at a<br />

height of 1 meter above ground level vs. time for neutral atmospheric conditions<br />

according to the ADE and ISC3 models. z 0 =0.01 m, D p =8 µm, u * = 0.3 m/s in all<br />

cases. ...................................................................................................................... 4-17<br />

Figure 4-7. Fraction of particles remaining in suspension for a given downwind distance.<br />

Panels a through e correspond to numerical solutions to the atmospheric diffusion<br />

equation with z 0 =0.01 m, D p =8 µm. under various conditions of atmospheric<br />

stability. The downwind distance is estimated based on the wind speed at 10 m<br />

assuming a logarithmic profile and z 0 = 0.01 m. Panel f shows the fraction<br />

remaining according to the box model for various assumed values of A G ............ 4-18<br />

Figure 4-8. Fraction of particles remaining in suspension vs. time since emission for 8<br />

µm particles falling under the influence of gravity in the absence of any dispersion<br />

(u * =0). .................................................................................................................... 4-19<br />

Figure 5-1. Schematic of equipment setup at Ft. Bliss during the April, 2002 field<br />

campaign. The vertical and horizontal components are drawn on different scales. 5-2<br />

Figure 5-2. Example of time series of DustTrak PM 10 concentrations after the passage of<br />

a vehicle on the unpaved road at Ft. Bliss. The arrows in the figure illustrate the start<br />

and stop times estimated for a baseline reading and the dust plume passing through<br />

the downwind towers DT_1, DT_2, and DT_3. ...................................................... 5-4<br />

Figure 5-3. Frequency distribution and cumulative distribution of u * for the 110 15-<br />

minute intervals corresponding to periods when testing was ongoing at Ft. Bliss. The<br />

u * values are based on a roughness height, z 0 equal to 0.005 m.............................. 5-5<br />

Figure 5-4. Examples of comparison between measured wind speed and wind speed<br />

calculated from curve fitting of Equation 2-9. a. light winds (u * =10 cm/s), b.<br />

medium winds (u * = 32 cm/s), c. high winds (u * = 54 cm/s). The roughness height,<br />

z 0 is equal to 0.005 m in all cases. ........................................................................... 5-5<br />

Figure 5-5. Comparison between collocated spinning cup and sonic anemometers. a.<br />

Example time series from 4/18/02; b. regression of 1-second-averaged and 1-<br />

minute-averaged data from 4/18/02; c. regression of 1-second-averaged and 1-<br />

minute-averaged data from 4/19/02; d. regression of 1-second-averaged and 1-<br />

minute-averaged data from 4/20/02......................................................................... 5-9<br />

Figure 5-6. Example of W-velocity as measured by the sonic anemometer in the vicinity<br />

of a vehicle pass. The vertical lines indicate the start and end times for the two<br />

background intervals and the influence interval. ................................................... 5-10<br />

Figure 5-7. Comparison of concentration profiles from the passage of two convoys with<br />

the concentration profiles at equivalent friction velocities based on the average from<br />

multiple passes of a moving point source. The line source is approximated by the<br />

passage of two convoys on two separate occasions. The moving point source<br />

average concentration profiles were calculated as the integrated concentration from<br />

Equation 5-5. Concentrations are normalized to the value at 1.26 meters in all cases.<br />

................................................................................................................................ 5-12<br />

xiii


Figure 5-8. Concentration profiles measured at the three towers downwind of an unpaved<br />

road and concentration profiles predicted by the ADE model for equivalent<br />

downwind distances a. for 32 passes with u * between 0.15 and 0.25 m/s and b. for 15<br />

passes with u * between 0.45 and 0.55 m/s. Tower distances from edge of road: T1 -<br />

7 m, T2 – 50 m, T3 – 100 m. Modeled profiles are based on neutral atmospheric<br />

conditions. Both measured and modeled concentration have been normalized to the<br />

concentration at 1.26 m.......................................................................................... 5-13<br />

Figure 5-9. Normalized measured and modeled gaussian concentration profiles at a.<br />

Tower 2, 50 meters downwind of unpaved road and b. Tower 3, 100 meters<br />

downwind............................................................................................................... 5-14<br />

Figure 5-10. Comparison of ISC3 gaussian concentration profile and profile from<br />

numerical solution of Atmospheric Diffusion Equation at multiple downwind<br />

distances. Solutions for the ADE are for neutral conditions................................. 5-15<br />

Figure 5-11. Particle removal rates measured at Ft. Bliss at tower 100 m downwind of an<br />

unpaved road. a. Comparison of horizontal flux among the three downwind towers.<br />

Data are averages over 27 individual passes when the wind direction was within 15<br />

degrees of perpendicular to the road. Concentrations measured with DustTrak<br />

monitors with PM 10 inlets. Fluxes calculated according to Equation 5-1 and<br />

normalized to value at Tower 1. Vertical bars indicate standard errors; b.<br />

Comparison of particle size distributions at Towers 1 and 3 measured by GRIMM<br />

1.108 OPC with model predictions. Data are averages over 137 individual passes. X-<br />

axis shows particle aerodynamic diameter calculated by assuming a density of 2.6<br />

g/cm 3 and geometric mean diameter representing each size bin. Particle<br />

concentrations at Towers 1 and 3 were each normalized by concentration of 3.9<br />

micron particles. Y-axis represents ratios of normalized concentrations. ............ 5-16<br />

Figure 5-12. Example of emission factor dependence on vehicle speed. The vertical bars<br />

are the standard deviations among the three downwind towers. ........................... 5-18<br />

Figure 5-13. The ratio of emission factors calculated according to AP-42 (USEPA, 1999)<br />

and those measured at Ft. Bliss. The upper line is the ratio based on a 7% silt<br />

content, while the lower line assumes 4% silt content. The top panel shows the ratios<br />

when all the vehicles in Table 5-2 are considered. The bottom panel shows the same<br />

data, but only for the passenger vehicles. .............................................................. 5-20<br />

Figure 5-14. Turbulence generated vs. height above ground for three vehicle types. The<br />

x-axis represents the difference between the vertical component of turbulence<br />

generated by the passing of the vehicle and the background fluctuations in vertical<br />

velocity normalized by the speed of the vehicle. The horizontal lines represent the<br />

background turbulence standard deviation. The dotted lines represent hand drawn<br />

curves to fit the data. The dotted line corresponding to the compact car was drawn<br />

based on the assumption that the height of the wake plume is proportional to the<br />

height of the vehicle (see Figure 5-15). ................................................................. 5-21<br />

Figure 5-15. Estimate of turbulent wake height vs. physical vehicle height. The dotted<br />

line represents a zero-intercept regression on the data from the Cargo Van and the<br />

Moving Truck. The hollow circle is an estimate of the wake height for the Compact<br />

Car based on the regression. .................................................................................. 5-21<br />

xiv


Figure 5-16. Comparison of fraction of particles removed according to the ISC3 model<br />

for various values of the injection height under a. moderately unstable conditions; b.<br />

neutral conditions; c. very stable conditions; and d. same as c. but x-axis extends to<br />

10 km. The injection height is manifested in the model by setting the initial value of<br />

σ z to 0.5×IH. .......................................................................................................... 5-23<br />

Figure 6-1. Plan view of the MUST site showing the array of shipping containers<br />

representing buildings and the location of the measurements. (1) 2-D sonic<br />

anemometers at 4, 8, 16 m; (2) DustTraks at 0.9, 1.7, and 3.7 m; (3) 3-D sonic<br />

anemometer at 1.6 m; (4) DustTraks at 1.8, 4.6, 9.1, and 18.3 m and 2-D sonic<br />

anemometers at 4, 8, 16, 32 m. ................................................................................ 6-1<br />

Figure 6-2. Wind Equation Fit. Both the logarithmic (dashed line) and power law (solid<br />

line) equations give a reasonable fit to the 15-minute average wind speed<br />

measurements (circles)............................................................................................. 6-6<br />

Figure 6-3. Dust concentration versus time as measured by DustTraks. Vehicle passes<br />

generated well-defined spikes at 3 m horizontal and 1 m vertical from the road<br />

(bottom). The spikes were broader but still well defined 95 m downwind and 4.6 m<br />

above grade (middle). Most vehicle trips did not cause a noticeable spike at 95 m<br />

downwind and 18.3 m above grade (top). Note that full scale for the top graph and<br />

middle graph are 1/1000 and 1/50 of full scale for the near road measurements<br />

respectively. ............................................................................................................. 6-7<br />

Figure 6-4 Collocated DustTraks. Five collocated DustTraks showed large differences in<br />

individual readings but little difference between the instruments when averaged over<br />

the entire experiment. Left: Histogram of maximum minus minimum divided by the<br />

average for sets of five coincident readings. Middle: outlier box plot of the same<br />

data. Right: a typical time trend showing poor correlation between individual<br />

readings.................................................................................................................... 6-8<br />

Figure 6-5. Comparison of fraction of PM 10 particles remaining in suspension 95 m<br />

downwind of an unpaved road during a nighttime test on Dugway Proving Grounds,<br />

UT. The white circles represent the fraction of PM 10 associated with each size bin.<br />

Mass fractions associated with each size bin were estimated from Ft. Bliss data from<br />

a different experiment. The gray circle is the fraction remaining in suspension<br />

according to the methods described in the previous section. The black squares and<br />

triangles correspond to modeled removal rates under very stable conditions. ...... 6-10<br />

xv


xvi


LIST <strong>OF</strong> TABLES<br />

Table 2-1. Parameters used in the Slinn (1982) model for estimating the deposition<br />

velocity over a bean crop. ........................................................................................ 2-9<br />

Table 3-1 Cumulative screen analysis for the soils samples collected before and after the<br />

sampling period........................................................................................................ 3-1<br />

Table 3-2. Log of all flat substrate samples..................................................................... 3-7<br />

Table 3-3. Log of all field vegetative samples................................................................. 3-8<br />

Table 3-4. Deposition velocity calculations for vegetative samples using different areas:<br />

1. Deposition velocity calculation using total area:.............................................. 3-12<br />

Table 3-5. Deposition Velocity calculated for vegetative samples................................ 3-13<br />

Table 4-1. Wind speed at 10 meters corresponding to given friction velocity and<br />

roughness height .................................................................................................... 4-15<br />

Table 5-1. Test dates, times, vehicles, data recovery, and conditions during the FT. Bliss,<br />

April 2002 field campaign. ...................................................................................... 5-6<br />

Table 5-2. Summary of vehicle emission factors measured at Ft. Bliss and Comparison<br />

with emission factors calculated as specified in AP-42 (USEPA, 1999) for two<br />

values of silt content (4% and 7%); moisture content was assumed to be 0.2%. The<br />

speed dependence of the measured emission factors is given in the third column and<br />

represents the slope of a linear regression between measured emission factor (g/vkt)<br />

and vehicle speed (mph); the R 2 value for the regression is given in the fourth<br />

column. For cases where a particular vehicle was tested more than once, the average<br />

of the two tests is used. All emission factors are reported in grams per vehicle<br />

kilometer traveled (g/vkt). ..................................................................................... 5-19<br />

Table 6-1. Screen analysis of test road surface material. Average of two soil samples. 6-2<br />

Table 6-2 Meteorological data from station DPG08, approximately 1 km south of test<br />

site, for September 26, 2002 01:00-03:00 MDT...................................................... 6-4<br />

Table 6-3. Mean ± one standard deviation wind and concentration data for the entire 1.5<br />

hr test period with 44 vehicle trips. Wind standard deviation based on variation<br />

between 15-minute averages. Concentration standard deviation based on variation<br />

between individual vehicle trips. Pulse area is the time-integrated concentration for<br />

the interval corresponding to one vehicle pass. ....................................................... 6-5<br />

xvii


xviii


1. INTRODUCTION<br />

1.1 Background<br />

The Western States Air Resources Council (WESTAR) contracted the Desert<br />

Research Institute (DRI) and the University of Utah (U of U), through subcontract, to<br />

perform an investigation of a model that has been proposed by Dale Gillette of<br />

NOAA(Countess, 2001) to account for near-source deposition of dust emissions from<br />

unpaved roads. This work was motivated by the well-documented disagreement between<br />

estimates of road and other geological dust obtained by emissions inventory methods and<br />

the actual amount of inorganic minerals observed on filter samples at ambient monitoring<br />

sites (Watson and Chow, 2000). This WESTAR-funded project leveraged two existing<br />

DoD funded studies (CP 1190, CP 1991) to increase the utility of data collection and<br />

analysis in the context of meeting the objectives of this study.<br />

1.2 Objectives<br />

Broadly, the goal of this study was to provide insight into the magnitude of PM 10<br />

dust removal by deposition close to the emission source. The analysis was intended to<br />

focus on emissions from unpaved roads, though the results may be pertinent to other<br />

sources of fugitive dust. The specific objectives of the study were:<br />

1. Provide recommendations for feasibility of the Gillette Box Model as a<br />

tool for correcting emissions inventories for fugitive dust from unpaved<br />

roads and suggest possible modifications that may improve model<br />

performance.<br />

2. Compare the results of field studies with the Box Model and with the<br />

algorithm used in ISC3.<br />

3. Provide recommendations for modeling PM 10 dust deposition and<br />

removal and suggest areas to target in future work<br />

1.3 Approach<br />

To achieve the stated objectives, this project drew on two field studies with robust<br />

datasets, three modeling approaches, and a review of the mechanisms of deposition and<br />

dispersion and their parameterization in air quality models. The first field test took place<br />

at the Ft. Bliss military facility near El Paso, TX. The setting was representative of the<br />

desert Southwestern United States. Atmospheric conditions ranged from near-neutral to<br />

moderately unstable. The second field test was conducted at the Dugway Proving<br />

Ground as part of the Mock Urban Experiment. The setting was representative of areas<br />

with large roughness elements, similar to suburban tract homes. Atmospheric conditions<br />

were stable.<br />

The three models evaluated were the Gillette Box Model, a Gaussian plume<br />

model similar to that used in EPA’s ISC, and a numerical solution to a simplified version<br />

of the Atmospheric Diffusion Equation (ADE). Predictions from all three models were<br />

compared with the data from the field studies.<br />

1-1


1.4 Report Organization<br />

This report is divided into six main Chapters. Chapter 2 reviews the literature<br />

with respect to models for particle deposition and turbulent dispersion. Pilot-scale<br />

measurement of deposition to surrogate surfaces and comparison of results with theory<br />

are summarized in Chapter 3. In Chapter 4, the three models considered in this report are<br />

introduced and analyzed. The assumptions of the Box Model and the characteristics of<br />

the ADE and ISC are examined in this Chapter. The Ft. Bliss field experiments are<br />

presented in Chapter 5. In addition to the horizontal flux and particle concentration<br />

profiles, this Chapter details the measurement of vehicle-induced turbulence and its effect<br />

on particle removal downwind of the source. Model predictions are compared with<br />

measurement and used to characterize the effect of the “injection height”. We summarize<br />

The Dugway Proving Ground experiments in Chapter 6 and analyze the related model<br />

results. A summary of findings and major conclusions is provided in Chapter 7.<br />

1-2


2.BACKGROUND<br />

Fugitive dust is emitted from an unpaved road when a vehicle passes and disturbs<br />

the soil. A cloud of dust is raised behind the vehicle and begins to travel downwind.<br />

Initially, the dust cloud is very dense, sometimes to the point of being opaque, but with<br />

travel downwind, the cloud is dispersed by mechanical mixing at the ground and by<br />

buoyant eddies in the atmosphere. Particles suspended in the cloud can be removed by<br />

interaction with the vegetative cover in the downwind fetch of the unpaved road.<br />

Figure 2-1 schematically illustrates the progression of a dust plume emitted<br />

behind a vehicle and advected downwind over a vegetative cover. In the region (A)<br />

where the dust cloud first meets the vegetation, dust particles may be removed by<br />

individual vegetation elements. Though possibly significant, none of the models or<br />

measurements discussed in this report is geared towards quantifying the extent of<br />

removal in this “impact zone”. Some work in the area of particle removal by windbreaks<br />

may be applicable to this region (Raupach and Leys, 1999), though the specific<br />

formulation would have to be adjusted for vegetative covers with a long fetch (as<br />

opposed to a windbreak that is only several meters deep).<br />

Region C: Far Downwind<br />

50 m < x < ∞<br />

Region B: Near Source<br />

5 m < x < 200 meters<br />

Region A: Impact Zone<br />

0 < x < 20 m<br />

Point of<br />

Emission<br />

Figure 2-1. Development of a dust plume downwind of an unpaved road<br />

Very far downwind (Region C in the figure), the dust plume approaches a steady<br />

concentration profile that does not change very much with transport downwind. In<br />

between the “impact zone” and the “Far downwind” region, the concentration profile of<br />

dust particles is changing with transport distance; specifically, the dust plume is<br />

expanding in the vertical direction. The models discussed in this report are most<br />

concerned with this region, where it is necessary to simulate the expansion of the dust<br />

plume in the vertical direction simultaneously with the removal of particles at the ground<br />

by the vegetative cover. These two processes are turbulent dispersion and dry deposition,<br />

respectively.<br />

2-1


2.1 Dry Deposition<br />

There are a number of models available for dry deposition of particles. Three of<br />

them are presented below. Because it is widely cited and because it has been suggested<br />

for use in the Gillette Box Model, the formulation provided by Slinn (1982) is discussed<br />

at some length. One of the difficulties of using Slinn’s model is that the parameters<br />

required are numerous and difficult to specify. Simpler methods that are based on similar<br />

reasoning, but requiring fewer parameters, are used in the Industrial Source Complex<br />

utility (USEPA, 1995) and by Raupach (1999).<br />

Dry deposition is frequently modeled in analogy to resistance in an electrical<br />

circuit. The flux to the ground is assumed to be equal to the concentration measured at a<br />

given height multiplied by the mass transfer coefficient, which is also dependent on<br />

height. The mass transfer coefficient is called the deposition velocity and is given the<br />

symbol v d . Thus, according to the resistance model<br />

F = C( z)<br />

v ( z)<br />

d<br />

Equation 2-1<br />

where F is the flux of material depositing to the ground (g/m 2 s) and C(z) is the heightdependent<br />

concentration. The deposition velocity is given as the inverse of three<br />

resistances in series. That is,<br />

1<br />

vd<br />

=<br />

ra<br />

+ rb<br />

+ rc<br />

Equation 2-2<br />

where r a is the resistance to aerodynamic transport through the surface layer, r b is the<br />

resistance to transport through the “quasi-laminar sublayer”, and r c is the resistance to<br />

collection on vegetative elements (The subscripts a, b, and c are unrelated to the Regions<br />

in Figure 2-1). Figure 2-2 shows the resistance model to deposition. It is built on the<br />

assumption that close to the ground (i.e. from the top of the surface layer down), the flux<br />

of a depositing species is constant with height. This is analogous to the flow of electrons<br />

through a circuit where the number of electrons that leave the positive terminal (top of<br />

the surface layer) must be matched by the number arriving at the negative terminal (the<br />

deposition surface). Related to the assumption of constant flux, the model also assumes<br />

that the system is approximately at steady-state. That is, the concentration profile through<br />

the surface and quasi-laminar layers is not changing significantly over time. This<br />

corresponds strictly only to region C in Figure 2-1, though it may also hold<br />

approximately for region B.<br />

It is usually assumed that for particles with aerodynamic diameters less than 10<br />

µm, r c is very small. That is, once a particle impacts on a collector, it sticks to that<br />

collector and is permanently removed from the atmosphere. Deviations from this ideal<br />

are addressed by Slinn (1982) as summarized in section 2.1.1.<br />

2-2


Planetary Boundary Layer (PBL)<br />

F = F(z)<br />

~200-2000 m<br />

C = C SL<br />

r a<br />

Surface Layer<br />

F = F s =V da [C SL -C QL ]=(1/r a ) [C SL -C QL ]<br />

~2-20 m<br />

r g<br />

r b<br />

C = C QL<br />

Quasi-laminar sublayer<br />

F = F s =V db [C QL -C 0 ]=(1/r b ) [C QL -C 0 ]<br />

C = C 0 =0<br />

~0.1-1 mm<br />

r c<br />

r c assumed = 0 for particles if particle rebound is<br />

negligible<br />

Figure 2-2. Illustration of resistance analogy to deposition. The numbers on the right side of the<br />

figure give an order of magnitude estimate of the physical depth of each layer. The arrow going<br />

through r g indicates that gravity only acts in the downward direction. The equations for flux in the<br />

figure do not include the effect of gravity.<br />

In addition to turbulent transport through the surface and quasi-laminar layers,<br />

particles are also influenced by gravity. This effect is accounted for in the resistance<br />

model by assuming that gravitational settling represents another resistor (r g ) acting in<br />

parallel with those shown in Figure 2-2. With the surface resistance r c equal to zero and<br />

the inclusion of gravity, the equation used to describe the deposition velocity is<br />

v 1<br />

d<br />

=<br />

v<br />

r + r + r r v<br />

+<br />

a<br />

b<br />

a b<br />

g<br />

g<br />

Equation 2-3<br />

where v g is the gravitational settling velocity and is equal to 1/r g . V g is a well-known<br />

function of the particle size and the characteristics of the surrounding fluid (air). For<br />

particles with aerodynamic diameter less than about 20 µm, it is given by (e.g. Seinfeld<br />

and Pandis, 1998)<br />

v<br />

g<br />

=<br />

1<br />

18<br />

2<br />

D ρ gC<br />

p<br />

p<br />

µ<br />

c<br />

Equation 2-4<br />

2-3


where D p is the particle diameter, ρ p is the particle density, g is the gravitational constant,<br />

µ is the viscosity of air, and C c is the Cunningham slip correction factor,<br />

C<br />

c<br />

2λ<br />

<br />

1.1D<br />

<br />

= + + −<br />

p<br />

1 1.257 0.4exp <br />

D <br />

2 λ<br />

p<br />

<br />

Equation 2-5<br />

with λ equal to the mean free path of air molecules.<br />

2.1.1 The Model of Slinn (1982)<br />

In a seminal paper, Slinn (1982) described the framework for a model to predict<br />

particle dry deposition to vegetative canopies. The model is based on the mass balance<br />

expression<br />

∂ ∂C<br />

<br />

=<br />

z <br />

Κ<br />

∂ ∂z<br />

<br />

<br />

( c αu) C<br />

d<br />

ξ c<br />

Equation 2-6<br />

where K is the turbulent diffusivity in the vertical direction, C is the mass or number<br />

concentration of particles, c d is the average drag coefficient for vegetation, α is the<br />

surface area of vegetation per unit volume, u is the mean wind speed, and ξ c is the<br />

canopy element collection efficiency for particles.<br />

The LHS of Equation 2-6 represents the flux of particles through a horizontal<br />

plane located at a height z above the ground while the RHS is the removal rate of<br />

particles by dry deposition. It is implied in the form of Equation 2-6 that turbulent<br />

dispersion only serves to bring particles from high above the canopy towards the ground.<br />

Practically, this means that the particle concentration profile above the vegetative canopy<br />

is well-developed and at steady state (i.e. not changing in time). This situation<br />

corresponds strictly only to Region C in Figure 2-1. However, if the deposition rate<br />

calculated is assumed to apply only locally (i.e. not constant in the downwind direction),<br />

then we may extend the applicability of Equation 2-6 into Region B.<br />

The final form of the deposition velocity provided by Slinn (1982) includes a term<br />

to account for gravitational settling:<br />

v<br />

d<br />

= v<br />

g<br />

+ C<br />

D<br />

uh<br />

ur<br />

1<br />

+<br />

<br />

ur<br />

<br />

<br />

ξ +<br />

1−ξ<br />

ξ tanhγ<br />

<br />

<br />

<br />

ζ <br />

−1<br />

Equation 2-7<br />

where C D is the overall drag coefficient for the canopy, u h is the wind speed at the top of<br />

the canopy (h being equal to the canopy height), u r is the wind speed at some reference<br />

height z r above the canopy, ξ is the location-independent canopy collection efficiency,<br />

and γ is a parameter that characterizes the wind profile within the canopy.<br />

2-4


C<br />

D<br />

u* =<br />

ur<br />

<br />

The overall drag coefficient C D is given by<br />

2<br />

Equation 2-8<br />

where u * is the friction velocity which may be estimated if the velocity profile above the<br />

canopy is approximately logarithmic with either<br />

u −<br />

= *<br />

z d<br />

u(<br />

z)<br />

ln<br />

<br />

κ z0<br />

or<br />

u z)<br />

= u<br />

*<br />

(<br />

r<br />

<br />

<br />

u zr<br />

−<br />

− ln<br />

κ<br />

<br />

z −<br />

( h − l)<br />

( ) <br />

h − l <br />

Equation 2-9<br />

Equation 2-10<br />

where κ is the von Karman constant and is equal to 0.4, d is the displacement height, z 0 is<br />

the roughness height, h is the canopy height, and l is the length scale of turbulence at the<br />

top of the canopy. The displacement height d can be thought of as the height above which<br />

the velocity profile is expected to have a logarithmic shape. Under some conditions, the<br />

displacement height and the roughness height z 0 may be difficult to estimate. For this<br />

reason, it may be preferable to use Equation 2-10, which utilizes characteristics of the<br />

canopy, namely the canopy height and the length scale of turbulence at the top of the<br />

canopy. Slinn suggests the approximation<br />

d = h − l .<br />

The ratio u h /u r in Equation 2-7 is given by Equation 2-9 with h-d=l<br />

Equation 2-11<br />

u<br />

u<br />

h<br />

r<br />

=<br />

u *<br />

l<br />

ln<br />

κur<br />

z<br />

0<br />

<br />

<br />

<br />

Equation 2-12<br />

or by Equation 2-10:<br />

u<br />

u<br />

h<br />

r<br />

=<br />

u z<br />

− ln<br />

u κ <br />

<br />

1<br />

*<br />

r<br />

r<br />

( h − l)<br />

−<br />

l<br />

<br />

<br />

.<br />

<br />

Equation 2-13<br />

2-5


The particle collection efficiency ξ is given by<br />

ξ = EB = ( E + E E )B<br />

B IN<br />

+<br />

IM<br />

Equation 2-14<br />

where E B , E IN , and E IM are the removal efficiencies due to Brownian diffusion, particle<br />

interception, and inertial impaction, respectively and E is the total removal efficiency due<br />

to the three processes operating concurrently. The factor B (≤1) represents a reduction in<br />

capture efficiency due to particles bouncing off of surfaces.<br />

E<br />

B<br />

The Brownian diffusion component can be estimated from<br />

cv<br />

= <br />

Sc<br />

cd<br />

<br />

− 2 3<br />

Equation 2-15<br />

where c v /c d is the ratio of viscous to total drag (equal to the sum of viscous drag and form<br />

drag) and assumed to be 1/3 by Slinn and Sc is the Schmidt number (Sc =ν/D, ν is the<br />

kinematic viscosity of air and D is the Brownian diffusivity of the particle given by the<br />

Stokes-Einstein relation). Note that Brownian diffusion is expected to be a negligible<br />

pathway for deposition for particles with diameters greater than 1 µm.<br />

E<br />

IN<br />

c<br />

=<br />

<br />

c<br />

The interception term E IN is calculated from<br />

v<br />

d<br />

<br />

<br />

F<br />

<br />

s <br />

<br />

D<br />

<br />

p<br />

Dp<br />

A<br />

+<br />

s<br />

2<br />

<br />

<br />

+<br />

<br />

<br />

( 1−<br />

F )<br />

s<br />

<br />

<br />

<br />

D<br />

<br />

p<br />

Dp<br />

A<br />

+<br />

L<br />

<br />

<br />

<br />

<br />

2 <br />

Equation 2-16<br />

where F s is the fraction of the total interception due to small collectors in the canopy such<br />

as hairs, A s is a characteristic width of the small collectors, A L is a characteristic radius of<br />

the large collectors such as blades, stalks, and needles.<br />

E<br />

IM<br />

The Impaction term E IM is equal to<br />

2<br />

Sta<br />

<br />

1+<br />

Sta<br />

=<br />

2<br />

<br />

<br />

where St a is an “average” Stokes number defined by<br />

Equation 2-17<br />

2-6


St<br />

a<br />

vg<br />

u<br />

g<br />

<br />

=<br />

<br />

cA<br />

L<br />

*<br />

Equation 2-18<br />

with c being an unspecified constant that is expected to be approximately equal to unity.<br />

B = exp − b<br />

The factor B is estimated from<br />

( )<br />

St a<br />

where b is another unspecified constant.<br />

Equation 2-19<br />

The factor γ used to represent the wind conditions within the canopy can be<br />

approximated by<br />

( u u )<br />

*<br />

γ ≅<br />

κ<br />

or<br />

h<br />

h<br />

h − d<br />

Equation 2-20<br />

1 2<br />

c d<br />

α<br />

γ ≅ h .<br />

κ l <br />

Equation 2-21<br />

Note that the height-dependent α is related to the bulk Leaf Area Index (LAI), defined as<br />

the ratio of upward facing leaf surface to ground surface in a canopy, according to<br />

1<br />

LAI =<br />

2<br />

<br />

0<br />

h<br />

α ⋅ dz .<br />

Equation 2-22<br />

The parameters that are used for an example of deposition to a bean crop are<br />

given by Slinn (1982) and summarized in Table 2-1. Figure 2-3 demonstrates the classic<br />

v-shaped deposition velocity curve as a function of particle size for unit density particles<br />

(ρ p =1 g/cm 3 ); particles that are less than 0.1 µm in diameter deposit primarily due to<br />

Brownian motion; particles in the 1 µm to 10 µm size range deposit by impaction;<br />

intermediate-sized particles (0.1 µm to 1 µm) are removed by a combination of Brownian<br />

motion, interception, and impaction. Figure 2-4 shows the relative contributions of E B ,<br />

E IN , and E IMP to the total collection efficiency E T for the example of flow over a bean<br />

crop. Though not shown in Figure 2-4, the gravitational settling term v g dominates the<br />

2-7


1E+01<br />

1E+00<br />

vd (cm/s)<br />

1E-01<br />

1E-02<br />

1E-03<br />

1E-03 1E-02 1E-01 1E+00 1E+01 1E+02<br />

Dp (microns)<br />

Figure 2-3. Deposition velocity vs. aerodynamic particle size using Slinn's (1982) model. See Table 2-1<br />

for parameters used.<br />

1E+01<br />

Etotal, Ebrownian, Einterception,<br />

Eimpaction<br />

1E+00<br />

1E-01<br />

1E-02<br />

1E-03<br />

E_total<br />

E_Brownian motion<br />

E_Interception<br />

E_Impaction<br />

1E-04<br />

1E-04 1E-03 1E-02 1E-01 1E+00 1E+01 1E+02<br />

Dp (microns)<br />

Figure 2-4. Contributions of Brownian motion, Interception, and inertial impaction to the total<br />

removal efficiency. See Table 2-1 for parameters used.<br />

RHS of Equation 2-3 for particles larger than 10 µm and for particles between 1 µm and<br />

10 µm under low wind conditions.<br />

It is clear from examination of Table 2-1 that use of Slinn’s model requires the<br />

specification of a number of parameters that may not be available. For example, the<br />

parameters F s , A s , A L , c, b, and c v /c d are generally unavailable for a wide range of<br />

vegetative covers. These parameters are roughly estimated by a best-guess approach.<br />

Figure 2-5 illustrates the sensitivity of the model with respect to uncertainties in these<br />

parameters. The Figure is confined to the vicinity of the particle size range relevant for<br />

PM 10 dust. F s and A s only affect the deposition velocity through the interception term E IN<br />

(Equation 2-16). Since interception is a minor pathway for super micron particles, the<br />

effect of varying these two parameters is small. In contrast, changing the size of A L from<br />

1 mm to 0.5 mm measurably improves the impaction efficiency and the overall collection<br />

efficiency. It is unclear how the coefficient b should be chosen for a given canopy.<br />

Rebound for particles with diameter less than ≈ 5 µm appears to be a minor<br />

consideration. The effect of increasing the value of b from 2 to 4 is to decrease the overall<br />

deposition velocity for 10 µm particles by about 30%. Varying the ratio of the viscous<br />

drag to the total drag between 0.25 and 0.45 exerts little effect on the deposition<br />

velocities for particles in the 1 – 10 µm size range.<br />

2-8


Table 2-1. Parameters used in the Slinn (1982) model for estimating the deposition velocity over a<br />

bean crop.<br />

Parameter Description Equation used Value in Base<br />

Case<br />

Fraction of total interception by<br />

User-specified 1%<br />

F s<br />

small collectors such as vegetative<br />

hairs<br />

A s<br />

Characteristic width of small<br />

collectors such as vegetative hairs<br />

User-specified<br />

10 -5 m<br />

(10 microns)<br />

Characteristic radius of large<br />

collectors such as grass blades,<br />

User-specified<br />

10 -3 m<br />

(1 mm)<br />

A L<br />

stalks, needles, etc<br />

Unknown numerical factor used in<br />

User-specified 1<br />

c<br />

Equation 2-18 and assumed to be<br />

near unity<br />

Numerical factor used to calculate<br />

User-specified 2<br />

b particle rebound in Equation 2-19<br />

Ratio of viscous to total drag<br />

User-specified 1/3<br />

c v/c d<br />

within the canopy<br />

h Height of the canopy User-specified 1.18 meters<br />

Fraction of canopy equal to<br />

User-specified 20%<br />

F l<br />

characteristic Eddy size<br />

u * Friction velocity User-specified 0.25 m/s<br />

Roughness height User-specified or estimated from knowledge of z r, u r, 0.054 m<br />

z 0<br />

h, and l<br />

Wind speed at reference height z r User-specified or estimated from u *, z 0, and d and 1.94 m/s<br />

u r<br />

Equation 2-9 or from u *, z r, u r, h, and l, and Equation<br />

2-10<br />

Wind speed at canopy height h Estimated from l and z 0 and Equation 2-12 or from u r, 0.81 m/s<br />

u h<br />

u *, h, and l and Equation 2-13<br />

d Displacement height User-specified or estimated from h and F l and Equation 0.94 m<br />

2-11<br />

l Characteristic Eddy size at the top<br />

Estimated from F l and h<br />

0.24 m<br />

γ<br />

of the canopy<br />

Parameter that describes the wind<br />

speed profile within the canopy<br />

expected to be between 2 and 5<br />

User-specified or estimated from u *, u h, h, and d and<br />

Equation 2-20 or from c d, l, h, and α and Equation<br />

2-21<br />

3.8<br />

Slinn (1982)<br />

suggests using<br />

γ=3. However,<br />

this assumption<br />

for γ is not<br />

consistent with<br />

Equation 2-20<br />

The deposition velocity is sensitive to the assumed wind conditions. The<br />

roughness height, z 0 , and the displacement height, d, are mathematical tools used to fit<br />

the wind speed to a logarithmic profile such as in Equation 2-9; they are not directly<br />

measurable quantities. In general, z 0 varies from 10 -5 m for smooth ice to 10 m for a<br />

high-density urban area (Seinfeld and Pandis, 1999). Vegetative covers span the range<br />

between 0.01 m for cut lawn to 1 m for a tree-covered area. The displacement height is<br />

related to the canopy height, but can only be roughly estimated without field<br />

measurement of the wind profile above the canopy. Slinn (1982) makes the<br />

approximation that the displacement height is equal to 80% of the canopy height and that<br />

the sum of d and the characteristic Eddy size, l, is equal to the canopy height (i.e. l=0.2h).<br />

U * is also a “fitted” quantity and specifying its value requires some knowledge of the<br />

wind speed profile. Figure 2-6 shows that there can be significant deviations in v d when<br />

slightly different values for z 0 , u * , and l are assumed. Thus, without accurate data for<br />

these parameters, the deposition velocity may only be known to within a factor of about<br />

two or three.<br />

2-9


1E+01<br />

Fs = 1%; As = 10 microns; AL=1 mm<br />

Fs = 0.1%; As = 10 microns; AL= 1 mm<br />

1E+00<br />

Fs = 1%; As = 10 microns; AL=0.5 mm<br />

vd (cm/s)<br />

1E-01<br />

1E-02<br />

1E-03<br />

1E-01 1E+00 1E+01 1E+02<br />

Dp (microns)<br />

a.<br />

1E+01<br />

1E+00<br />

b=1<br />

b=2<br />

b=4<br />

vd (cm/s)<br />

1E-01<br />

1E-02<br />

1E-03<br />

1E-01 1E+00 1E+01 1E+02<br />

Dp (microns)<br />

b.<br />

1E+01<br />

1E+00<br />

cv/cd=0.25<br />

cv/cd=0.33<br />

cv/cd =0.45<br />

vd (cm/s)<br />

1E-01<br />

1E-02<br />

1E-03<br />

1E-01 1E+00 1E+01 1E+02<br />

Dp (microns)<br />

c.<br />

Figure 2-5. Variation of Slinn (1982) deposition velocity with estimated parameters. a. with assumed<br />

length scales of vegetation; b. with assumed rebound coefficient b; c. with assumed viscous to total<br />

drag ratio c v /c d . See Table 2-1 for base case.<br />

2-10


1E+01<br />

1E+00<br />

z0 = 4.4 cm<br />

z0 = 6.4 cm<br />

z0 = 8.4 cm<br />

vd (cm/s)<br />

1E-01<br />

1E-02<br />

1E-03<br />

1E-01 1E+00 1E+01 1E+02<br />

Dp (microns)<br />

a.<br />

1E+01<br />

1E+00<br />

u* = 0.2 m/s<br />

u* = 0.25 m/s<br />

u* = 0.3 m/s<br />

vd (cm/s)<br />

1E-01<br />

1E-02<br />

1E-03<br />

1E-01 1E+00 1E+01 1E+02<br />

Dp (microns)<br />

b.<br />

1E+01<br />

1E+00<br />

l = 0.15 h<br />

l = 0.2 h<br />

l = 0.3 h<br />

vd (cm/s)<br />

1E-01<br />

1E-02<br />

1E-03<br />

1E-01 1E+00 1E+01 1E+02<br />

Dp (microns)<br />

c.<br />

Figure 2-6. Variation of Slinn (1982) deposition velocity with wind conditions that obey a logarithmic<br />

profile as in Equation 2-9. a. with roughness height z 0 while u * and d held constant; b. with friction<br />

velocity u * while z 0 and d held constant c. with assumed fraction of canopy height equal to turbulence<br />

scale at the top of the canopy while u * and z 0 held constant. See Table 2-1 for base case<br />

2-11


Aside from the error associated with improperly specifying parameters, it is<br />

unclear if the assumptions in Slinn’s model apply to sparsely vegetated landscapes where<br />

inter-plant distances may be several times the height of individual plants. The model was<br />

derived for a “horizontally-homogeneous” canopy, where quantities are dependent only<br />

on the distance z above the ground.<br />

2.1.2 Dry Deposition in the ISC3<br />

For a stable atmosphere, the aerodynamic resistance r a is given by (Byun and<br />

Dennis, 1995):<br />

1 zr<br />

ra = ln<br />

<br />

+ 4. 7<br />

κu*<br />

z0<br />

<br />

z <br />

<br />

L<br />

Equation 2-23<br />

while for an unstable atmosphere, it is given by<br />

r<br />

a<br />

1<br />

=<br />

κu<br />

<br />

ln<br />

<br />

<br />

( 1+<br />

16z<br />

)( )<br />

( )( ) <br />

r<br />

/ L −1<br />

1+<br />

16z0<br />

/ L + 1<br />

1+<br />

16zr<br />

/ L + 1 1+<br />

16z0<br />

/ −1<br />

<br />

* L<br />

Equation 2-24<br />

where L is the Monin-Obukhov length scale (L > 0 stable conditions, L < 0 unstable<br />

conditions, ⏐L⏐>> 1 neutral conditions). The reference height z r is taken to be the<br />

larger of 1 m and 20×z 0 .<br />

1995):<br />

r<br />

b<br />

=<br />

The quasi-laminar resistance is given by (Pleim et al. 1984; Slinn, 1982; USEPA,<br />

1<br />

−<br />

( 2 3 −3<br />

St<br />

Sc + 10 ) u *<br />

Equation 2-25<br />

where Sc is the Schmidt number (Sc =ν/D, ν is the kinematic viscosity of air and D is the<br />

Brownian diffusivity of the particle given by the Stokes-Einstein relation) and St is the<br />

Stokes number in a slightly different form than Equation 2-18<br />

St =<br />

( v g)<br />

g<br />

υ<br />

2<br />

u *<br />

Equation 2-26<br />

The term containing the Schmidt number in Equation 2-25 accounts for Brownian<br />

motion while the term containing the Stokes number accounts for inertial impaction.<br />

Interception is not explicitly accounted for in this formulation, and neither is particle<br />

rebound. However, it is clear from Figure 2-4 that inertial impaction is the dominant<br />

2-12


pathway for deposition for particles with diameters relevant to dust emission, i.e. 1 – 10<br />

µm in diamter.<br />

This model for dry deposition requires knowledge of u * , z 0 , and L.<br />

2.1.3 The Model of Raupach (1999)<br />

Unlike the multi-layer, multi-resistance approach characterized by Equation 2-3, Raupach<br />

(1999) has proposed that a single layer with multiple conductances can, under some<br />

conditions, adequately represent deposition according to<br />

v = v + G + G<br />

d<br />

g<br />

IM<br />

B<br />

Equation 2-27<br />

where v d and v g are defined as before and G IM and G B are the bulk conductances for<br />

impaction and Brownian motion, respectively. Raupach (1999) uses the portion of the<br />

bulk aerodynamic conductance G aM (G aM = u * 2 /u r ) associated with viscous drag to scale<br />

G B and the portion associated with form drag to scale G IM . Thus,<br />

G = a<br />

IM<br />

f<br />

E<br />

IM<br />

( G f )<br />

aM<br />

form<br />

Equation 2-28<br />

and<br />

G<br />

B<br />

v<br />

B<br />

[ G ( − f )]<br />

= a E 1<br />

aM<br />

form<br />

Equation 2-29<br />

where E IM and E B are the particle impaction and the Brownian diffusion efficiencies,<br />

respectively, f form is the fraction of the total drag that is form drag and is assumed to be<br />

approximately 75% (the remainder is viscous drag) and a v and a f are numerical<br />

parameters of order unity that account for inter-element sheltering. Raupach (1999)<br />

suggests using the values 8 and 2, respectively based on a best-fit to Chamberlain’s<br />

(1967) wind tunnel data. As in the earlier examples, the efficiency of removal by<br />

Brownian motion is given by the Schmidt number raised to the –2/3 power:<br />

−2 3<br />

E B<br />

= Sc<br />

E IM is given by a form similar to that in Slinn (1982) (see Equation 2-17)<br />

Equation 2-30<br />

E<br />

IM<br />

Std<br />

=<br />

<br />

Std<br />

+<br />

p <br />

q<br />

Equation 2-31<br />

2-13


where p and q are numerical factors estimated to be 0.8 and 2, respectively from earlier<br />

work (Peters and Eiden, 1992; Bache, 1981). St d is a form of the Stokes number<br />

St<br />

d<br />

2<br />

=<br />

( v g)<br />

g<br />

d<br />

c<br />

U<br />

c<br />

Equation 2-32<br />

with U c equal to the flow velocity around the vegetation elements and d c equal to a<br />

characteristic width of canopy elements upon which particles impact. Raupach (1999)<br />

suggests that U c is “characterized” by the friction velocity u * . Note that if U c is assumed<br />

to be equal to u * and d c is set equal to 2A L , then Equation 2-32 and Equation 2-18 are<br />

equivalent.<br />

With the approximations stated, this model only requires the specification of three<br />

parameters, namely u * , u r , and d c .<br />

2.2 Dispersion in the atmosphere<br />

Whereas deposition serves to permanently remove particles from the air stream,<br />

turbulent dispersion causes particles to dynamically redistribute through the mixed layer<br />

in the atmosphere. Dispersion causes pockets of air with comparatively high<br />

concentrations of particles to be diluted with pockets of air that have lower<br />

concentrations.<br />

There are a number of mathematical tools to describe dispersion in the<br />

atmosphere. Most commonly, in analogy to mixing length theory for momentum<br />

transport, it is quantified as a dispersion coefficient multiplied by a concentration<br />

gradient. Thus, the amount of particles crossing a horizontal plane is given by an equation<br />

of the form<br />

F<br />

z<br />

= −K<br />

zz<br />

∂C<br />

∂z<br />

Equation 2-33<br />

where F z is the flux of particles at a height z and is equal to the vertical (z-direction)<br />

dispersion coefficient K zz (m 2 /s) multiplied by the negative of the vertical concentration<br />

gradient (g/m 4 ). Similar equations are used to describe dispersion in the two horizontal<br />

directions utilizing the coefficients K xx and K yy which may be close in magnitude to one<br />

another but can be quite different from K zz . Horizontal dispersion may be ignored if the<br />

source is approximated by an infinitely long line (e.g. a long road).<br />

In the surface layer (nominally 5 – 50 m above ground), where the flux of<br />

momentum to the ground is constant with height, K zz is well-specified. At greater<br />

heights, measurements are more difficult, and the functional form of K zz is inferred from<br />

wind tunnel tests, theory, or numerical simulations.<br />

2-14


An alternative method for modeling diffusion is to specify the spread of a<br />

pollutant as a function of downwind distance and atmospheric stability. This is the<br />

approach used by a number of Gaussian Plume Models, including the ISC3 (USEPA,<br />

1995). The concentration of an airborne species is assumed to be normally distributed in<br />

each of the three ordinal directions, with standard deviations of σ x , σ y , and σ z For a<br />

continuous line source, dispersion in the horizontal plane can be ignored. The parameter<br />

σ z has been specified by a number of authors including Gifford (1961), Klug (1963),<br />

Turner (1969, 1970), and Martin (1976). The ISC3 model uses the formulation of Turner<br />

(1970). The Gaussian approach is useful because of its simplicity in implementation.<br />

However, it has its limitations, most notably, that the assumption of a normally<br />

distributed concentration profile may be too simple to adequately account for the range of<br />

land use and atmospheric conditions that can be encountered.<br />

Alternative methods for quantifying turbulent dispersion are available from the<br />

literature (e.g. Venkatram, 1993; Du and Venkatram, 1997; Schopflocher and Sullivan,<br />

2002), some of which may be more accurate than mixing-length or Gaussian models.<br />

2-15


2-16


3. MEASUREMENT <strong>OF</strong> DEPOSTION VELOCITY<br />

The goal of this WESTAR-funded, University of Utah subproject was to estimate<br />

the magnitude of particle deposition from vehicle-generated dust clouds downwind of an<br />

unpaved road. In this chapter, we discuss the methods and results of direct measurement<br />

of deposition velocity v d . Measurements were conducted in the “impact zone” (refer<br />

Figure 2-1). The measured values of v d are compared to values previously reported in the<br />

literature and analyzed in the context of the resistance model for deposition.<br />

Two types of surfaces for particle deposition, flat substrates and artificial<br />

vegetation, were analyzed with microscopic and gravimetric techniques. Comparison of<br />

these measurements with prior work suggests that there are challenges to bridging direct<br />

measurements with canopy-wide deposition velocities.<br />

3.1 Methods<br />

All the deposition velocity testing was conducted during April 2002 at Ft. Bliss<br />

TX, on a spur road south of Loyalty Lane at GPS coordinates 13 3 70 544 E, 35 28 119<br />

N. The field experiments were conducted in cooperation with the SERDP-funded<br />

projects CP1190 and CP1191. The site is arid with sparse vegetation cover and exposed<br />

sand. The prevailing winds were from the west and nearly perpendicular to the road.<br />

Road surface samples were collected and size distribution was measured by standard<br />

sieve analysis (refer Table 3-1).<br />

Table 3-1 Cumulative screen analysis for the soils samples collected before and after the sampling<br />

period<br />

Facility & GPS Units/<br />

Sample Site Location<br />

Well Test<br />

Road<br />

13 3 70 544 E<br />

35 28 119 N<br />

DRI/JMV ID Notes Silt %<br />

+ 3/8 + 4 + 20 + 30 + 50 + 100 + 200 - 200<br />

mesh mesh mesh mesh mesh mesh mesh mesh<br />

Heavy sand deposits<br />

K408-1A from previous day’s 7.0 1.00 1.00 1.00 0.97 0.85 0.64 0.36 0.07<br />

windstorm<br />

Well Test<br />

Road<br />

13 3 70 544 E<br />

35 28 119 N<br />

K419-TR-1<br />

After extensive driving<br />

compared to previous 4.1 1.00 0.99 0.98 0.87 0.81 0.59 0.28 0.04<br />

samples<br />

Dust Concentration was measured using a GRIMM (Dust Monitor Series 1.108®,<br />

GRIMM Technologies Inc., Douglasville, GA), which recorded particle-count<br />

concentration for 16 size ranges logging every six seconds. This is an optical instrument<br />

that determines particle size and count by light scattering. The sample inlet was placed<br />

next to the surfaces used for particle deposition sampling. A 2.5 m long PVC tube<br />

connected the inlet to the GRIMM instrument, which was placed with the data-logging<br />

computer in a sealed box. One test was conducted using a different GRIMM that had a<br />

direct sample inlet and that was programmed to report calculated mass. Quality checks<br />

for the GRIMM included an internal calibration cycle and cleaning of sample inlet head.<br />

The data from the GRIMM were used to obtain a particle number size distribution timeaveraged<br />

over the duration of each dust deposition experiment.<br />

3.1.1 Deposition on Flat Substrates<br />

A series of tests was conducted to measure the passive deposition of vehicle<br />

generated dust particles on flat surfaces placed near the road. The technique involved<br />

3-1


counting by scanning electron microscopy (SEM) the number of particles deposited on<br />

polycarbonate membranes which were exposed using a three-axis support stand (Figure<br />

3-1). Six samples corresponding to the stream-wise, crosswind and vertical directions<br />

(+x, -x, +y, -y, +z, and –z) could be exposed simultaneously. The +x direction was<br />

downwind perpendicular to the road and the +z direction was vertical upward. The<br />

substrates were exposed to dust from vehicles driving on the test road at speeds from 5 to<br />

50 mph. Alternative locations, exposure durations, and substrate materials were tested as<br />

discussed in method development, below. The best results were obtained on<br />

polycarbonate substrates (Millipore type GTTP, Fisher Scientific) located 5 m downwind<br />

of the road and exposed for 15 vehicle trips.<br />

6-Direction Passive Deposition<br />

Polycaronate<br />

Polycarbonate<br />

membrane on on<br />

Carbon Tape<br />

Petri Dish Detail<br />

Typical for 6<br />

Support Stand<br />

with Clamps<br />

Grimm<br />

Figure 3-1 The experimental setup of the passive deposition on flat surfaces for examination by<br />

electron microscopy. Three-axis (corresponding to the stream-wise, crosswind, and vertical<br />

components of wind direction i.e. +x, +y, +z, -x, -y and –z directions) passive collection substrates<br />

were mounted on a ring stand. Standard right hand coordinates are used with the positive x-axis<br />

downwind perpendicular to the road. A GRIMM inlet was placed right next to it to monitor the dust<br />

concentration. The substrates were placed 1 or 2m high and 5 or 10m away from the unpaved road.<br />

The size and number of particles deposited on the flat substrates were determined<br />

by quantitative electron microscopy. The flat substrates were mounted on double-sided<br />

carbon tape in a disposable 50 mm diameter Petri dish. Different substrate types were<br />

tested side by side in a single Petri dish. After exposure in the field the Petri dish covers<br />

were replaced and the set of samples was boxed for shipment to the laboratory. To<br />

prepare for microscopy, a section of the substrate assembly was cut from the dish and the<br />

bottom layer of conductive double-sided carbon tape was used to mount the sample on an<br />

aluminum SEM stub (Hitachi S-450®, Redding, CA) for gold coating. All samples were<br />

evaluated qualitatively for dust loading using the Hitachi S3000N® scanning electron<br />

microscope (Hitachi, Pleasanton, CA) located in the Material Science and Engineering<br />

3-2


Department at the University of Utah. The selected filters that passed QA underwent full<br />

quantitative microscopy to provide the images for particle counting.<br />

SEM images for counting were taken in a non-overlapping structured grid pattern<br />

at three magnifications (100X, 500X, 2500X). By combining data from multiple<br />

magnifications both the infrequent, large, high-mass particles and the ubiquitous, less<br />

massive sub-micron particles could be quantified. Particles d>20 µm were measured on<br />

100X images, particles 5


Note that all the size bins measured by electron microscopy and by the GRIMM<br />

are physical diameters and not the aerodynamic diameters. In addition, Equation 3-2<br />

requires that the concentration time average be calculated over the same interval as the<br />

exposure time.<br />

3.1.2 Deposition on Artificial Vegetation<br />

Dust deposition on artificial vegetation (AV) surfaces placed near the edge of the<br />

road was measured by collecting the deposited material, separating the particles by size,<br />

and weighing the material. Plastic "fir" Christmas garland was used since real vegetation<br />

is subject to artifacts from natural variation, biogenic debris, and moisture changes. The<br />

pre-washed AV was wrapped around a 17”x19” PVC frame and mounted on a support<br />

frame above a clean plastic tub (20"x 14"= 0.18 m 2 ) that was placed downwind of the<br />

road, Figure 3-2. The "fir" consisted of a central wire with close-spaced needle-covered<br />

branches. The multiple wraps of the garland around the frame created a dense but porous<br />

obstruction to the dust flow. Total surface area of the branches and needles was 3.7 m 2<br />

compared to the horizontal projected area of the assembly outline, which was 0.18 m 2 ,<br />

and the frontal projected area above the tub, which was 0.26 m 2 . For manufactured AV,<br />

total surface area could be determined accurately by measuring dimension components<br />

(needles, branches) and multiplying by number of components per length.<br />

Artificial<br />

Vegetation<br />

Grimm<br />

Collection<br />

Tub<br />

Figure 3-2. The experimental set up for passive deposition on artificial vegetation (Fir Christmas<br />

garland). The plastic foliage was wrapped around a PVC frame, pre-washed and installed above a<br />

clean plastic tub placed 5 m away from the road and 1 m above grade. A GRIMM inlet was placed<br />

alongside it to monitor the dust concentration.<br />

A control tub of similar dimensions was placed next to the AV to measure particle<br />

settling on horizontal surface in the absence of vegetation. During sampling runs, the AV<br />

was exposed to the vehicle-generated dust. Exposure ranged from 1 hour at 5 m from the<br />

road to 8 hours at 50 m from the road. After sampling, the assembly was covered and<br />

taken indoors, where the deposited material was collected in two fractions: the dry<br />

material dislodged by shaking the vegetation and the material removed by spray washing<br />

with distilled water.<br />

3-4


In the laboratory, the deposited material was weighed then separated into four size<br />

bins (d 30 µm, 10 d < 30 µm, 3 d < 10 µm and d < 3 µm) by gravity settling in<br />

water, based on the Stokes velocity difference between various size particles (see<br />

Equation 2-4). The dry and wet fractions collected in the field were analyzed separately<br />

and the results were added to get the total mass deposited. The size-separated fractions<br />

were dried and weighed and the amount of deposition on the AV in each size range was<br />

calculated by mass balance. The gravity separation method was validated in preliminary<br />

experiments using surrogate mixtures of particles with known size and the separations of<br />

the field samples were run in duplicate after splitting the sample using a riffler. The<br />

quality of the size separations obtained from the field samples was also verified using<br />

SEM.<br />

For the artificial vegetation (AV) experiments, the deposition velocity was<br />

determined using Equation 3-2 expressed on a mass basis. The GRIMM particle number<br />

data were converted to particle mass. The amount deposited was determined from<br />

weighing the 3 d < 10 µm and d < 3 µm wet separated fractions. The deposition<br />

velocity was calculated based on the horizontal projected area of the collector outline.<br />

3.1.3 Flat Substrate Method Testing<br />

This study of field measurement of vehicle-generated dust deposition required<br />

testing to evaluate alternative procedures, materials, sampling locations, sampling times,<br />

and data analysis methods. Twenty experiments measuring particle deposition on flat<br />

substrates were collected using various locations and numbers of vehicle trips. Samples<br />

were collected 1 or 2 m above grade at 5 and 10 m from the unpaved road. In addition, a<br />

blank sample was collected 100 m upwind from the road.<br />

Four substrates were tested to determine which surface gave the best data:<br />

1. Dry Polycarbonate filters (Isopore Membrane filter, 47 mm diameter,<br />

0.2 µm GTTP, Millipore)<br />

2. Oiled Polycarbonate filters, created by spraying DOW 316 (Slipicone®<br />

316 release agent, Dow Corning Corp., Midland, MI) on the filters and drying them<br />

3. Carbon tape (Catalog # 16073®, 8 mm, Ted Pella Inc, Redding, CA)<br />

4. Double sided clear tape (Scotch ®, 3M, St. Paul, MN)<br />

3.1.4 Vegetative Method Testing<br />

Six experiments were conducted to evaluate deposition velocities by the artificial<br />

vegetation method. A combination of different distances from the road (5 and 50 meters)<br />

and different artificial vegetation substrates (Fir Garland and artificial Ivy) were used.<br />

Deposition velocity was calculated using different areas in Equation 3-2:<br />

horizontal, total and projected. The horizontal projection area was used because it<br />

compares to what Slinn used in his calculations. It measures the enhancement of removal<br />

provided by vegetation. It is also easy to measure and is applicable to flux studies. The<br />

frontal projection area of the vegetation was used since it relates to the SEM flat substrate<br />

experiments. It represents the area of direct contact with the bulk of the dust cloud. The<br />

total area represents all surfaces available for deposition. Since not all exposed surfaces<br />

3-5


will achieve equal deposition rates due to shadowing and other issues, the overall<br />

deposition velocity calculated using this area would be low.<br />

Logs of all the flat substrate samples and the artificial vegetation experiments,<br />

including samples not completely analyzed due to method or QA issues, are given in<br />

Table 3-2 and Table 3-3 respectively. This includes method development tests, samples<br />

where data were lost, samples that did not pass QA, and spare field samples that were not<br />

analyzed. The incremental cost of collecting spare field samples during the campaign<br />

was low compared to the cost of microscopy and image analysis.<br />

3.2 Results: Deposition velocities measured by SEM and gravimetric techniques<br />

Microscopy counting of particles deposited on flat substrates gave a deposition<br />

velocity based on projected area ranging from 0.3 cm/s for submicron particles to 2 cm/s<br />

for 3 µm particles. The artificial vegetation suspended above a tub gave a deposition<br />

velocity based on horizontal surface area of about 10 cm/s. Deposition velocities by<br />

electron microscope particle counting of flat substrates and by gravimetric analysis of the<br />

deposits on artificial vegetation are presented in Figure 3-3. These values are also<br />

compared to predictions for dry deposition into a forest canopy in a theoretical study by<br />

Slinn (1982). The model prediction curve plotted in Figure 3-3 corresponds to a wind<br />

velocity of 5 m/s and a value of gamma (a measure of wind profile in the canopy) of 3.5,<br />

which represent the middle of the range considered by Slinn. A major difference is that<br />

Slinn's model calculates vertical flux from the atmosphere to a horizontal area of forest<br />

canopy while the data from this study are deposition from a horizontally moving cloud<br />

impinging on the projected area of a surface. Figure 3-4 gives the directional variation in<br />

the deposition velocities on flat surfaces (polycarbonate filter) normalized by the<br />

deposition velocity measured on the horizontal surface facing up (+z). The flux on the +z<br />

surface corresponds to the deposition velocity measured by Slinn. In addition,<br />

gravitational settling would be expected to deposit particles on this surface only. The<br />

vertical plane surfaces (-x, +y, -y) show enhanced deposition due to impaction caused by<br />

wind and vehicle wake whereas the deposition on surfaces facing down and away from<br />

the road (+x, -z) is presumed to be due to turbulence. Figure 3-5 shows the directional<br />

deposition velocities over the size range studied.<br />

Complete analysis was done on four samples (dry polycarbonate filters). They<br />

were SEM ID 12, 17, 18 and 19. SEM 17, 18 and 19 were 5 m away from the road and<br />

SEM 12 was 10 m away from the road. A grand average deposition velocity (17, 18, 19)<br />

curve versus each size bin plot is given in Figure 3-6. The particle concentration from<br />

the GRIMM measured coincident with the four SEM samples has been plotted in Figure<br />

3-7. There appear to be systematic differences between the size distribution measured<br />

with a direct inlet (SEM 12) and the size distributions measured with an inlet tube.<br />

Particle losses increase the calculated deposition velocity, but data for d < 5 µm appear to<br />

be valid.<br />

Testing of alternative methods increased the chance of success for this experiment<br />

and provided data on materials and sampling strategies that can be used to design future<br />

3-6


Table 3-2. Log of all flat substrate samples<br />

Date SEM ID #<br />

Location<br />

(x, y)<br />

# Of<br />

Trips<br />

Vehicle type<br />

03/26/02 1N 10, 2 10 Chevy Impala<br />

Comments<br />

03/26/02 1F 40, 2 20 Chevy Impala Very lightly loaded<br />

03/27/02 2N 10, 2 10 Chevy Impala<br />

Only PM10 mass, no<br />

concentration by size<br />

03/28/02 3N 10, 2 10 Chevy Impala Very lightly loaded<br />

03/28/02 3F 40, 2 20 Chevy Impala Very lightly loaded<br />

04/08/02 1– (+x, +y, +z)<br />

-100, 1<br />

-<br />

Patriot<br />

Battery<br />

04/08/02 2 5, 1 40 Voyager<br />

Missile<br />

04/11/02 3 5, 1 17 Dodge Caravan<br />

04/12/02 4 5, 1 20 Voyager<br />

04/12/02 5 5, 2 20 Voyager<br />

Decided not to use oiled filters<br />

for counting<br />

Decided not to use oiled filters<br />

for counting<br />

-X filter lost, oiled base<br />

formed cake-like structure<br />

Decided not to use oiled filters<br />

for counting<br />

Decided not to use oiled filters<br />

for counting<br />

04/12/02 8 5, 1 20 Voyager Very lightly loaded<br />

04/13/02 1- (-x, -y, -z) -100, 1 -<br />

1 hr upwind<br />

control<br />

04/16/02 13 - - Voyager Not analyzed<br />

Decided not to use oiled filters<br />

for counting<br />

13-1 - 0 - Filter pores clearly visible<br />

13-2 - 0 -<br />

Oil forms a cracked cake-like<br />

base<br />

13-3 (-x) 5, 1 1 - Very lightly loaded<br />

13-4 (-x) 5, 1 5 - Loading greater than 13-3<br />

13-5 (-x) 5, 1 5 -<br />

Cake base formed, Particles<br />

barely distinguishable<br />

13-6 (-x) 5, 1 25 - Loading greater than 13-4<br />

04/17/02 7 5, 1 50 2 ½ ton truck<br />

04/17/02 14 5, 1 50 2 ½ ton truck<br />

Decided not to use oiled filters<br />

for counting<br />

Carbon tape seems to be a<br />

good substrate for future SEM<br />

experiments<br />

04/17/02 12 10, 2.3 50 2 ½ ton truck Deposition velocity calculated<br />

04/18/02 15 -100, 1<br />

1 hr<br />

control<br />

Deposition velocity calculated<br />

04/18/02 16 5, 1 5 Hemmet Spare – not analyzed<br />

04/18/02 17 5, 1 25 Hemmet Deposition velocity calculated.<br />

04/19/02 18 5, 2 15<br />

04/19/02 19– (+x, +y, +z) 5, 1 15<br />

04/19/02 20 50, 1.5 15<br />

18 wheel Flatbed<br />

military truck<br />

18 wheel Flatbed<br />

military truck<br />

18 wheel Flatbed<br />

military truck<br />

Deposition velocity calculated.<br />

Deposition velocity calculated.<br />

Spare – not analyzed<br />

studies. Four different substrates were tested for flat surface sampling. The reported<br />

deposition velocities are based on SEM examination of dry polycarbonate filter<br />

membranes. The oiled filters developed a cracked surface film and it was difficult to<br />

distinguish the particles. The sample collected on carbon tape indicated that this might<br />

be a suitable substrate. The clear tape has high collection efficiency for large size<br />

particles, but it is difficult to distinguish the smaller size particles from artifacts on the<br />

adhesive surface. A comparison of particle counts done on a clear tape and a dry filter<br />

sample is given in Figure 3-8.<br />

3-7


Table 3-3. Log of all field vegetative samples<br />

Sample ID Date<br />

Location Total Run Time Comments<br />

(x, y) - m (min)<br />

F = Fir C = control<br />

F001 04/08/02 5, 1 20 Problem in lab processing<br />

5, 1 52 Unable to open GRIMM data<br />

C001 04/08/02 5, 1 52 Unable to open GRIMM data<br />

F002 04/12/02 5, 1 460 Deposition velocity result<br />

C002 04/12/02 5, 1 460 Deposition velocity result<br />

F003 04/17/02 5, 0.75 194 Gravimetric data suspect<br />

C003 04/17/02 5, 0.75 194 Gravimetric data suspect<br />

F004 04/17/02 50, 0.75 185<br />

04/18/02 50, 0.75 330<br />

I001 04/18/02 5, 0.75 220<br />

Deposition velocity result<br />

No GRIMM data for d


16<br />

Normalized Deposition Velocity<br />

11<br />

6<br />

1<br />

-4<br />

-9<br />

NX NY PY NZ PX<br />

Direction<br />

Figure 3-4. Directional variation in deposition velocity normal to a flat surface. Data given are mean<br />

and one standard deviation, normalized by the deposition velocity on the horizontal surface facing up<br />

(PZ) averaged for the range 0.5-7.5 m (physical diameter) based on four determinations (SEM 12,<br />

17, 18, 19). The vertical surfaces (NX, NY, PY) show enhanced deposition due to impaction.<br />

Deposition on surface facing down (NZ) and away from the road (PX) is presumed to be due to<br />

turbulence.<br />

100<br />

Normalized Deposition Velocity<br />

10<br />

1<br />

0.1<br />

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15<br />

Physical Diameter - microns<br />

NX NY NZ PX PY PZ<br />

Figure 3-5. Directional variation in deposition velocity normal to a flat surface as a function of<br />

particle size. Data are the mean of four determinations (SEM 12, 17, 8, 19), normalized by the<br />

deposition velocity on the horizontal surface facing up (PZ). The plot indicates no consistent size<br />

dependence for particles less than 10 µm in diameter.<br />

3-9


100<br />

Deposition velocity - cm/s<br />

10<br />

1<br />

0.1<br />

0.3<br />

0.4<br />

0.5<br />

0.65<br />

0.8<br />

1<br />

1.6<br />

2<br />

3<br />

4<br />

5<br />

7.5<br />

10<br />

15<br />

20<br />

Physical Diameter - microns<br />

Grand Avg of all the filters excluding SEM 12<br />

Stdev<br />

Figure 3-6. . The grand average deposition velocity (SEM 17, 18, 19) for the SEM samples versus the<br />

particle size. The dashed line represents one standard deviation of the three determinations.<br />

Deposition velocity and standard deviation of replicates data outside the 0.5 m to 5 m range are<br />

not shown due to uncertainty in the accuracy of the data. For d5 m, the results are uncertain due to inlet line<br />

loss.<br />

1.E+10<br />

Concentration - N/m3<br />

1.E+08<br />

1.E+06<br />

1.E+04<br />

1.E+02<br />

1.E+00<br />

0.3<br />

0.4<br />

0.5<br />

0.65<br />

0.8<br />

1<br />

1.6<br />

2<br />

3<br />

4<br />

Physical Diameter - microns<br />

5<br />

7.5<br />

10<br />

15<br />

20<br />

DRI - 12 Utah - 17 Utah - 18, 19<br />

Figure 3-7. Variation in particle concentration between sampling days and between instruments with<br />

different inlet configurations. SEM 12 data were obtained from a GRIMM with a direct sample inlet<br />

and SEM 17, 18, 19 data were obtained from a GRIMM with a 2.5 m long PVC tubing attached to its<br />

sample inlet. For diameters less than 5 m (physical), both the inlets agree but at particle diameters<br />

greater than 5 m, the GRIMM with the long inlet tube exhibited sampling losses. Particle losses<br />

affect deposition velocity calculations.<br />

3-10


1.E+05<br />

Deposition Velocity - cm/s<br />

1.E+03<br />

1.E+01<br />

1.E-01<br />

1.E-03<br />

0.1 1 10 100<br />

Physical Diameter - microns<br />

Clear Tape<br />

Dry Filter<br />

Figure 3-8. Deposition velocities calculated from SEM particle count for a dry polycarbonate<br />

membrane and a clear tape surface that were collected simultaneously (SEM 18). From 5 m to 10<br />

m both the substrates give similar particle counts. The count is increased (and therefore deposition<br />

velocity is increased) for particles d20 m) that fell off the dry substrate (data not shown).<br />

Deposition velocity for the vegetative samples was calculated using three<br />

different area counts, given in Table 3-4. Using the horizontal projection area resulted in<br />

a higher deposition velocity for the AV and control combination when compared to the<br />

control alone. The first calculation given in Table 3-4 used the entire area of the AV plus<br />

the area of the tub underneath it. The calculated deposition velocity for d< 3 µm using<br />

this area was 0.7 cm/s compared to 8.3 cm/s for the control. The second deposition<br />

velocity calculation used the frontal projection area of the AV plus the area of the tub<br />

underneath. The resulting deposition velocity for d< 3 µm using this area was 6.5 cm/s<br />

vs. 8.3 cm/s for the control. The final calculation used the horizontal AV projected area<br />

plus the area of the tub (essentially the area of the tub). The resulting deposition velocity<br />

for d


Table 3-4. Deposition velocity calculations for vegetative samples using different areas:<br />

1. Deposition velocity calculation using total area:<br />

Sample ID<br />

Size Range<br />

(physical)-m<br />

Particle<br />

Concentration-g/m 3 Gravimetric Wt. - g Area – m2 Time - sec Velocity–cm/s Comments<br />

F001 d


are deposition from a horizontally moving cloud impinging on the projected area of a<br />

surface. Slinn reasoned that because momentum transfer from the wind to ground is<br />

higher over a forest, then downward particle transfer should also be higher, and<br />

developed a model that used the wind velocity profile in the forest canopy as a parameter.<br />

The experiments in this study are more closely analogous to the deposition predicted by<br />

models of single fiber filtration efficiency.<br />

Table 3-5. Deposition Velocity calculated for vegetative samples<br />

Sample ID<br />

Location<br />

(x, y) - m<br />

Size Range Particle<br />

(physical)- µ Concentration- g/m 3 Gravimetric Wt. - g Area – m2 Time - sec<br />

F002 5, 1 d


polycarbonate surfaces and those calculated for similar conditions according to Equation<br />

3-3 and Equation 2-25. The difference between measured and predicted deposition<br />

velocities is quite large, ranging from one order of magnitude for large particles to two<br />

orders of magnitude for small particles. This difference is expected to be manifested in<br />

the overall deposition velocity (I.e. considering both r b and r a ) since the resistance to<br />

transport through the surface layer r a is very small compared to r b . That is, under the<br />

atmospheric conditions of Ft. Bliss, the deposition velocity is controlled almost entirely<br />

by the transport through the quasi-laminar sublayer.<br />

100<br />

Deposition Velocity (cm/s)<br />

10<br />

1<br />

0.1<br />

Vd=1/rb calculated<br />

Ft Bliss<br />

Expon. (Ft Bliss)<br />

0.01<br />

0 1 2 3 4 5 6 7 8 9 10 11<br />

Aerodynamic Dp (microns)<br />

Figure 3-9. Measured deposition velocities on polycarbonate substrates and deposition velocities<br />

calculated based on the boundary layer resistance r b only vs. particle aerodynamic diameter. The<br />

average u * for the sample periods represented by the measured data and used for the theoretical<br />

prediction was 0.35 m/s.<br />

There is a subtle point that explains some of the differences between the measured<br />

and modeled deposition velocities. The measured values in this study represent<br />

deposition to a surrogate surface (artificial plant or flat substrate). It is likely that the<br />

density of the vegetation in the canopy, characterized perhaps by the fraction of ground<br />

area covered by plants should enter in the calculation of overall deposition to the canopy.<br />

That is, the surrogate surfaces used in this study may adequately simulate the removal of<br />

particles by individual vegetative elements, but do not account for the density of those<br />

elements. Some work in this area has been done by Raupach and Leys (1999), though<br />

testing of their hypotheses against data is not complete.<br />

3.4 Conclusions<br />

An important point, directly relevant to this WESTAR project as a whole, is that<br />

the quantification of the deposition velocity is not a straightforward matter. Most models,<br />

including the ones discussed in (Chapter 2) are only tested against limited data sets, most<br />

notably those of Chamberlain (1967). One reason for this is that direct measurement of<br />

deposition velocities is quite difficult, and each of the available techniques has some<br />

significant shortcomings as noted by Dabberdt et al. (1993). Though use of surrogate<br />

surfaces for deposition is a common approach (e.g. Wu and Davidson ,1992; Etyemezian<br />

et al., 1999), it also suffers from shortcomings. Most notably, the surrogate surface is a<br />

3-14


single element that is intended to induce similar transfer of material (momentum,<br />

particles, gases, etc) as an entire canopy. While there have been recent novel approaches<br />

that attempt to measure deposition in-situ (e.g. Gillies et al., 2001), at present there<br />

remain substantial uncertainties in the true value of deposition velocities. The solution to<br />

the difficulty of quantifying deposition is outside the scope of the present study. We note<br />

simply that values of v d calculated by any of the methods described in Chapter 2 may be<br />

erroneous in some cases by a factor of two or more, based on the input parameters alone.<br />

The results of this portion of the study, which utilized surrogate surfaces for<br />

deposition, indicate that measured deposition velocities are difficult to relate to<br />

theoretical formulation. This is illustrative of the ongoing challenge in reconciling data<br />

with theory in the field of dry deposition.<br />

3-15


3-16


4. MODELS USED TO STUDY PARTICLE REMOVAL BY DRY<br />

<strong>DEPOSITION</strong> DOWNWIND <strong>OF</strong> AN UNPAVED ROAD<br />

One of the objectives of this study was to provide an analysis of the Gillette Box<br />

Model in terms of its utility in determining the removal of dust particles downwind of an<br />

unpaved road. This was accomplished by considering two other models, one based on<br />

the Atmospheric Diffusion Equation (ADE), and the other based on the Gaussian plume<br />

approximation. The former provides the underlying basis for many regional air quality<br />

models, while the latter is used by the EPA’s ISC3 dispersion utility. In this chapter, we<br />

present the ADE, Gaussian, and the Gillette Box Models. Simulations using all three<br />

models are compared to one another. Special consideration is given to the assumptions<br />

of the Box Model, and the conditions for their validity.<br />

4.1 Presentation of Models<br />

4.1.1 The one-dimensional Atmospheric Diffusion Equation (ADE)<br />

The Atmospheric Diffusion Equation (ADE) describes the transport and removal<br />

of airborne material. It is derived from a mass balance on a control volume (CV), where<br />

species (gases or small particles) are allowed to diffuse in and out of the CV by the<br />

turbulent motions of the atmosphere, move through the CV by advection, and react<br />

chemically. Invoking the K-theory for turbulent dispersion, the full Eulerian equation is:<br />

∂c<br />

∂c<br />

∂ ∂c<br />

∂ ∂c<br />

∂ ∂c<br />

<br />

+ u = K<br />

xx + K<br />

yy<br />

+ K<br />

zz + S<br />

∂t<br />

∂x<br />

∂x<br />

∂x<br />

∂y<br />

∂y<br />

∂z<br />

∂z<br />

<br />

Equation 4-1<br />

where c is the concentration of the species of interest, u is the velocity in the x-direction,<br />

K ii are the turbulent diffusivities in the x, y, and z directions, and S is a chemical reaction<br />

term. For an approximate continuous line source oriented in the y-direction, the net<br />

transverse dispersion is zero, and the concentration at any one location is invariant with<br />

time. In the x-direction, for any appreciable value of u (u > 1 m/s), advection is much<br />

greater than dispersion. Dust particles do not react chemically, and the source/sink term<br />

S is also set to zero. In summary,<br />

K<br />

∂c<br />

∂c<br />

∂c<br />

∂ ∂c<br />

<br />

= 0 , = 0, u >> K<br />

xx ,<br />

S = 0<br />

∂y<br />

∂t<br />

∂x<br />

∂x<br />

∂x<br />

<br />

yy<br />

.<br />

Equation 4-2<br />

reducing Equation 4-1 to<br />

∂c<br />

∂ ∂c<br />

<br />

u = K<br />

zz .<br />

∂x<br />

∂z<br />

∂z<br />

<br />

Equation 4-3<br />

4-1


At this point, it is possible to remove the dependence of c on x if it is assumed that<br />

the wind speed u is nearly constant with height and by expressing u as dx/dt. This casts<br />

the ADE in a Lagrangian frame where the concentration c is dependent only on the height<br />

z and the time since emission, yielding the one-dimensional, time dependent ADE:<br />

∂c<br />

∂ <br />

= K<br />

∂t<br />

∂z<br />

<br />

∂c<br />

<br />

<br />

∂z<br />

<br />

zz<br />

.<br />

Equation 4-4<br />

Note that the time derivative in Equation 4-4 is not the same as that in Equation 4-1.<br />

Specifically, in the latter equation δc/δt is evaluated at a stationary point in space,<br />

whereas in the former, it is implied that the derivative is evaluated in a frame of reference<br />

that moves downwind with the plume.<br />

The downwind distance at a given time can be estimated by multiplying the time<br />

with the wind speed at a 10 m height, obtained, for example, from a logarithmic profile<br />

(e.g. Equation 2-9). Values of K zz as a function of height have been provided by a<br />

number of investigators. The functional forms used in this model are<br />

Unstable Conditions (Lamb and Duran, 1977):<br />

K<br />

zz<br />

w z<br />

*<br />

i<br />

4<br />

<br />

1<br />

3<br />

<br />

4<br />

<br />

<br />

z z z<br />

2.5<br />

<br />

<br />

κ<br />

1<br />

− 15 → 0 ≤ < 0.05<br />

zi<br />

L zi<br />

<br />

<br />

<br />

2<br />

<br />

z z <br />

<br />

0.021+<br />

0.408<br />

<br />

+ 1.351<br />

<br />

↵ <br />

<br />

zi<br />

zi<br />

<br />

<br />

<br />

3<br />

4<br />

<br />

z z z<br />

= −<br />

4.096<br />

<br />

+ 2.560<br />

<br />

→ 0.05 ≤ < 0.6<br />

zi<br />

zi<br />

zi<br />

<br />

<br />

<br />

<br />

<br />

z z<br />

0.2exp6<br />

−10<br />

→ ≤ < <br />

<br />

<br />

0.6 1.1<br />

z <br />

<br />

i<br />

zi<br />

<br />

<br />

z<br />

<br />

<br />

0.0013 → ≥ 1.1<br />

z<br />

<br />

i<br />

<br />

<br />

<br />

<br />

Equation 4-5<br />

Neutral Conditions (Shir, 1973):<br />

8 fz <br />

K zz = κ u<br />

−<br />

<br />

*<br />

z exp<br />

u*<br />

<br />

Equation 4-6<br />

Stable conditions (Businger and Arya):<br />

4-2


κu <br />

( ) <br />

*<br />

z<br />

8 fz<br />

K zz<br />

=<br />

exp<br />

−<br />

0.74<br />

+ 4.7 z L u*<br />

<br />

Equation 4-7<br />

where z i is the mixing height, L is the Monin-Obukhov length scale for turbulence, u * is<br />

the friction velocity, f is the Coriolis parameter, κ is the Von Karman constant (0.41), and<br />

w * is the convective velocity scale given approximately by<br />

w<br />

1<br />

3<br />

zi<br />

<br />

*<br />

= u*<br />

− L <br />

(L is negative for unstable conditions).<br />

Equation 4-8<br />

To solve Equation 4-4, it is necessary to specify boundary conditions on z and<br />

initial conditions for t=0. The top of the modeling domain is set high enough (z top > 2×z i )<br />

so that there is no net flux of particles. That is,<br />

∂c<br />

K<br />

zz<br />

= 0 , z = z top<br />

.<br />

∂z<br />

Equation 4-9<br />

At the ground, the downward flux of particles by dispersion must equal the removal by<br />

deposition<br />

∂c<br />

K<br />

zz<br />

= vd<br />

c, z = 0<br />

∂z<br />

Equation 4-10<br />

where v d is calculated from Equations 2-3, 2-25, and 2-26 with r a set to zero, since the<br />

depositing particles are already at the ground (z=0).<br />

The initial condition is simply that the concentration of particles is uniform up to<br />

a specified injection height. That is<br />

c z)<br />

= c , z < IH<br />

(<br />

0<br />

c(<br />

z)<br />

= 0, z ≥ IH<br />

Equation 4-11<br />

In practice, the transition across z=IH is made mathematically smooth to avoid<br />

difficulties in the numerical solution. With these initial and boundary conditions, the<br />

ADE is solved numerically at 1,000 time steps spaced logarithmically from 1 second to<br />

10,000 seconds.<br />

4-3


4.1.2 The Integrated Source Complex Model 3 (ISC3)<br />

The ISC3 Short-Term dispersion model (ISC3 hereafter) is built on the<br />

approximation that a plume dispersing in the atmosphere assumes a shape similar to a<br />

Gaussian distribution (USEPA, 1995). The ISC3 utility allows for simulation of point,<br />

area, and volume sources with the aid of hourly meteorological data. Point and volume<br />

sources are treated in essentially the same way; volume sources are assumed to be point<br />

sources that originated at some distance upwind. Area sources are treated as multiple<br />

point sources. Line sources can be approximated by using multiple area or volume<br />

sources. The ISC3 also includes a number of optional parameters that can be used for<br />

adjusting the height of a plume (e.g. in the case of hot or high-speed stack gases),<br />

accounting for building downwash, and dispersion in complex terrain. For simulating a<br />

road dust plume generated by the movement of a vehicle on an unpaved road, the volume<br />

source approach is most directly applicable; accordingly, we will restrict the presentation<br />

of the ISC3 model to volume sources and omit material that is not directly applicable to<br />

unpaved road dust emissions.<br />

The gaussian representation of the concentration profile in the ISC3 model is<br />

actually a result of an approximate analytical solution to the Atmospheric Diffusion<br />

Equation under certain atmospheric conditions. It works best for sources that are high<br />

enough above ground that the dispersion parameter (the equivalent of K zz in Equation<br />

4-3) can be assumed roughly constant. The applicability and limitations of Gaussian<br />

approaches to dispersion modeling have been considered by other investigators (e.g.<br />

Seinfeld and Pandis, 1999) and an in depth discussion is omitted here.<br />

The basic equation that is solved for the ground-level concentration C (µg/m 3 )<br />

downwind of a point source is<br />

C =<br />

−6<br />

( 1×<br />

10 )<br />

QVD<br />

<br />

<br />

y<br />

exp − 0.5 <br />

2πu<br />

<br />

sσ<br />

yσ<br />

z<br />

σ<br />

y <br />

2<br />

<br />

<br />

<br />

<br />

Equation 4-12<br />

where<br />

Q = pollutant emission rate (g/s)<br />

V = a vertical term that includes the effects of ground reflection, dry deposition, and<br />

vertical mixing<br />

D = chemical decay term (assumed equal to 1 since road dust is non-reactive)<br />

σ y , σ z = standard deviation of vertical and lateral concentration distribution<br />

u s = mean wind speed at release height.<br />

For a continuous line source of non-reactive material (i.e. D=1); Equation 4-12 may be<br />

integrated over -∞ < y < +∞ to obtain<br />

4-4


C =<br />

−6<br />

( 1×<br />

10 )<br />

QV <br />

2π<br />

u s<br />

σ<br />

z<br />

Equation 4-13<br />

where the dot over the Q indicates that the source strength is expressed in terms of a unit<br />

crosswind distance (i.e. (g/s)/m). The most important parameter in the ISC3 model with<br />

regard to a continuous line source is the vertical standard deviation σ z. In the ADE,<br />

turbulent dispersion is calculated from the vertical gradient of the concentration whereas<br />

in the ISC3, σ z. completely specifies the concentration distribution and therefore, also the<br />

vertical gradient. This is accomplished through the vertical term V in Equation 4-13<br />

which, in the absence of deposition is specified as<br />

V<br />

z <br />

r<br />

− h<br />

= exp−<br />

0.5<br />

<br />

<br />

<br />

σ<br />

z <br />

2<br />

z <br />

r<br />

+ h<br />

+ exp−<br />

0.5<br />

<br />

<br />

<br />

<br />

σ<br />

z <br />

2<br />

<br />

+<br />

<br />

[.....]<br />

Equation 4-14<br />

where z r is the height at which the concentration is to be evaluated, and h is the initial<br />

height of the release. To account for the reflection of the plume at ground level, the first<br />

two terms on the RHS of the equation actually represent two sources, one located at h,<br />

and one at –h. For road dust, it may be assumed that the release height of the plume is<br />

the ground. The release height is different from the initial vertical depth of the plume<br />

(discussed below). The unspecified bracketed term on the far RHS of Equation 4-14 is<br />

used to account for multiple reflections between the ground and the mixing height. This<br />

term can be ignored for a ground-level release provided that the analysis does not proceed<br />

too far downwind (i.e. σ z.


When a vehicle passes over a road, the plume generated behind the vehicle has a<br />

discrete depth in the vertical direction. This depth is a measure of the initial breadth of<br />

the plume and is different from the plume release height. To account for the fact that the<br />

dust plume is initially dispersed by the turbulent wake of the vehicle, a “virtual” distance<br />

is added to the value of x in Equation 4-15. For example, the virtual distance x 0 is<br />

calculated by solving the equation for the initial value σ z0 . Since σ z is the standard<br />

deviation of the concentration in the vertical direction assuming a gaussian (Normal)<br />

distribution, then 95% of the plume is initially below the height 2×σ z0 . Therefore, we<br />

may approximately define σ z0 as one half the “injection height” or the height of the<br />

influence of turbulence generated in the wake of a vehicle. The concept of the “injection<br />

height” is discussed in detail in Chapter 5.<br />

4.1.3 The Gillette Box Model This section of the text was provided by Dr. Dale<br />

Gillette (NOAA) with some minor adjustments by DRI<br />

4.1.3.1 Formulation and solution<br />

A “lumped control volume” approach was used by Gillette to gain understanding of<br />

fugitive dust sources. This case considers dust generated from a road surface. By letting<br />

the road be directed into and out of the page while wind is directed from right to left,<br />

two-dimensionality or symmetry in the direction into the page was invoked. That is,<br />

fluxes are equal into and out of the direction into the page. Figure 4-1 shows the<br />

geometry of the control volume. A dirt road exists at the right side of the control volume.<br />

To the left of the road is a surface that is grass or shrub-covered and does not emit<br />

particles. The ceiling of the control volume is the surface of primary interest as to<br />

vertical flux.<br />

dm/dt up<br />

Wind<br />

dm/dt ceil<br />

dm/dt ambout<br />

dm/dt ambin<br />

dm/dt road<br />

X<br />

X=X o<br />

dm/dt depos<br />

dm/dt ceil<br />

dm/dt ambin<br />

dm/dt out<br />

dm/dt road<br />

∆Z<br />

Z=0<br />

∆L=1<br />

Figure 4-1. Control Volume for Fugitive Dust Model Depicting Vertical and Horizontal Fluxes.<br />

The quantities shown in the figure are as follows:<br />

4-6


oad.<br />

M is the mass of particles in the control volume (CV),<br />

dm up /dt is the mass per unit time passing out vertically at the top of the CV,<br />

dm depos /dt is the mass per unit time depositing to the floor of the CV,<br />

dm ambout /dt is the mass per unit time passing out of the CV through the left wall,<br />

dm ambin /dt is the mass per unit time passing into the CV from the right wall,<br />

dm road /dt is the mass per unit time emitted from the road, and<br />

dm ceil /dt is the mass per unit time passing out of the ceiling directly above the<br />

A conservation of mass equation can be written for the control volume shown in<br />

the figure as:<br />

dM/dt + dm up /dt + dm depos /dt + dm ambout /dt - dm ambin /dt - dm road /dt + dm ceil /dt = 0<br />

Equation 4-16<br />

where, M is the mass of particles in the control volume (CV),<br />

dm up /dt is the mass per unit time passing out vertically at the top of the CV,<br />

dm depos /dt is the mass per unit time depositing to the floor of the CV,<br />

dm ambout /dt is the mass per unit time passing out of the CV through the left wall,<br />

dm ambin /dt is the mass per unit time passing into the CV from the right wall,<br />

dm road /dt is the mass per unit time emitted from the road, and<br />

dm ceil /dt is the mass per unit time passing out of the ceiling directly above the<br />

road.<br />

A simplifying assumption is that of steady state emissions, that is, dM/dt = 0. For this<br />

case:<br />

dm up /dt = - dm depos /dt - dm ambout /dt + dm ambin /dt + dm road /dt - dm ceil /dt<br />

4.1.3.1a<br />

road<br />

Equation 4-17<br />

Relationship of mass per unit time to the horizontal mass flux from the<br />

A subvolume of the control volume is shown in Figure 4-1 as that part that<br />

contains the road but no area downwind of the road. The left wall of this part of the<br />

control volume is the surface through which all of the dust emitted from the road passes.<br />

That is, we specify that none of the road dust is part of the ceiling flux (dm ceil /dt). For the<br />

condition of steady state, a description of the conservation of mass for this part of the<br />

control volume is:<br />

dm road /dt = dm out /dt - dm ambin /dt + dm ceil /dt<br />

Equation 4-18<br />

where dm out /dt is the horizontal mass per unit time passing out through an imaginary wall<br />

just left of the left edge of the road reaching from the surface to the top of the control<br />

4-7


volume, and dm ceil /dt is the mass per unit time passing through the ceiling of the control<br />

volume Because the road dust is usually the overwhelming source of dust, that is,<br />

dm road /dt >> dm ambin /dt + dm ceil /dt , the horizontal flux dm out /dt for these conditions is<br />

approximately:<br />

dm road /dt = dm out /dt<br />

which is the horizontal mass flux of dust from the road.<br />

Equation 4-19<br />

4.1.3.1b Deposition at the floor and vertical flux of dust through the ceiling of the<br />

control volume downwind of the road<br />

The total loss of material diffusing vertically through the ceiling and depositing<br />

on the floor of the control volume to the left of the road may be calculated by first<br />

calculating the effective concentration at position x to the left of x 0 (i.e., the left edge of<br />

the road). We use the equation:<br />

V<br />

dC(<br />

x)<br />

dx<br />

=<br />

− C(<br />

x)[<br />

V<br />

∆z<br />

d<br />

+ K]<br />

Equation 4-20<br />

where, V is the wind speed that carries the dust through the control volume (CV),<br />

C(x) is the concentration of dust mass at position x in the CV,<br />

x is the horizontal position in the CV, increasing to the left,<br />

z is the vertical position in the CV, increasing from the floor to the ceiling,<br />

V d is the deposition velocity, and<br />

K is a coefficient having the dimensions of velocity.<br />

A solution to Equation 4-20 is:<br />

dmroad<br />

= Vd<br />

+ K<br />

C(<br />

x)<br />

dt<br />

exp−<br />

x − x<br />

V∆z∆L<br />

V∆z<br />

o<br />

[ ] ( )<br />

<br />

Equation 4-21<br />

Multiplying C(x) by V d L (i.e., the deposition velocity times the unit length of the road)<br />

and integrating with respect to x from x = x 0 to x = gives:<br />

dm<br />

depos<br />

dt<br />

V<br />

=<br />

d<br />

dm<br />

dt<br />

( V + K )<br />

d<br />

road<br />

Equation 4-22<br />

4-8


By making the approximation that dm ambout /dt - dm ambin /dt 0 and ignoring dm ceil /dt we<br />

may rewrite Equation 4-17 as:<br />

dmup<br />

dm dm<br />

road depos<br />

= −<br />

dt dt dt<br />

Equation 4-23<br />

Watson (personal communication, 2000) stated that the condition dm ambout /dt - dm ambin /dt<br />

0 was observed in the field for a distance x of about 200 meters in the absence of any<br />

intervening dust sources.<br />

4.1.3.1c Ratio of vertical flux of road dust into the atmosphere to the horizontal<br />

flux of road dust<br />

The ratio of the mass per unit time emitted through the ceiling of the CV<br />

downwind of the road (dm up /dt) to the mass emitted from the road (dm road /dt) is called the<br />

“transportable fraction” and represents the fraction of emitted road dust that can be<br />

considered to be regionally transported. This fraction Φ is given in Equation 4-24.<br />

dmup<br />

dt <br />

Φ = =<br />

dm<br />

1<br />

−<br />

road <br />

dt<br />

V<br />

K<br />

( V + K ) ( V + K )<br />

d<br />

d<br />

<br />

=<br />

<br />

d<br />

Equation 4-24<br />

The value of K was derived by using data of Porch and Gillette (1977) and<br />

(Gillette, 1974). Porch and Gillette (1977) provided data on fast-response concentrations<br />

of diffusing dust simultaneously taken with fluctuation vertical wind speeds at the same<br />

location. Analysis of the high-speed data showed that the aerosol flux could be expressed<br />

approximately as 0.04u * [C] where [C] is the mean mass concentration. Analysis of<br />

gradients of dust concentration in wind-eroding fields (Gillette, 1974) showed that the<br />

aerosol flux could be expressed as approximately 0.07u * [C]. The mean of the<br />

coefficients of [C] for these analyses gave a value of 0.06 u * which is the suggested<br />

value for K.<br />

Using Equation 4-24 with K equal to 0.06 u * , yields the “transportable fraction”<br />

of fugitive dust emitted from the road. Practical application of Equation 4-24 can use<br />

values of V d chosen to represent particle size and environmental conditions, for example,<br />

those of Slinn (1982). Values of u * are chosen by the user to represent the environmental<br />

conditions of interest.<br />

4.1.3.2 Discussion<br />

The expression given in Equation 4-24 may explain part of why observed<br />

concentrations are smaller than that predicted by regional scale models that do not correct<br />

for the transportable fraction of emitted dust but rather use the entire amount of dust<br />

emitted by roads. However, there are limitations in applying this model to natural<br />

4-9


situations. Some of these limitations are listed here with a detailed discussion given in a<br />

later section.<br />

1. If the flux through the ceiling above the road dm ceil /dt is not negligible. This<br />

implies that the thickness of the first layer of a regional air quality model is<br />

important to the transport of dust.<br />

2. If the flux upwind of the road dm ambin /dt is not negligible compared to the flux<br />

from the road. If the flux upwind from a source is not negligible, a regional<br />

model must accommodate carrying the dust from upwind to downwind.<br />

3. If the approximation of a well mixed dust concentration from the ground to the<br />

top of the control volume is not a good description of the natural situation.<br />

4. If the assignment of the flux through the ceiling of the control volume as KC(x) is<br />

different from the diffusion taking place in the natural situation.<br />

5. If the assignment of the deposition as V d C(x) is different from the deposition<br />

taking place in the natural situation.<br />

6. If steady state conditions not observed.<br />

7. If the downwind flux dm ambout /dt is not of the order of dm ambin /dt within a few<br />

hundred meters to a kilometer, the result of the model would not be applicable to<br />

most regional scale regional models. Since the grid-scale of regional models is of<br />

the order of a kilometer, dm ambout /dt should be of the order of dm ambin /dt within a<br />

few hundred meters. The downwind flux equilibration distance may vary<br />

depending on local surface variables. This may have some implications for the<br />

minimum spatial resolution of a regional dust transport model. Perhaps in<br />

modeling, the distance could be varied as a function of the soil and vegetative<br />

cover—which in turn implies the need of good geographic data bases of such<br />

information.<br />

8. The analysis above was done for the wind perpendicular to the road. For wind at<br />

an angle to but not parallel the road, the analysis would be almost the same, but<br />

with a wider “effective road width.” For an infinitely long road parallel to the<br />

wind, however, the assumptions of the analysis would be violated and the solution<br />

would not be useful<br />

4.1.3.2a<br />

Weaknesses of the Box Model<br />

1. The model continues to pass dust mass through the CV ceiling even though realworld<br />

situations exist where no vertical flux takes place or there is vertical flux<br />

from above the ceiling into the CV.<br />

2. The model assumes dust to be well mixed in the control volume even though there<br />

are situations where dust continues to lie close to the ground (possibly caused by<br />

atmospheric stratification.)<br />

4.2 Inter-comparison of the ADE, ISC3, and the Gillette Box Model<br />

Since the model Gillette has proposed is mechanistically different from the ADE<br />

and the ISC3, we begin with a discussion of the assumptions in the Box Model.<br />

Modeling results from the ADE and the ISC3 are then compared with those obtained<br />

from the Box Model. Note that except were noted, all simulations for deposition apply to<br />

particles with aerodynamic diameters of 8 µm. PM 10 is the mass of all particles with<br />

4-10


diameters less than 10 microns. Discussing model results for all particle sizes requires<br />

more space in the text than is warranted. Eight microns is close to the Mass Mean<br />

Diameter (MMD) of dust particles in the PM 10 size range near the point of emission.<br />

Thus, the behavior of 8 micron particles, at least in the area close to the source is a<br />

reasonable surrogate for PM 10 as a whole.<br />

4.2.1 Analysis of the Gillette Box Model<br />

It is useful in analyzing the Gillette box model to consider the assumptions made<br />

in deriving that model. Invoking the resistance analogy, we consider the deposition and<br />

dispersion of particles contained in a box with very small vertical extent equal to ∆z and<br />

at an arbitrary height that falls somewhere in the surface layer. This is shown<br />

schematically in Figure 4-2. We can describe the rate of change of the concentration in<br />

the box as<br />

∂C<br />

=<br />

∂t<br />

( F + F )<br />

dep<br />

∆z<br />

Box−SL<br />

Equation 4-25<br />

where F Box-SL is the flux of particles out the top of the box and F dep is the deposition flux<br />

towards the ground. Replacing the fluxes with the concentration multiplied by the<br />

appropriate resistance and noting that the wind speed V is equal to dx/dt, we obtain<br />

<br />

<br />

C ra<br />

V<br />

<br />

=<br />

∂x<br />

1<br />

+ r + r r v<br />

<br />

+ v <br />

<br />

( C − C ) + − v ( C − C )<br />

∆z<br />

1<br />

<br />

ra<br />

∂<br />

g 0 Box<br />

1 b a1<br />

b g<br />

2<br />

g<br />

<br />

<br />

<br />

SL<br />

Box<br />

<br />

<br />

<br />

Equation 4-26<br />

where the variables in the equation have been defined in Figure 4-2. If we assume that the<br />

concentration at the ground C 0 is equal to zero (i.e. there is no particle rebound and the<br />

surface resistance is zero) and that the concentration at the top of the surface layer C SL is<br />

equal to zero then Equation 4-26 reduces to<br />

∂C<br />

− C<br />

V =<br />

∂x<br />

Box<br />

[ v + K ]<br />

d<br />

∆z<br />

G<br />

Equation 4-27<br />

with v d equal to the deposition velocity given in Equation 2-3 and the dispersion<br />

coefficient K G given by<br />

K<br />

G<br />

1<br />

=<br />

r<br />

a2<br />

− v<br />

g<br />

Equation 4-28<br />

4-11


where the subscript “G” is used to indicate that the dispersion coefficient is defined as in<br />

the Gillette box model with units of m/s. Equation 4-27 is identical to Equation 4-20, and<br />

forms the basis for the Gillette Box model. Note that the gravity resistance has been<br />

included for the transport of material through the top of the box model (i.e. Equation<br />

4-28).<br />

Planetary Boundary Layer (PBL)<br />

F = F(z)<br />

C SL Surface Layer2<br />

r a2<br />

F Box-SL =V da2 [C SL -C Box ]=(1/r a1 ) [C SL2 -C SL1 ]<br />

C Box ∆z<br />

r Surface Layer1<br />

a1 F dep = F QL = F Box-SL1 =V da2 [C Box -C QL ]=(1/r a2 ) [C Box -C QL ]<br />

C QL<br />

Quasi-laminar Layer<br />

F r dep = F Box-SL1 = F QL =V db [C QL -C 0 ]=(1/r b ) [C QL -C 0 ]<br />

b<br />

C 0 = 0<br />

Ground or Vegetative Surface<br />

r c r c assumed = 0 for particles if particle rebound is<br />

negligible<br />

r g<br />

Figure 4-2. Representation of the Gillette Box Model in the context of the resistance model to<br />

transport. The dashed horizontal line effectively divides the resistance r a into two smaller resistances,<br />

r a1 and r a2 . The equations for fluxes in the figure do not include the effect of gravity.<br />

In arriving at Equation 4-27 we have made three assumptions that warrant closer<br />

inspection. First, we have assumed that there is no resistance to deposition at the ground<br />

or vegetative surface (r c =0, C 0 =0). This assumption is frequently invoked for particles<br />

that are 10 µm in diameter or less, and we accept it as being reasonable (Seinfeld and<br />

Pandis, 1999; Raupach et al., 1999). Second, we have assumed that the concentration<br />

profile through the surface layer is sufficiently well-developed that F box-SL is constant<br />

from the box to the top of the surface layer. The validity of this assumption depends on<br />

the atmospheric stability. Under unstable conditions, most of the resistance to<br />

aerodynamic transport resides near the surface (see Figure 4-3). When conditions are<br />

stable, the resistance to transport does not drop off as dramatically with height.<br />

Therefore, assuming that the constant flux layer is instantly well-developed is probably a<br />

reasonable approximation for unstable conditions, less so for neutral conditions, and not<br />

valid for stable conditions.<br />

The magnitude of K G cannot be estimated for a stable atmosphere since r a does<br />

not converge on a maximum value. Under unstable conditions, r a converges on a<br />

maximum value, and we can in principle calculate K G . However, since r a is dependent on<br />

height above ground, we must make an assumption about which height to use. Selection<br />

4-12


of such a height is somewhat arbitrary. For example, in the formulation of the box model,<br />

it is suggested that K G = Au * with the value of A equal to 0.06. Neglecting the gravity<br />

term for the moment, we can use Equation 4-28 to infer the corresponding value for r a2<br />

1<br />

r a 2<br />

= =<br />

Au<br />

*<br />

1<br />

0.06u<br />

*<br />

Equation 4-29<br />

If we now use Equation 2-24 to infer the height z at which K G was calculated, we find<br />

that we need to specify the roughness height z 0 and the Monin-Obukhov length L. That is,<br />

our specification of K G (0.06u * ) is not only dependent on the height z, but is also<br />

dependent on the stability parameters of the flow. Thus, in general, our formulation for<br />

K G is very approximate and only reasonably applicable under neutral to unstable<br />

conditions.<br />

100<br />

aerodynamic resistance r a from the<br />

ground up to indicated height (s/m)<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

very unstable L=-2<br />

very unstable L=-10<br />

unstable L=-100<br />

slightly unstable L=-1,000<br />

slightly unstable L=-10,000<br />

neutral L=-100,000<br />

slightly stable L=1,000 stable L=100<br />

very stable L=10 very stable L=2<br />

0 2 4 6 8 10<br />

height above ground (m)<br />

Figure 4-3. The total aerodynamic resistance from the ground up to the height indicated on the x-<br />

axis. R a calculated after Byun and Dennis (1995). L is the Monin-Obukhov length and indicates<br />

stability. (100,000 > ⏐L⏐ > 10,000 near neutral, 10,000 > L > 10 stable, 10 > L > 0 very<br />

stable, 0 > L > -100 very unstable, -100 > L > -10,000 unstable). U * = 0.2 m/s and z 0 =0.01 m in<br />

all cases.<br />

Third, we have assumed that the concentration at the top of the surface layer (C SL ) is so<br />

small compared to the concentration in the box (C Box ) that it is effectively zero (i.e. C SL -<br />

C Box ≈ -C Box in Equation 4-27). Again, this approximation is only reasonably valid under<br />

unstable conditions when most of the resistance to transport within the surface layer is<br />

confined to the lowest few meters. Even then, the assumption only holds over a limited<br />

distance downwind. At some point, particles that were originally lofted to a high<br />

elevation are driven back down towards the ground. This occurs because particles are<br />

removed at the surface and the concentration gradient that originally (near the source)<br />

drove the vertical dispersion reverses direction as the concentration of particles at the<br />

surface decreases.<br />

4-13


4.2.2 Comparison of results for particle removal between predictions from ADE,<br />

ISC3, and the Gillette Box Model<br />

The results of numerically solving the Atmospheric Diffusion Equation are<br />

segregated by atmospheric stability and shown in Figure 4-4. The fraction of particles<br />

remaining in suspension increases as the atmospheric conditions vary from very stable to<br />

very unstable. For all atmospheric stability conditions, the fraction remaining in<br />

suspension decreases with increasing friction velocity. This is a very important point.<br />

Increasing the friction velocity not only enhances turbulent dispersion, but it also<br />

increases the impaction of particles and therefore the deposition velocity. Table 4-1 gives<br />

approximate values for the wind speed at 10 meters corresponding to the friction<br />

velocities shown in Figure 4-4. For a roughness height of 0.01 m and friction velocity of<br />

0.5 m/s, the equivalent 10 m wind speed is 13.8 m/s (30 mph), a value that corresponds to<br />

a fairly windy day at most locales. For low to moderate winds in arid to semi-arid areas<br />

(i.e. u * < 0.5 m/s), the fraction of particles remaining in suspension after 500 seconds is<br />

greater than 70% except for the very stable case. Note however that in the case of a<br />

stable atmosphere, the wind speed is unlikely to exceed a few meters per second, since at<br />

higher wind speeds, mechanical mixing at the surface would disallow the formation of a<br />

stable layer to any appreciable height (i.e. more than a few tenths of a meter).<br />

The trend of reduced deposition with increasing instability is consistent with<br />

expectation. During stable conditions, dust particles stay close to the ground for longer<br />

periods of time. Therefore, they are removed more quickly. In contrast, during unstable<br />

conditions, particles are lofted to higher elevations quickly by buoyant eddies. This<br />

means that those particles are not immediately available for deposition to ground<br />

surfaces. The concentration profiles are compared among different conditions of<br />

atmospheric stability in Figure 4-5. It is clear from the figure that unstable conditions<br />

result in the rapid vertical dispersion of particles.<br />

The removal of dust in the area near a source is often assumed to be significant<br />

because concentrations of particles downwind of a source rapidly attenuate to the<br />

background. This is an erroneous deduction. Watson and Chow (2000) and Countess<br />

(2001) both cite an earlier study (Watson et al., 1996) where the concentration of PM 10<br />

dust was found to decrease by 90% only 50 m downwind from the source. The<br />

concentration of particles may decrease rapidly with downwind distance, but it is<br />

incorrect to assume that the decrease is due solely to deposition. This concept is<br />

illustrated in Figure 4-6. The figure clearly shows that while concentrations decrease<br />

rapidly downwind of the source, the actual fraction of particles remaining in suspension<br />

may be quite high. For example, according to the ISC3 model, under neutral conditions,<br />

the concentration at a height of 1 meter 200 meters downwind of the road is 6% of its<br />

value near the road (within a few meters). However, 90% of the particles are still in<br />

suspension at the same downwind distance. Clearly, the decrease in concentration can be<br />

due primarily to the vertical mixing of particles and is not necessarily due to deposition.<br />

4-14


1<br />

0.9<br />

0.8<br />

1<br />

0.9<br />

0.8<br />

Fraction remaining<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3 ADE u*=0.1<br />

ADE u*=0.3<br />

0.2<br />

ADE u*=0.5<br />

0.1 ADE u*=0.8<br />

ADE u*=1.2<br />

0<br />

0 50 100 150 200 250 300 350 400 450 500<br />

Time since release (s)<br />

Fraction remaining<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3 ADE u*=0.1<br />

ADE u*=0.3<br />

0.2<br />

ADE u*=0.5<br />

0.1 ADE u*=0.8<br />

ADE u*=1.2<br />

0<br />

0 50 100 150 200 250 300 350 400 450 500<br />

Time since release (s)<br />

a. very stable (L=10 m) b. stable (L=1,000 m)<br />

1<br />

0.9<br />

0.8<br />

1<br />

0.9<br />

0.8<br />

Fraction remaining<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

ADE u*=0.1<br />

ADE u*=0.3<br />

ADE u*=0.5<br />

ADE u*=0.8<br />

ADE u*=1.2<br />

0 50 100 150 200 250 300 350 400 450 500<br />

Time since release (s)<br />

Fraction remaining<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

ADE u*=0.1<br />

ADE u*=0.3<br />

ADE u*=0.5<br />

ADE u*=0.8<br />

ADE u*=1.2<br />

0 50 100 150 200 250 300 350 400 450 500<br />

Time since release (s)<br />

c. neutral |L|>100,000 d. unstable (L=-1,000 m)<br />

Fraction remaining<br />

1<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

ADE u*=0.1<br />

ADE u*=0.3<br />

ADE u*=0.5<br />

ADE u*=0.8<br />

ADE u*=1.2<br />

0 50 100 150 200 250 300 350 400 450 500<br />

Time since release (s)<br />

e. very unstable (L=-10 m)<br />

Figure 4-4. Fraction of particles remaining in suspension vs. time since emission in the near field<br />

(0


200<br />

180<br />

160<br />

140<br />

Very Stable<br />

Stable<br />

Neutral<br />

Unstable<br />

Very Unstable<br />

Height z (m)<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

-0.1 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1<br />

Concentration Normalized to Initial Release<br />

200<br />

a. t = 10 seconds<br />

180<br />

160<br />

140<br />

Very Stable<br />

Stable<br />

Neutral<br />

Unstable<br />

Very Unstable<br />

Height z (m)<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

-0.1 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7<br />

Concentration Normalized to Initial Release<br />

200<br />

b. t=30 seconds<br />

180<br />

160<br />

140<br />

Very Stable<br />

Stable<br />

Neutral<br />

Unstable<br />

Very Unstable<br />

Height z (m)<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

-0.1 0.0 0.1 0.2 0.3 0.4 0.5<br />

Concentration Normalized to Initial Release<br />

C. t=100 seconds<br />

Figure 4-5. Concentration profiles according to the solution of the atmospheric diffusion equation at<br />

a. t=10 seconds , b. t=30 seconds, c. t=100 seconds. U * = 0.3 m/s in all cases.<br />

4-16


Fraction of Particles Remaining in Suspension or<br />

Normalized concentration<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

ADE-Normalized concentration<br />

ADE-Fraction in suspension<br />

ISC-Normalized concentration<br />

ISC-Fraction in suspension<br />

0 100 200 300 400 500 600 700 800 900 1000<br />

Distance Downwind (m)<br />

Figure 4-6. The fraction of particles remaining in suspension and the Concentration at a height of 1<br />

meter above ground level vs. time for neutral atmospheric conditions according to the ADE and ISC3<br />

models. z 0 =0.01 m, D p =8 µm, u * = 0.3 m/s in all cases.<br />

The fraction of particles removed is shown for multiple downwind distances in<br />

Figure 4-7 a-e. Results are shown for both the ISC method and the ADE. For the ADE,<br />

the distances are computed by multiplying the wind speed at 10 meters by the time in<br />

seconds (The one-dimensional ADE is only dependent on time). The wind speed at 10<br />

meters is in turn based on the assumption of a logarithmic wind profile with z 0 = 0.01 m.<br />

One interesting result that is common to both models is that when the friction<br />

velocity is 0.1 m/s, the removal of particles is greater than when it is 0.3 m/s; however, as<br />

u * continues to increase, the removal of particles also increases. The key to<br />

understanding this behavior is that the deposition velocity for particles also increases with<br />

friction velocity. For values of u * that are higher than a certain threshold (near 0.3 m/s<br />

for the conditions of Figure 4-7) the greater mixing that occurs is more than offset by the<br />

greater deposition rate, resulting in a net increase in particle removal. The opposite is<br />

true for values of u * that are less than the threshold.<br />

Both ADE and ISC3 models indicate that under neutral to unstable conditions, the<br />

fraction of particles remaining in suspension for values of u * ≤ 0.5 m/s is greater than<br />

78% at a downwind distance of 1 km and greater than 59% at a downwind distances of<br />

10 km. Even for a very high friction velocity (u * =1.2 m/s) the fraction of particles<br />

remaining in suspension 1 km downwind is greater than 50%. Note that in arid regions,<br />

very high values of u * initiate wind blown dust storms, dwarfing the emissions of dust<br />

from an unpaved road. For example, the unpaved road tests at Ft. Bliss were suspended<br />

on several occasions because of windblown dust events. These events occurred when u *<br />

was greater than 0.6 m/s, approximately equivalent to a wind speed of 14 m/s at a height<br />

of 10 m.<br />

4-17


Fraction Remaining<br />

Fraction Remaining<br />

1<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

1<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

ISC<br />

ISC<br />

ISC<br />

ISC<br />

ISC<br />

ISC<br />

ISC<br />

ISC<br />

ISC<br />

ISC<br />

ADE, u*=0.1<br />

ADE, u*=0.3<br />

ADE, u*=0.5<br />

ADE, u*=0.8<br />

ADE, u*=1.2<br />

100 500 1000 5000 10000<br />

Distance Downwind(m)<br />

Fraction Remaining<br />

1<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

ISC<br />

ISC<br />

ISC<br />

ISC<br />

ISC<br />

ADE, u*=0.1<br />

ADE, u*=0.3<br />

ADE, u*=0.5<br />

ADE, u*=0.8<br />

ADE, u*=1.2<br />

100 500 1000 5000 10000<br />

Distance Downwind(m)<br />

a. very stable (L=10 m) b. stable (L=1,000 m)<br />

ADE, u*=0.1<br />

ADE, u*=0.3<br />

ADE, u*=0.5<br />

ADE, u*=0.8<br />

ADE, u*=1.2<br />

100 500 1000 5000 10000<br />

Distance Downwind(m)<br />

Fraction Remaining<br />

1<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

ISC<br />

ISC<br />

ISC<br />

ISC<br />

ISC<br />

ADE, u*=0.1<br />

ADE, u*=0.3<br />

ADE, u*=0.5<br />

ADE, u*=0.8<br />

ADE, u*=1.2<br />

100 500 1000 5000 1000<br />

Distance Downwind(m)<br />

c. neutral |L|>100,000 d. unstable (L=-1,000 m)<br />

Fraction Remaining<br />

1<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

ISC ADE, u*=0.1<br />

ISC ADE, u*=0.3<br />

ISC ADE, u*=0.5<br />

ISC ADE, u*=0.8<br />

ISC ADE, u*=1.2<br />

100 500 100 500 100<br />

Distance Downwind(m)<br />

e. very unstable (L=-10 m)<br />

Fraction Remaining<br />

1<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

u*=0.1<br />

u*=0.3<br />

u*=0.5<br />

u*=0.8<br />

u*=1.2<br />

A=0.04 A=0.08 A=0.12 A=0.2 A=0.3<br />

Dispersion Parameter in Box Model<br />

f. Box model for various values of the<br />

parameter A G and V d (1 m)<br />

Figure 4-7. Fraction of particles remaining in suspension for a given downwind distance. Panels a<br />

through e correspond to numerical solutions to the atmospheric diffusion equation with z 0 =0.01 m,<br />

D p =8 µm. under various conditions of atmospheric stability. The downwind distance is estimated<br />

based on the wind speed at 10 m assuming a logarithmic profile and z 0 = 0.01 m. Panel f shows the<br />

fraction remaining according to the box model for various assumed values of A G .<br />

The degree of particle removal is greater under stable conditions than neutral and<br />

unstable conditions. This is intuitive, since in the absence of buoyant mixing, particles<br />

are likely to remain closer to the ground where they can be removed by impaction and<br />

4-18


gravitational settling. Stable conditions cannot exist when the friction velocity is high<br />

because mechanical mixing does not allow for stratification to occur to any appreciable<br />

extent (there is always a small stable layer near the ground at night, though it may only be<br />

a few centimeters in depth when wind speeds are high). Therefore, conservatively<br />

assuming that u * does not exceed 0.5 m/s when conditions are stable, the removal of<br />

particles 1 km downwind of the unpaved road is not likely to be greater than about 50%.<br />

There is an exception to this result that occurs when the friction velocity is nearly zero<br />

(i.e. very little vertical mixing and downwind transport) and particles settle to the ground<br />

under the influence of gravity. In this case, the dispersion models presented here are not<br />

applicable, and the fraction of particles remaining in suspension decreases linearly with<br />

time since emission (e.g. Figure 4-8). It seems unlikely that there would be a significant<br />

amount of motor vehicle traffic on unpaved roads during nighttime stable conditions.<br />

Thus, though removal of particles under stable conditions is considered for completeness,<br />

we note that for most emissions from unpaved roads, conditions are probably neutral to<br />

unstable.<br />

1.2<br />

Fraction remaining in suspension<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0 200 400 600 800 1000 1200<br />

Time since Release (s)<br />

Figure 4-8. Fraction of particles remaining in suspension vs. time since emission for 8 µm particles<br />

falling under the influence of gravity in the absence of any dispersion (u * =0).<br />

Comparing the ADE and ISC model predictions, the predictions of fraction<br />

remaining in suspension are very similar, though some systematic differences exist. Very<br />

near the source (first few hundred meters), the assumptions of the Atmospheric Diffusion<br />

Equation are violated, because vertical gradients in concentration are large compared to<br />

the absolute values of concentration. This results in a significant over prediction of the<br />

initial dispersion of the plume. In the ISC model, the vertical extent of the plume is<br />

specified as a function of downwind distance (through the parameter σ z ) and is therefore<br />

not dependent on the steepness of the concentration gradient. The difference in initial<br />

downwind dispersion between the ADE and the ISC is examined further in the next<br />

section. Another reason for the difference between ADE and ISC predictions is that the<br />

ADE accounts for dispersion explicitly through the friction velocity u * . That is, the<br />

4-19


downwind concentration profile is dependent on u * . In contrast, the ISC model<br />

prescribes the concentration profile based on downwind distance only. In practice, this<br />

means that the ISC does not allow for alteration of the concentration profile due to<br />

variations in u * . This point is illustrated in the next chapter.<br />

Figure 4-7 f. shows the fraction remaining in suspension according to the box<br />

model for various values of the parameter A in Equation 4-29. The deposition velocity<br />

for the box model was computed assuming a height of 1 meter and neutral stability. Note<br />

that the box model gives a result that is independent of distance downwind of the source.<br />

For all values of A, the box model captures the general behavior of particle deposition to<br />

a surface; the greater the dispersion rate (as characterized by A and u * ), the smaller the<br />

fraction of particles that deposit. However, comparison of Figure 4-7f. with the other<br />

five panels in the figure (ADE and ISC3 model results) underscores the basic difficulty of<br />

using the box model: Though in its derivation the box model does not depend on the<br />

height of the box or the downwind distance, the parameter A used in the model depends<br />

on both. A also depends on atmospheric stability.<br />

It is clear from Figure 4-5 that for a line source, the concentration profile through<br />

the surface layer is dynamic, becoming flatter in the vertical direction with downwind<br />

distance. This means that the concentration at the top of the surface layer (C SL in the<br />

previous section) is neither constant in time, nor is it negligibly small compared to the<br />

concentration in the box (C Box ). Thus, the formulation for dispersion in the Box Model is<br />

not valid for a plume that is expanding in the vertical direction with concurrent decay in<br />

the concentration gradients within the surface layer.<br />

In a dust storm (Gillette, 1974), the ground acts like an area source, which<br />

behaves quite differently in the context of the Box Model from a line source for three<br />

reasons. First, an area source is horizontally homogeneous. Therefore, the vertical<br />

concentration profile is, on average, invariant with downwind distance. Second, the<br />

ground is acting as a strong source and the net flux of material is always in the upward<br />

direction. This means that we do not have to account for particles re-entering the Box<br />

through the ceiling. Third, because in a dust storm, the ground is a very strong source,<br />

the concentration near the ground is much higher than it is aloft (i.e. C Box >> C SL ), or at<br />

least, the ratio C Box /C SL is constant over large distances in the horizontal. For these three<br />

reasons, dispersion in a dust storm can be estimated based only on the concentration near<br />

the ground (as in Equation 4-27), without considering the concentration profile aloft. For<br />

a dust plume that is expanding while traveling downwind, the assumptions behind<br />

Equation 4-27 do not hold as well.<br />

4.3 Summary<br />

For unpaved road dust emissions, the Box Model provides an order of magnitude<br />

estimate of the dust particle removal due to deposition. For a more accurate assessment, a<br />

model that accounts for changes in the concentration profile with downwind distance is<br />

required. Among the models considered in this section, the algorithm of the ISC3 is most<br />

suited for simulating near-source dispersion. The ADE performs poorly in the near field,<br />

4-20


though it is expected to improve at larger downwind distances (500 - 1,000 m). In the<br />

next two chapters, the ISC3 and ADE models are compared with field data.<br />

4-21


4-22


5. DETERMINATION <strong>OF</strong> PARTICLE REMOVAL RATES,<br />

EMISSION FACTORS, <strong>AND</strong> INJECTION HEIGHTS ON AN<br />

UNPAVED ROAD IN AN ARID SETTING<br />

As part of an ongoing DoD-funded contract, leveraged by this WESTAR project,<br />

field experiments were performed at the Ft. Bliss military facility near El Paso TX.<br />

Towers instrumented with particle monitors were used to measure the vertical<br />

concentration profiles and horizontal flux of PM 10 dust particles emitted from an unpaved<br />

road. These data provided both an estimate of emission factors for a number of vehicles<br />

and an estimate of the fraction of PM 10 emissions that is regionally transportable. During<br />

the same field campaign, the turbulence behind moving vehicles was examined because<br />

of a hypothesis that the spatial extent of the wake may have a strong influence on the<br />

transportable fraction of emitted dust particles. In this Chapter, the results of the Ft. Bliss<br />

field tests are presented. Those results are compared with predictions from both the ADE<br />

and the ISC models.<br />

5.1 Methods<br />

5.1.1 Upwind/Downwind Towers<br />

From April 11 through 24, 2002 unpaved road emissions experiments were<br />

conducted at Ft. Bliss. The procedure was based on an upwind/downwind technique that<br />

has been used by other investigators (e.g., Gillies et al., 1999; Cowherd, 1999). Three<br />

towers were set up collinearly and perpendicular to a 1,000 m section of unpaved road on<br />

the open range of Ft. Bliss, TX, which was oriented in a north-south direction. Historical<br />

meteorological data indicated that winds at this time of year in this area were<br />

predominantly from the west.<br />

The three towers were all downwind of the road at distances of 7 m, 50 m, and<br />

100 m. The tower closest to the road was 6 m high with the other two being 12 m in<br />

height(Figure 5-1). Each downwind tower was instrumented with four DustTraks (Model<br />

8520, TSI Inc., St. Paul, MN) that were spaced logarithmically in the vertical direction.<br />

The DustTrak is a portable, battery-operated, laser-photometer that uses light scattering<br />

technology to determine mass concentration in real-time. In applied studies of dust<br />

emissions from playa surfaces and partially vegetated surfaces Houser and Nickling<br />

(2001), and Nickling et al. (1997), respectively, found these instruments to be superior to<br />

the traditional method of collecting suspended sediment on filters. Etyemezian et al.<br />

(submitted) and Kuhns et al. (2001; submitted) used DustTraks to measure PM 10<br />

emission from paved and unpaved roadways using an on-board vehicle measurement<br />

system.<br />

The tower at 7 m downwind (DT_1) had measurement positions at 0.76, 1.28,<br />

2.66, and 5.18 m above ground level (AGL). The second tower (DT_2) at 50 m had<br />

measurement positions at 1.25, 2.6, 5.7, and 12.2 m. The third tower at 100 m (DT_3)<br />

had the same measurement positions as DT_2 with an additional sampling location at 0.4<br />

m. The DustTraks were equipped with PM 10 inlets and measured particle concentrations<br />

at intervals of 1 second. Four anemometers, one wind vane, and one temperature probe<br />

were mounted on the DT_3 tower in order to characterize the local meteorological<br />

conditions. During the course of the study four GRIMM model 1.108 particle size<br />

5-1


analyzers (GRIMM Tech. Inc., Atlanta, GA) were deployed at various locations and<br />

different measurement periods to provide information on the particle size characteristics<br />

of the emitted dust.<br />

-DT<br />

-SA<br />

-DT<br />

Top View<br />

Legend<br />

DT_3<br />

DT_2<br />

DT_1<br />

1220<br />

DT: DustTrak<br />

GR: GRIMM<br />

SA: Sonic Anemometer<br />

: Wind Vane<br />

: Cup Anemometer<br />

: Laptop computer<br />

1220<br />

Trailer with<br />

visibility<br />

equipment<br />

Generator<br />

200 meters<br />

LIDAR 3,000<br />

meters<br />

-DT<br />

-GR<br />

-DT<br />

-GR<br />

-DT<br />

570<br />

570<br />

517<br />

-DT<br />

-GR<br />

-DT<br />

-DT-GR<br />

260<br />

260<br />

125<br />

40<br />

-DT<br />

-DT<br />

-GR<br />

125<br />

-DT<br />

-GR<br />

266 128<br />

76<br />

-DT<br />

-DT<br />

10,000<br />

5,000<br />

700<br />

Figure 5-1. Schematic of equipment setup at Ft. Bliss during the April, 2002 field campaign. The<br />

vertical and horizontal components are drawn on different scales.<br />

The concentration data from the DustTraks, particle size distribution data from the<br />

GRIMMs, and meteorological data were collected in 1-second intervals on a PC located<br />

at the base of each tower running a custom-designed LabView (National Instruments)<br />

data acquisition program. The PC clocks were set to a master clock at the beginning of<br />

each day to ensure the data collected from the towers could be merged and synchronized<br />

into a master database. The data files were created in Microsoft Access readable format.<br />

At the beginning of each day, the zero baselines were set for each DustTrak on the<br />

flux towers with a HEPA filter. Analysis of the DustTrak data indicated that baseline<br />

drift over the course of a day, generally less than 50 µg/m 3 , did not affect all instruments<br />

equally. Thus, a baseline for each instrument was calculated approximately every 20<br />

minutes over the course of the day from background concentrations measured during<br />

periods when there were no dust emissions from the unpaved road. PM 10 concentrations<br />

measured during periods when the tower was influenced by a passing dust plume were<br />

corrected with the most recent baseline value that was determined. This ensured that the<br />

resulting concentration was due only to road dust emissions generated by passing<br />

vehicles.<br />

A portable filter sampler with a PM 10 inlet (Aerometrics, MiniVol) was collocated<br />

with a DustTrak at 1.5 m on DT_1. The filter sampler was operated for 6 hours during<br />

testing. The Concentration of PM 10 measured with on the filter was compared to the<br />

average concentration measured by the DustTrak as a means of calibrating the DustTrak<br />

light scattering measurement with a mass measurement. The results indicated that the<br />

5-2


PM 10 measured by the DustTrak must be multiplied by 1.35 ± 0.15 in order to obtain an<br />

equivalent PM 10 mass-based measurement. The DustTrak-based emission factors<br />

presented in this report have been corrected to reflect an equivalent mass-based<br />

measurement (i.e. the DustTrak signal has been multiplied by 1.35).<br />

PM 10 emissions fluxes were calculated for each downwind tower using the<br />

assumption that each DustTrak represented the PM 10 concentration over a height that<br />

spanned half the distance to the next lowest monitor to half the distance to the next<br />

highest monitor. For example, the lowest monitor at 0.76 m AGL on DT_1 was<br />

considered representative of the PM 10 concentration from the ground surface to 1.02 m<br />

AGL. The second monitor at 1.28 m AGL represented the PM 10 concentrations from<br />

1.02 m agl to 1.94 m agl, and so on. The time series of PM 10 concentrations were<br />

examined for each DustTrak at each location. Each peak in concentration was associated<br />

with an individual vehicle pass and each pass was visually assigned a start and stop time.<br />

Figure 5-2 shows an example of a concentration time series and the determination of peak<br />

start and stop times. The emissions factor per vehicle pass for each downwind tower was<br />

calculated using the sum of the 1-second PM 10 fluxes with the equation:<br />

EF =<br />

endofpeak<br />

<br />

startofpeak<br />

<br />

cos(θ<br />

)<br />

<br />

4<br />

<br />

i=<br />

1<br />

u C ∆z<br />

∆t<br />

i<br />

i<br />

i<br />

<br />

<br />

<br />

Equation 5-1<br />

where EF is the emissions factor of PM 10 in grams per vehicle kilometer traveled, θ is the<br />

angle (1-second measurement) between the wind direction and a line perpendicular to the<br />

road, i is one of the four positions (five for T_3) of the monitors on the tower, u i is the<br />

average wind speed in m/s over the interval represented by the i th monitor (1-second<br />

measurement), C i is the baseline-corrected PM 10 concentration in mg/m 3 as measured by<br />

the i th monitor over the period ∆t (1-second measurement), ∆z in m is the vertical interval<br />

represented by the i th monitor, ∆t is equal to 1 second. Note that if multiplied by the<br />

length of the unpaved road, the bracketed term following the first summation symbol<br />

would be the amount of PM 10 that passes through a vertical plane in an interval of time<br />

equal to ∆t (1 second).<br />

The friction velocity, roughness height, and displacement height (u * , z 0 , and d,<br />

respectively) where estimated from the vertical distribution of wind speeds. The wind<br />

profile was assumed to be logarithmic as in Equation 2-9 with the displacement height, d,<br />

equal to zero. This is a reasonable assumption considering the sparseness of vegetation at<br />

Ft. Bliss and that d is expected to be much less than the height of the first anemometer<br />

(1.25 m), making it difficult to accurately determine (since z-d≈ z). In order to obtain<br />

meaningful estimates of the bulk quantities u * and z 0 , data were averaged over 15-minute<br />

intervals to obtain the vertical distribution of wind speeds. There are 110 15-minute<br />

intervals that corresponded to times during which sampling was in progress. For each 15-<br />

minute interval and at each height where the wind speed was measured, the friction<br />

velocity, u * was calculated from an initial guess for z 0 and the measured values of u(z),<br />

yielding four calculated values of u * . These four values were averaged and the average<br />

friction velocity was used to obtain a velocity according to Equation 2-9 for each of the<br />

5-3


four heights. The root mean sum of squares (RMS) was then computed from the relative<br />

difference between the calculated velocities and the measured velocities. The sum of the<br />

RMS over all 110 intervals was used as a weighting function. The final estimate for z 0<br />

was obtained by minimizing the weighting function. Note that this method is superior to<br />

performing a regression for each 15-minute interval, since it provides a single value for z 0<br />

that is applicable to all intervals.<br />

20<br />

15<br />

DustTrak Reading (mg/m3)<br />

10<br />

5<br />

Vehicle passes by DT_1<br />

0<br />

Baseline<br />

DT_1 DT_2<br />

DT_3<br />

-5<br />

18:02:10 18:02:27 18:02:44 18:03:01 18:03:19 18:03:36 18:03:53 18:04:11<br />

Time<br />

Figure 5-2. Example of time series of DustTrak PM 10 concentrations after the passage of a vehicle on<br />

the unpaved road at Ft. Bliss. The arrows in the figure illustrate the start and stop times estimated<br />

for a baseline reading and the dust plume passing through the downwind towers DT_1, DT_2, and<br />

DT_3.<br />

The final value of 0.005 m obtained for z 0 lies between those expected for “level<br />

desert” (0.001 m) and “lawn” (0.01 m) (Seinfeld and Pandis, 1999). Figure 5-3 shows the<br />

distribution of u * values over the 111 intervals. Ninety-five percent of the values of u *<br />

fell between 10 and 50 cm/s, with values between 30 and 40 cm/s occurring most<br />

frequently (35%). Examples of how the wind speed fitted to a logarithmic vertical profile<br />

compares with the measured wind speed are given in Figure 5-4 for conditions<br />

corresponding to the lightest, most frequent, and heaviest winds observed over the 110<br />

15-minute intervals. In general, the logarithmic fit is reasonable for all cases, though the<br />

wind speed at 12.2 meters is overestimated in the case of light winds. This may be caused<br />

by small differences in the resistance to motion among the four cup anemometers used.<br />

The fact that in Figure 5-4a the measured wind speed at 12.2 meters is slightly less than<br />

that at 5.7 meters supports this hypothesis.<br />

5.1.1.1 Emission Tests<br />

Road dust emissions were created by having a test vehicle travel back and forth<br />

along the roadway for a number of passes. This ranged from a minimum of 20 to a<br />

maximum of 102. The mean number of vehicle passes was 49. The test vehicles traveled<br />

at set speeds of 16, 24, 32, 40, 48, 56, 64, 72, and 81 km/hr. Not all vehicles attained the<br />

highest speeds for safety considerations. After two passes (one heading south and one<br />

returning north) at the same speed, the vehicle speed was increased incrementally to the<br />

5-4


100%<br />

90%<br />

80%<br />

Relative Frequency (probability)<br />

Cumulative %<br />

Relative Frequency<br />

70%<br />

60%<br />

50%<br />

40%<br />

30%<br />

20%<br />

10%<br />

0%<br />

0 -10 10 - 20 20 - 30 30 - 40 40 - 50 50 - 60 > 60<br />

u* (cm/s)<br />

Figure 5-3. Frequency distribution and cumulative distribution of u * for the 110 15-minute intervals<br />

corresponding to periods when testing was ongoing at Ft. Bliss. The u * values are based on a<br />

roughness height, z 0 equal to 0.005 m.<br />

u*= 10 cm/s<br />

u*= 32 cm/s<br />

u*= 54 cm/s<br />

14<br />

u(z) measured<br />

u(z) fitted<br />

14<br />

u(z) measured<br />

u(z) fitted<br />

14<br />

u(z) measured<br />

u(z) fitted<br />

12<br />

12<br />

12<br />

Height (z) in meters<br />

10<br />

8<br />

6<br />

4<br />

Height (z) in meters<br />

10<br />

8<br />

6<br />

4<br />

Height (z) in meters<br />

10<br />

8<br />

6<br />

4<br />

2<br />

2<br />

2<br />

0<br />

12 16 20 24<br />

u(z)/u *<br />

0<br />

12 16 20 24<br />

u(z)/u*<br />

a. b. c.<br />

0<br />

12 16 20 24<br />

u(z)/u*<br />

Figure 5-4. Examples of comparison between measured wind speed and wind speed calculated from<br />

curve fitting of Equation 2-9. a. light winds (u * =10 cm/s), b. medium winds (u * = 32 cm/s), c. high<br />

winds (u * = 54 cm/s). The roughness height, z 0 is equal to 0.005 m in all cases.<br />

5-5


highest attainable value and then decreased incrementally to the minimum value. This<br />

cycle was then repeated as time allowed. Test vehicles paused for approximately one and<br />

a half minutes between passes to allow for the dust plume to clear all three downwind<br />

towers. The schedule and test conditions are summarized in Table 5-1.<br />

Table 5-1. Test dates, times, vehicles, data recovery, and conditions during the FT. Bliss, April 2002<br />

field campaign.<br />

Date<br />

4/11/2002<br />

4/11/2002<br />

4/12/2002<br />

4/12/2002<br />

4/15/2002<br />

4/15/2002<br />

4/16/2002<br />

4/16/2002<br />

4/16/2002<br />

4/17/2002<br />

4/18/2002<br />

4/19/2002<br />

4/22/2002<br />

4/23/2002<br />

4/24/2002<br />

4/24/2002<br />

Time<br />

13:44 -<br />

14:51<br />

15:11 -<br />

16:06<br />

14:41 -<br />

15:21<br />

15:35 -<br />

16:13<br />

10:38 -<br />

12:37<br />

12:48 -<br />

13:30<br />

10:57 -<br />

12:26<br />

13:53 -<br />

15:35<br />

17:18 -<br />

19:20<br />

10:22 -<br />

12:36<br />

10:56 -<br />

13:15<br />

11:06 -<br />

14:29<br />

9:47 -<br />

10:08<br />

13:38 -<br />

15:11<br />

10:50 -<br />

12:33<br />

13:41 -<br />

15:16<br />

Equipment<br />

status<br />

DT3 only, No<br />

wind<br />

direction data<br />

DT3 only, No<br />

wind<br />

direction data<br />

DT3 only<br />

DT3 only<br />

DT1, DT2,<br />

DT3<br />

DT1-partial,<br />

DT2, DT3<br />

DT1, DT2,<br />

DT3<br />

DT2, DT3<br />

DT1, DT2,<br />

DT3<br />

DT1, DT2,<br />

DT3<br />

DT1, DT2,<br />

DT3<br />

DT1-partial,<br />

DT2, DT3<br />

DT1, DT2,<br />

DT3<br />

DT1-partial,<br />

DT2, DT3<br />

DT1, DT2,<br />

DT3<br />

DT1, DT2,<br />

DT3<br />

Vehicle<br />

HUMVEE<br />

Dodge<br />

Caravan<br />

Dodge<br />

Caravan<br />

Dodge<br />

Neon<br />

Ford Taurus<br />

U-Haul<br />

Dodge<br />

Caravan<br />

5-ton truck<br />

1979 Chevy<br />

van<br />

(TRAKER)<br />

LMTV<br />

Hemmet<br />

22<br />

Freightliner<br />

Dodge<br />

Neon<br />

22<br />

Freightliner<br />

1979 Chevy<br />

van<br />

(TRAKER)<br />

Vehicle<br />

speeds<br />

(km/hr)<br />

8, 16, 24, 32,<br />

40, 48, 56,<br />

64, 72<br />

16, 32, 48,<br />

64, 81<br />

Total<br />

number of<br />

valid passes<br />

26<br />

Wind conditions<br />

Light and variable with small<br />

westerly componenet<br />

40 South -West<br />

48 30 Light and variable<br />

16, 24, 32,<br />

48, 64, 81<br />

16, 32, 48,<br />

64, 81<br />

16, 32, 48,<br />

64<br />

16, 32, 48,<br />

64, 81<br />

16, 32, 48<br />

,64<br />

16, 24, 32,<br />

40, 48, 64<br />

16, 32, 48,<br />

64, 81<br />

16, 32, 48,<br />

64<br />

16, 24, 32,<br />

40, 48, 56<br />

16, 32, 48,<br />

64<br />

16, 24, 32,<br />

40, 48, 56<br />

16, 24, 32,<br />

40, 48, 56<br />

HUMVEE 16, 64 2<br />

12 Light and variable<br />

23 South -West<br />

11 South -West<br />

29 South -West<br />

27 South -West<br />

46 South -West<br />

16 South -West<br />

16 South -West, with gusts<br />

11 South -West, with gusts<br />

0 South: Test suspended<br />

18 West, with gusts<br />

0<br />

Light winds from the North.<br />

No valid data obtained<br />

Winds light and variable. Data<br />

recovery poor.<br />

The unpaved road at Ft. Bliss was 1 kilometer in length and oriented in the northsouth<br />

direction, with the three towers lined perpendicular to and east of the road. Thus,<br />

when winds did not have a strong westerly component, the dust plume generated by<br />

passing vehicles did not travel in the direction of the towers. The transport time between<br />

the road and the far downwind tower (100 m) was estimated to be approximately 30<br />

seconds. One-second wind direction data were averaged for the 30 seconds prior to plume<br />

impact at the far downwind tower. If the average direction fell in the quadrant spanned<br />

by the northwest and southwest vectors, the corresponding vehicle pass was considered<br />

valid. Vehicle passes corresponding to wind directions that were not in the northwestsouthwest<br />

quadrant were considered suspect and not included in subsequent data analysis.<br />

Note that for westerly winds, a 30-second transport time corresponds to a wind speed of<br />

3.3 m/s while for southwesterly or northwesterly winds, it corresponds to 4.7 m/s. The<br />

5-6


average wind speed measured at 12 meters over all emissions tests was 6.2 m/s. Thus, a<br />

transport time of 30 seconds is probably, on average, a conservative overestimate.<br />

The wind speed and direction were measured only at the far downwind tower<br />

(DT_3). Therefore, for using Equation 5-1, it was necessary to use wind speeds and<br />

directions measured at DT_3 for DT_1 and DT_2. The possibility of introducing errors<br />

by this method was examined. The horizontal PM 10 fluxes were calculated for DT_3 only<br />

and averaged over the test vehicle speed. This procedure was performed twice, once with<br />

wind data that corresponded to the time of the passing of the dust plume, and once with<br />

wind data that was retarded by 30 seconds. For example, the flux at 13:30:00 was<br />

calculated using wind data measured at 13:30:00 and also with wind data measured at<br />

13:29:30. The purpose of this exercise was to subject the flux calculations for DT_3 to<br />

the same uncertainties that the data from DT_1 and DT_2 would be experiencing. In<br />

comparing the two resultant sets of horizontal fluxes, no significant difference were<br />

found (slope = 1.01, R 2 =0.98, n=53, intercept was forced to 0), indicating that the use of<br />

DT_3 wind direction and wind speed to calculate horizontal fluxes of PM 10 at DT_1 and<br />

DT_2 would not introduce errors.<br />

5.1.2 Sonic Anemometer Tests<br />

A three-dimensional sonic anemometer (“A” style probe, applied Technologies<br />

Inc) was used to assess the effect of vehicle size on the initial distribution of the dust<br />

plume generated. The anemometer was first collocated with the cup anemometer and the<br />

wind vane at the 12.2 m height on DT_3 for three sample days (4/18/02 – 4/20/02). The<br />

sonic anemometer was set to measure the U, V, and W components (corresponding to the<br />

x, y, and z directions) of the velocity at a frequency of 10 Hz. Figure 5-5 shows the<br />

comparison between wind speeds measured with the cup anemometer and a sonic<br />

anemometer; part a. shows a time series; parts b. through d. show comparisons of 1-<br />

second averages and 1-minute averages. In general, the cup and the sonic anemometers<br />

track each other well, though the sonic anemometer gives smaller values of wind speed as<br />

indicated by the slopes of the regressions in Figure 5-5 b-d (0.91 – 0.94). The significant<br />

improvement in R 2 values between the 1-second (0.84-0.89) and 1-minute averaged data<br />

(≥ 0.98) is to be expected since the cup anemometer has a nominal response time of 1<br />

second (approximately the time required to register 1/3 of the difference between changes<br />

in wind speed) while the sonic anemometer response is practically instantaneous.<br />

On 4/21/02 between 19:44 and 22:15 a series of tests were conducted to assess the<br />

magnitude of the turbulent wake behind a passing vehicle. The tests occurred at night<br />

because under stable nighttime atmospheric conditions, the turbulence generated by a<br />

vehicle can be more readily identified with a sonic anemometer than when the<br />

atmosphere is unstable and the background turbulence level is much higher. The<br />

downwind tower closest to the unpaved road was moved to the edge of the road. Two<br />

markers were placed on the unpaved road, one at a distance of 2 meters from the edge of<br />

the road (and the tower) and the other at a distance of 6 meters. The sonic anemometer<br />

was mounted on the tower at a height of 0.9 meters above ground level (AGL) and<br />

protruded 1.1 meters towards the road (i.e. approximately 1 meter from the near marker).<br />

Three vehicles with considerably different profiles were tested: a Dodge Neon, a 1979<br />

Chevy cargo van, and a 24-foot moving truck. Each vehicle was driven through the two<br />

markers on the road once heading north and once heading south at 16, 32, 48, and 64<br />

5-7


km/hr for a total of eight passes per vehicle. The time, accurate to within two seconds,<br />

was recorded whenever the vehicle passed immediately in front of the sonic anemometer.<br />

After all three vehicles completed this cycle, the sonic anemometer was moved to 2.4<br />

meters and then to 5.5 meters AGL. The driving pattern (three vehicles, four speeds, two<br />

passes at each speed) was repeated at each of those two heights.<br />

Under stable atmospheric conditions, natural fluctuations in the vertical<br />

component of velocity (W) are expected to be small. Thus, W was chosen to assess the<br />

vertical extent of the turbulent wake generated by the passage of the vehicles. This was<br />

done by making use of Reynold’s decomposition<br />

W ( t)<br />

= Wave + W '<br />

Equation 5-2<br />

where W(t) is the instantaneous velocity at time t, W ave is the average component of the<br />

vertical velocity, and W’ is the fluctuating component of the velocity. Upon inspection of<br />

the data, it became apparent that the effect of the passing vehicle could be seen from 7<br />

seconds prior to the recorded vehicle pass time to 15 seconds after the recorded time.<br />

This was not the case for every vehicle pass, but this “influence” interval (-7 to +15<br />

seconds) was large enough to be inclusive of all vehicle passes. W ave in Equation 5-2 was<br />

obtained by averaging the W-component of the velocity for 30 seconds prior to the<br />

beginning of the interval influenced by the vehicle (i.e. –37 to –7 seconds) and for 30<br />

seconds following the interval (+15 to +45 seconds). These two 30-second intervals were<br />

considered “background” conditions. The background turbulence was quantified by<br />

taking the average and standard deviation of J b over each 0.1 second measurement<br />

obtained during the entire “background” period where<br />

J<br />

b<br />

=<br />

( W t)<br />

−W<br />

) 2<br />

(<br />

b ave<br />

1<br />

Equation 5-3<br />

and the subscript b indicates that this operation is performed for the background intervals.<br />

The turbulence generated by the passing of the vehicle is likewise quantified by obtaining<br />

the average of J i<br />

J<br />

i<br />

=<br />

( W t)<br />

−W<br />

) 2<br />

(<br />

i ave<br />

Equation 5-4<br />

1 As an aside, note that J can be thought of as the vertical component of the turbulence<br />

kinetic energy (KE) which is customarily defined as<br />

2 2<br />

KE = U ' + V ' +<br />

W<br />

2<br />

'<br />

where U and V represent the x- and y- component velocities.<br />

5-8


where the subscript i indicates that this operation is performed for the influence interval (-<br />

7 to + 15 seconds). Figure 5-6 shows an example of the data resulting from the passage of<br />

a vehicle and the corresponding background and influence intervals.<br />

12<br />

10<br />

Wind speed (m/s)<br />

8<br />

6<br />

4<br />

2<br />

Sonic anemometer<br />

Cup anemometer<br />

0<br />

9:49:41 9:50:24 9:51:07 9:51:50 9:52:34 9:53:17<br />

Time of Day<br />

a.<br />

1-Second Wind Speed Measured<br />

with Sonic Anemometer (m/s)<br />

1-Second Wind Speed Measured<br />

with Sonic Anemometer (m/s)<br />

1-Second Wind Speed Measured<br />

with Sonic Anemometer (m/s)<br />

20<br />

15<br />

10<br />

5<br />

y = 0.93x<br />

R 2 = 0.89<br />

4/18/02<br />

0<br />

0 5 10 15 20<br />

20<br />

15<br />

10<br />

5<br />

1-Second Wind Speed Measured with Cup Anemometer (m/s)<br />

y = 0.94x<br />

R 2 = 0.84<br />

4/19/02<br />

0<br />

0 5 10 15 20<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

1-Second Wind Speed Measured with Cup Anemometer (m/s)<br />

y = 0.91x<br />

R 2 = 0.89<br />

4/20/02<br />

0<br />

0 5 10 15 20 25 30<br />

1-Second Wind Speed Measured with Cup Anemometer (m/s)<br />

b.<br />

c.<br />

d.<br />

1-Minute Wind Speed Measured<br />

with Sonic Anemometer (m/s)<br />

1-Minute Wind Speed Measured<br />

with Sonic Anemometer (m/s)<br />

1-Minute Wind Speed Measured<br />

with Sonic Anemometer (m/s)<br />

20<br />

15<br />

10<br />

5<br />

y = 0.94x<br />

R 2 = 0.99<br />

4/18/02<br />

0<br />

0 5 10 15 20<br />

20<br />

15<br />

10<br />

5<br />

1-Minute Wind Speed Measured with Cup Anemometer (m/s)<br />

y = 0.94x<br />

R 2 = 0.98<br />

4/19/02<br />

0<br />

0 5 10 15 20<br />

20<br />

15<br />

10<br />

5<br />

1-Minute Wind Speed Measured with Cup Anemometer (m/s)<br />

y = 0.91x<br />

R 2 = 0.99<br />

4/20/02<br />

0<br />

0 5 10 15 20<br />

1-Minute Wind Speed Measured with Cup Anemometer (m/s)<br />

Figure 5-5. Comparison between collocated spinning cup and sonic anemometers. a. Example time<br />

series from 4/18/02; b. regression of 1-second-averaged and 1-minute-averaged data from 4/18/02; c.<br />

regression of 1-second-averaged and 1-minute-averaged data from 4/19/02; d. regression of 1-secondaveraged<br />

and 1-minute-averaged data from 4/20/02.<br />

5-9


0.8<br />

Vertical velocity W (m/s)<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

-0.1<br />

-0.2<br />

Background<br />

interval #1:<br />

-37 to -7<br />

seconds<br />

Influence<br />

interval: -7<br />

seconds to<br />

+15 seconds<br />

Recorded<br />

pass time<br />

Background<br />

interval #2:<br />

+15 to +45<br />

seconds<br />

00.5 10.0 19.5 29.0 38.5 48.0 57.5 07.0 16.5 26.0 35.5<br />

Time (seconds)<br />

Figure 5-6. Example of W-velocity as measured by the sonic anemometer in the vicinity of a vehicle<br />

pass. The vertical lines indicate the start and end times for the two background intervals and the<br />

influence interval.<br />

5.2 Results<br />

5.2.1 Dispersion and deposition downwind of the unpaved road at Ft. Bliss<br />

For a ground-level release, it is as important to accurately represent dispersion as<br />

deposition. The two processes act simultaneously and the removal of particles downwind<br />

of a dust source is strongly dependent on both. The parameterization used for dispersion<br />

in the box model was discussed in Chapter 4. This section examines the ability of the<br />

numerical solution of the Atmospheric Diffusion Equation (ADE) and the Gaussian<br />

distribution used in the ISC3 model (USEPA, 1995) to represent the concentration profile<br />

downwind of a ground-level source. The results of the Ft. Bliss experiments are<br />

compared with predicted ADE and ISC concentration profiles. This comparison is only<br />

applicable to the distance between the unpaved road and the furthest downwind tower<br />

(100 m). Data for comparison of concentration profiles over longer distances may be<br />

obtained from a long-range scanning LIDAR that was operating concurrently with the<br />

tower measurements. However, those data were not available at the time of the writing<br />

of this report.<br />

Both the ADE and the ISC3 are designed to handle continuous release sources<br />

and consequently, downwind concentrations that do not vary with time (steady-state).<br />

The experiments at Ft. Bliss examined the downwind plume from a vehicle that traversed<br />

an unpaved road, first in one direction, then in the other. The result is that the time series<br />

of the concentration measured at the downwind towers more resembles an instantaneous<br />

puff release than a continuous source (see Figure 5-2). To compare the Ft. Bliss data<br />

with the models, the concentration time series was integrated for each vehicle pass and<br />

each location on the tower. That is<br />

5-10


k<br />

C int<br />

=<br />

peakend<br />

k<br />

<br />

C<br />

peeakstart<br />

dt<br />

Equation 5-5<br />

where C k int is the integral of the concentration C k at tower location k. For the purpose of<br />

comparison, it was necessary to use a common scale for concentration. This was<br />

accomplished by normalizing the concentrations to a common value for each tower.<br />

While the location of the particle monitoring instruments (DustTraks) varied slightly<br />

from one tower to the next, all towers were equipped with one monitor at a height of 1.26<br />

m. Thus, the integrated concentrations for each individual pass and at each height on a<br />

tower were normalized by the integrated concentration at 1.26 m for that tower. The<br />

concentration profile from each of 109 passes was placed in a category according to the<br />

value of the friction velocity, u * , associated with that pass. The normalized concentration<br />

profiles were then averaged within each friction velocity category. There were four<br />

categories in all for u * , corresponding to bins centered at 0.2 m/s, 0.3 m/s, 0.4 m/s, and<br />

0.5 m/s. For example, all concentration profiles with a u * between 0.15 m/s and 0.25 m/s<br />

were lumped together under the u * =0.2 category.<br />

The validity of using Equation 5-5 to estimate the concentration profile from a<br />

line source based on multiple passes of a moving point source (one vehicle traversing the<br />

road) was assessed. On two occasions during the field campaign, large military convoys<br />

traveled on the road near the towers. The convoys, consisting of tens of vehicles spaced<br />

approximately 10 meters apart, are an excellent approximation to a true line source.<br />

Figure 5-7 shows a comparison between the concentration profiles from a line source and<br />

those from multiple moving point sources. Allowing for some scatter, especially in the<br />

line source curves where only one data point was used for each convoy, the two sets of<br />

profiles are very similar for both the 50 m (T2) and 100 m (T3) downwind towers. This<br />

indicates that the treatment of an unpaved road as a line source is reasonable at the scale<br />

of 100 meters or so. At larger distances, the observed differences between a true line<br />

source and multiple moving point sources is expected to be small.<br />

The integrated concentration profile for three towers at multiple downwind<br />

distances from the road (7, 50, and 100 m) are shown in Figure 5-8 for two values of u * .<br />

All tests were performed on sunny days after 10:00 AM. Therefore atmospheric stability<br />

ranges from neutral to moderately unstable. As expected, the concentration profile is<br />

steep near the road at the first tower. The profile becomes less steep as particles mix<br />

vertically while being transported further downwind past the second and third towers.<br />

The profiles for the u * =0.3 m/s case indicate a greater degree of dispersion than those for<br />

the u * =0.5 m/s. This seems counter-intuitive since it is expected that higher values of<br />

friction velocity would enhance dispersion. The data shown in Figure 5-8 do not<br />

contradict that expectation; higher values of friction velocity also imply higher ambient<br />

wind speeds and shorter transport times. Thus, while on a time basis, dispersion<br />

increases with friction velocity, on a distance basis, the opposite appears to be true, at<br />

least for the conditions of the experiment at Ft. Bliss.<br />

5-11


14<br />

12<br />

10<br />

Puff_average T2 u*=0.2<br />

Convoy T2_u*=0.19<br />

Puff_average T2 u*=0.5<br />

Convoy T2_u*=0.47<br />

Height z (m)<br />

8<br />

6<br />

4<br />

2<br />

0<br />

0 0.2 0.4 0.6 0.8 1 1.2<br />

Normalized Concentration<br />

a. Tower 2 (50 m)<br />

14<br />

12<br />

10<br />

Height z (m)<br />

8<br />

6<br />

4<br />

Puff_average T3 u*=0.2<br />

Convoy T3_u*=0.19<br />

Puff_average T3 u*=0.5<br />

Convoy T3_u*=0.47<br />

2<br />

0<br />

0 0.2 0.4 0.6 0.8 1 1.2<br />

Normalized Concentration<br />

b. Tower 3 (100 m)<br />

Figure 5-7. Comparison of concentration profiles from the passage of two convoys with the<br />

concentration profiles at equivalent friction velocities based on the average from multiple passes of a<br />

moving point source. The line source is approximated by the passage of two convoys on two separate<br />

occasions. The moving point source average concentration profiles were calculated as the integrated<br />

concentration from Equation 5-5. Concentrations are normalized to the value at 1.26 meters in all<br />

cases.<br />

Figure 5-8 also shows the predicted concentration profile for each tower resulting<br />

from the numerical solution of the one-dimensional Atmospheric Diffusion Equation. It<br />

is clear from the figure that the ADE significantly over predicts dispersion in the nearfield.<br />

This in turn means that the ADE is expected to under predict deposition over the<br />

same distance scale. The ADE depends on mixing length theory for its validity. In<br />

particular, the equation adequately describes dispersion provided that the length scale for<br />

changes in the mean concentration is much larger than the length scale for turbulent<br />

transport (Seinfeld and Pandis, 1999). This condition does not hold true for the region<br />

near an isolated source such as an unpaved road, with the result that dispersion is<br />

overestimated near the unpaved road.<br />

5-12


Height z (m)<br />

14<br />

12<br />

10<br />

8<br />

6<br />

T1_data<br />

T2_data<br />

T3_data<br />

T1_ADE<br />

T2_ADE<br />

T3_ADE<br />

4<br />

2<br />

0<br />

0 0.2 0.4 0.6 0.8 1 1.2 1.4<br />

Normalized Concentration<br />

a. u * = 0.3<br />

Height z (m)<br />

14<br />

12<br />

10<br />

8<br />

6<br />

T1_data<br />

T2_data<br />

T3_data<br />

T1_ADE<br />

T2_ADE<br />

T3_ADE<br />

4<br />

2<br />

0<br />

0 0.2 0.4 0.6 0.8 1 1.2 1.4<br />

Normalized Concentration<br />

b. u * = 0.5<br />

Figure 5-8. Concentration profiles measured at the three towers downwind of an unpaved road and<br />

concentration profiles predicted by the ADE model for equivalent downwind distances a. for 32<br />

passes with u * between 0.15 and 0.25 m/s and b. for 15 passes with u * between 0.45 and 0.55 m/s.<br />

Tower distances from edge of road: T1 - 7 m, T2 – 50 m, T3 – 100 m. Modeled profiles are based on<br />

neutral atmospheric conditions. Both measured and modeled concentration have been normalized to<br />

the concentration at 1.26 m.<br />

It is interesting that the Gaussian model used in the ISC3 better emulates the<br />

concentration profiles observed in the experiment, especially since the gaussian plume<br />

equations are actually approximate analytical solutions to the Atmospheric Diffusion<br />

Equation. This is illustrated in Figure 5-9 where normalized concentrations at Towers 2<br />

and 3 are plotted on the same graph as the predicted gaussian concentration profile. The<br />

value of σ z in the ISC3 model is based only on atmospheric stability and downwind<br />

distance. Thus, friction velocity is not represented. This is partly why experimental<br />

conditions corresponding to low values of u * are better represented by the “moderate” to<br />

“very” unstable curves. Higher values of u * correspond to “slight” to “moderate”<br />

5-13


Height z (m)<br />

14<br />

12<br />

10<br />

8<br />

6<br />

T2 u*=0.2<br />

T2 u*=0.3<br />

T2 u*=0.4<br />

T2 u*=0.5<br />

T2_ISC_neutral<br />

T2_ISC_slightly unstable<br />

T2_ISC_moderately unstable<br />

T2_ISC_very unstable<br />

4<br />

2<br />

0<br />

0 0.2 0.4 0.6 0.8 1 1.2<br />

Normalized Concentration<br />

a. Tower 2 (50 m)<br />

Height z (m)<br />

14<br />

12<br />

10<br />

8<br />

6<br />

T3 u*=0.2<br />

T3 u*=0.3<br />

T3 u*=0.4<br />

T3 u*=0.5<br />

T3_ISC_neutral<br />

T3_ISC_slightly unstable<br />

T3_ISC_moderately unstable<br />

T3_ISC_very unstable<br />

4<br />

2<br />

0<br />

0 0.2 0.4 0.6 0.8 1 1.2<br />

Normalized Concentration<br />

b. Tower 3 (100 m)<br />

Figure 5-9. Normalized measured and modeled gaussian concentration profiles at a. Tower 2, 50<br />

meters downwind of unpaved road and b. Tower 3, 100 meters downwind.<br />

unstable conditions. The difference between the ADE and the ISC3 is that in the latter,<br />

the dispersion parameter (embodied in σ z ) is allowed to vary with downwind distance,<br />

while in the former (embodied in K zz ) the dispersion is only allowed to vary with height.<br />

In practical terms, this means that the ISC3 has a correction factor built into it to account<br />

for the shortcomings of the ADE close to a source. Figure 5-10 shows the downwind<br />

progression of the concentration profiles according to the ADE and the ISC3 model. The<br />

results for the ADE are shown for neutral conditions at two values of u * and those for the<br />

ISC3 are shown for three stability conditions ranging from neutral to very unstable. At<br />

far downwind distances (i.e. > 500 m), the profile for the ADE curve lies between<br />

moderately unstable and neutral profiles from the ISC3. This is consistent with the<br />

expectation that at further from the source, the assumptions of the ADE are better met<br />

than in the immediate vicinity of the source.<br />

5-14


Horizontal fluxes of PM 10 measured at each of the three downwind towers during<br />

the 2002 Ft. Bliss experiments appear in Figure 5-11a. The horizontal flux was<br />

calculated according to Equation 5-1 using PM 10 concentrations measured with<br />

DustTraks at multiple heights on each of the downwind towers. For each pass, the<br />

horizontal flux measured at each tower was normalized by the average flux for the three<br />

towers. Because of the great variability among fluxes for individual vehicle passes, the<br />

columns in the figure represent averages of the normalized flux over 27 passes when the<br />

wind direction was known to be within 15 degrees of the vector perpendicular to the road.<br />

The vertical bars are standard errors. Figure 5-11a indicates that there is no measurable<br />

reduction in PM 10 horizontal fluxes for at least 100 meters downwind of the unpaved<br />

road. The fluxes measured at T1 and T2 (50 m) are nearly identical, while those<br />

measured at T3 (100 m) are slightly higher, but within two standard errors of T1. This<br />

indicates that there is no statistically significant difference (α=0.05) between the<br />

horizontal fluxes measured at any of the three towers and that for the conditions at Ft.<br />

Bliss, the removal of PM 10 particles over a distance of 100 m downwind of the road is<br />

negligible.<br />

250<br />

200<br />

ISC3_neutral<br />

ISC3_moderately_unstable<br />

ISC3_Very_unstable<br />

ADE_u*=0.3<br />

ADE_u*=0.3<br />

250<br />

200<br />

ISC3_neutral<br />

ISC3_moderately_unstable<br />

ISC3_Very_unstable<br />

ADE_u*=0.3<br />

ADE_u*=0.3<br />

hieght z (m)<br />

150<br />

100<br />

hieght z (m)<br />

150<br />

100<br />

50<br />

50<br />

0<br />

0<br />

0.0 0.2 0.4 0.6 0.8 1.0 1.2<br />

0.0 0.2 0.4 0.6 0.8 1.0 1.2<br />

Normalized Concentration<br />

Normalized Concentration<br />

a. 50 m b. 100 m<br />

250<br />

200<br />

ISC3_neutral<br />

ISC3_moderately_unstable<br />

ISC3_Very_unstable<br />

ADE_u*=0.3<br />

ADE_u*=0.5<br />

250<br />

200<br />

ISC3_neutral<br />

ISC3_moderately_unstable<br />

ISC3_Very_unstable<br />

ADE_u*=0.3<br />

ADE_u*=0.5<br />

hieght z (m)<br />

150<br />

100<br />

hieght z (m)<br />

150<br />

100<br />

50<br />

50<br />

0<br />

0.0 0.2 0.4 0.6 0.8 1.0 1.2<br />

Normalized Concentration<br />

0<br />

0.0 0.2 0.4 0.6 0.8 1.0 1.2<br />

Normalized Concentration<br />

c. 500 m d. 1,000 m<br />

Figure 5-10. Comparison of ISC3 gaussian concentration profile and profile from numerical solution<br />

of Atmospheric Diffusion Equation at multiple downwind distances. Solutions for the ADE are for<br />

neutral conditions.<br />

5-15


Horizontal flux of PM10 by DustTrak normalized to the FLux at<br />

Tower1<br />

1.5<br />

1.4<br />

1.3<br />

1.2<br />

1.1<br />

1<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

T1 (7 m) T2 (50 m) T3 (100 m)<br />

a. Comparison of horizontal flux among the three downwind towers.<br />

Relative concentration of particles in size range<br />

normalized to mass in 3.9 micron size bin<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

Normalized concentration at 100 m (T3) divided by initial value<br />

(T3) of ratio of mass in bin to mass in 3.9 micron size bin<br />

ISC prediction (u* =0.35 m/s)<br />

ADE prediction (u* = 0.35 m/s)<br />

3.9 5.6 7.2 9.9 14 19.7<br />

Particle Diameter (microns)<br />

b. Comparison of particle size distributions at Towers 1 and 3<br />

Figure 5-11. Particle removal rates measured at Ft. Bliss at tower 100 m downwind of an unpaved<br />

road. a. Comparison of horizontal flux among the three downwind towers. Data are averages over 27<br />

individual passes when the wind direction was within 15 degrees of perpendicular to the road.<br />

Concentrations measured with DustTrak monitors with PM 10 inlets. Fluxes calculated according to<br />

Equation 5-1 and normalized to value at Tower 1. Vertical bars indicate standard errors; b.<br />

Comparison of particle size distributions at Towers 1 and 3 measured by GRIMM 1.108 OPC with<br />

model predictions. Data are averages over 137 individual passes. X-axis shows particle aerodynamic<br />

diameter calculated by assuming a density of 2.6 g/cm 3 and geometric mean diameter representing<br />

each size bin. Particle concentrations at Towers 1 and 3 were each normalized by concentration of<br />

3.9 micron particles. Y-axis represents ratios of normalized concentrations.<br />

The finding that there is no measurable removal of PM 10 particles can be further<br />

supported by examining the particle size distributions measured at T1 and T3. Figure<br />

5-11b shows the ratio of particle concentrations measured at the same height (2.66 m) on<br />

both towers. The figure represents data from 137 individual passes. The x-axis shows<br />

the mean geometric diameter for the size bin measured. Aerodynamic diameters have<br />

5-16


Table 5-2 enumerates the speed-dependent emission factors estimated from the<br />

measurements at Ft. Bliss alongside those determined according to Equation 5-6.<br />

Samples of road dirt were obtained both prior to and at the end of the field campaign.<br />

The silt contents for those samples were 7% and 4%, respectively. The higher silt<br />

content at the beginning of the study reflects two phenomena. First, on the day prior to<br />

sample collection, there was a windstorm that may have deposited some fine soil material<br />

onto the unpaved road. Second, the field tests themselves may have resulted in the<br />

depletion of fine soil material from the road. Thus, during the field tests, it is reasonable<br />

to assume that the silt content on the unpaved road was somewhere in between 4% and<br />

7%. The AP-42 factors calculated using these two values as lower and upper bounds<br />

appear in Table 5-2.<br />

04/23_22 Freightliner<br />

EF (g/vkt)<br />

2500<br />

2000<br />

1500<br />

1000<br />

y = 38.397x<br />

R 2 = 0.9458<br />

Average<br />

Linear (Average)<br />

500<br />

0<br />

0 10 20 30 40 50 60<br />

Speed (mph)<br />

Figure 5-12. Example of emission factor dependence on vehicle speed. The vertical bars are the<br />

standard deviations among the three downwind towers.<br />

In general, the AP-42 factors overestimate PM 10 emissions for vehicles traveling<br />

at low speeds compared to the measurements. This is illustrated in Figure 5-13, where<br />

the ratios of the AP-42 values to measured values are plotted as a function of vehicle<br />

speed. The top panel in the figure summarizes the data for all the vehicles in Table 5-2.<br />

The figure illustrates that, depending on the assumed silt content, for speeds<br />

approximately less than 20 mph, AP-42 overestimates PM 10 emissions while for speeds<br />

greater than 25 mph, AP-42 underestimates emissions. It is noteworthy that most<br />

emissions inventories assume that vehicles traveling on unpaved roads average speeds of<br />

either 20 or 25 mph. This suggests, that at least from the standpoint of emissions<br />

inventories, use of the data from the Ft. Bliss experiments would give results similar to<br />

the AP-42. However, the vehicles in this study are on average much heavier than one<br />

would expect to find on rural unpaved roads (industrial haul roads may be an exception).<br />

The bottom panel of Figure 5-13 shows that if only passenger-sized vehicles are<br />

considered, then the AP-42 may substantially overestimate PM 10 emissions for speeds up<br />

to about 35 mph. In fact, in the 20 or 25 mph range, the AP-42 values are 1.5 to 2 times<br />

those measured at Ft. Bliss. This suggests that the AP-42 emission factors may be<br />

positively biased for passenger-sized vehicles. This bias can be a source of discrepancy<br />

5-18


etween PM 10 emission inventories for crustal material and the amount of such material<br />

found on ambient filter samples.<br />

Table 5-2. Summary of vehicle emission factors measured at Ft. Bliss and Comparison with emission<br />

factors calculated as specified in AP-42 (USEPA, 1999) for two values of silt content (4% and 7%);<br />

moisture content was assumed to be 0.2%. The speed dependence of the measured emission factors<br />

is given in the third column and represents the slope of a linear regression between measured<br />

emission factor (g/vkt) and vehicle speed (mph); the R 2 value for the regression is given in the fourth<br />

column. For cases where a particular vehicle was tested more than once, the average of the two tests<br />

is used. All emission factors are reported in grams per vehicle kilometer traveled (g/vkt).<br />

vehicle/ speed<br />

GVWR<br />

(kg)<br />

Emission EF<br />

(g/vkt/ mph) R 2 mph<br />

Potential 10<br />

EF<br />

15<br />

mph<br />

Dodge Neon 1,622 9.8 0.37 82 123 164 205 246 287 328 369 410 247 387<br />

Dodge<br />

Caravan Test<br />

1 2,495 17.5 0.85<br />

Dodge<br />

Caravan Test<br />

1 2,341 15.5 0.70<br />

Dodge<br />

Caravan<br />

Average 2,418 16.5 138 208 277 346 415 484 553 623 692 294 460<br />

Ford Taurus 2,124 9.7 0.90 81 122 163 203 244 285 325 366 407 275 431<br />

Humvee Test<br />

1 2,445 10.0 0.93<br />

Humvee Test<br />

2 2,445 12.5 0.76<br />

Humvee<br />

Average 2,445 11.3 94 141 189 236 283 330 377 424 472 291 456<br />

EF<br />

20<br />

mph<br />

EF<br />

25<br />

mph<br />

EF<br />

30<br />

mph<br />

EF<br />

35<br />

mph<br />

EF<br />

40<br />

mph<br />

EF<br />

45<br />

mph<br />

EF<br />

50<br />

mph<br />

5-ton 9,818 76.0 0.82 637 956 1275 1593 1912 2230 2549 2868 3186 508 795<br />

LMTV 8,060 30.0 0.80 252 377 503 629 755 880 1006 1132 1258 469 735<br />

Freightliner<br />

M915A4 Test<br />

1 23,636 38.8 0.92<br />

Freightliner<br />

M915A4 Test<br />

2 23,636 38.4 0.95<br />

Freightliner<br />

M915A4<br />

Average 23,636 38.6 324 485 647 809 971 1133 1295 1456 1618 722 1130<br />

Average for<br />

all vehicles 230 345 460 574 689 804 919 1034 1149 401 627<br />

Average for<br />

small<br />

vehicles 99 149 198 248 297 347 396 446 495 277 433<br />

AP-<br />

42<br />

4%<br />

AP-<br />

42<br />

7%<br />

In addition to the neglect of vehicle speed, some of the inconsistencies in the AP-<br />

42 emission factors for PM 10 dust from unpaved roads come from the “lumping” of the<br />

vehicle weights in Equation 5-6. Emission factors calculated in this way are not selfconsistent.<br />

For example, consider three scenarios a. 1,000 vehicles weighing 1 ton each,<br />

b. 1,000 vehicles weighing 10 tons each, and c. a mix of 500 1-ton and 500 10-ton<br />

vehicles. According to Equation 5-6, the relative emissions of PM 10 for the 1,000<br />

vehicles in each of these three scenarios are respectively 1, 2.51, and 1.97. This goes<br />

against the logical result that one would expect, namely that the emissions in case c.<br />

should be equal to the average of those in a. and b (1.25).<br />

5-19


6<br />

Ratio of AP-42 Emission Factor to<br />

Measured Emission Factor<br />

5<br />

4<br />

3<br />

2<br />

silt content = 7%<br />

1<br />

Silt content = 4%<br />

0<br />

10 15 20 25 30 35 40 45 50<br />

Speed (mph)<br />

a. All vehicles in Table 5-2<br />

6<br />

Ratio of AP-42 Emission Factor to<br />

Measured Emission Factor<br />

5<br />

4<br />

3<br />

2<br />

1<br />

silt content = 4%<br />

silt content = 7%<br />

0<br />

10 15 20 25 30 35 40 45 50<br />

Speed (mph)<br />

b. same as a, but only considering Dodge Neonb, Dodge Caravan, Ford Taurus, and<br />

Humvee (GMC)<br />

Figure 5-13. The ratio of emission factors calculated according to AP-42 (USEPA, 1999) and those<br />

measured at Ft. Bliss. The upper line is the ratio based on a 7% silt content, while the lower line<br />

assumes 4% silt content. The top panel shows the ratios when all the vehicles in Table 5-2 are<br />

considered. The bottom panel shows the same data, but only for the passenger vehicles.<br />

5.2.3 Injection Height<br />

5.2.3.1 Estimates of injection heights<br />

The results of the nighttime wake turbulence tests are shown in Figure 5-14. Because of<br />

the inherent variability in the data, it was necessary to combine the passes at all speeds<br />

for each vehicle. This was accomplished by normalizing the square root of the difference<br />

between the vehicle-influenced vertical fluctuations in velocity and the background<br />

fluctuation with the vehicle speed. The figure illustrates that the size of the turbulent<br />

wake behind a vehicle increases with physical size. The 24-foot moving truck creates a<br />

measurable wake up to a height of approximately 6 meters. The smaller cargo van wake<br />

5-20


8<br />

7<br />

6<br />

Compact<br />

Cargo Van<br />

Box Truck<br />

height (z)<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

-0.005 0 0.005 0.01 0.015 0.02 0.025<br />

(j iave -j bave ) 1/2 /Vehicle speed<br />

Figure 5-14. Turbulence generated vs. height above ground for three vehicle types. The x-axis<br />

represents the difference between the vertical component of turbulence generated by the passing of<br />

the vehicle and the background fluctuations in vertical velocity normalized by the speed of the<br />

vehicle. The horizontal lines represent the background turbulence standard deviation. The dotted<br />

lines represent hand drawn curves to fit the data. The dotted line corresponding to the compact car<br />

was drawn based on the assumption that the height of the wake plume is proportional to the height of<br />

the vehicle (see Figure 5-15).<br />

Turbulent Wake height (m)<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

Compact Car<br />

estimated from<br />

regression<br />

y = 1.7x<br />

Cargo Van<br />

0 0.5 1 1.5 2 2.5 3 3.5 4<br />

Vehicle Height (m)<br />

24-foot Moving<br />

Truck<br />

Figure 5-15. Estimate of turbulent wake height vs. physical vehicle height. The dotted line represents<br />

a zero-intercept regression on the data from the Cargo Van and the Moving Truck. The hollow circle<br />

is an estimate of the wake height for the Compact Car based on the regression.<br />

only goes up to about 2.5 meters. The compact car is apparently too small to cause a<br />

measurable wake, even at the minimum setting for the anemometer height (0.9 meters).<br />

The dotted line drawn for the compact car is based on a linear fit of the wake heights of<br />

the box truck and the cargo van vs. their physical heights (3.2 and 2.0 meters,<br />

respectively). The compact car is 1 meter in height (measured from the top of the trunk<br />

5-21


to the ground) giving an approximate wake height of 1.7 meters according to this<br />

regression (see Figure 5-15).<br />

Turbulence is a surrogate for mixing. The wake height measured in this portion<br />

of the study is important because it gives an indication of the height to which the dust<br />

plume is mixed behind a vehicle traversing an unpaved road. As a first approximation,<br />

we may assume that the dust emitted behind a vehicle is well-mixed up to the height of<br />

the turbulent wake (approximately 1.7 times the height of the vehicle itself). This<br />

“injection” height may play an important role when determining the fraction of<br />

particulate matter that deposits close to the road vs. the fraction that is transportable; as a<br />

rule, the further a particle is from the ground initially, the less likely it is to deposit in a<br />

given period of time. As an aside, the injection height also poses a lower limit on the<br />

height that can be used in Gillette’s model of a hypothetical box.<br />

The approximation for the injection height suggested here is quite crude and<br />

simplistic. The height of the wake is likely to depend on a number of parameters, most<br />

notably, the atmospheric stability, the shape of the vehicle, and the angle of the ambient<br />

wind with respect to the direction of vehicle travel. The analysis is based on data<br />

obtained under stable conditions. Under unstable conditions, turbulent mixing is not<br />

inhibited by buoyancy forces. Therefore, it is likely that the effective injection height will<br />

be larger than under stable conditions. For the present purpose, it is sufficient to note that<br />

the height to which a dust plume is thoroughly mixed in the wake of a vehicle is<br />

dependent on the size of the vehicle, increasing in a roughly linear relationship with<br />

height.<br />

Detailed discussion of vehicle wake characteristics is outside the scope of this<br />

project. Hosker (1982) provides a thorough literature review of flows around bluff<br />

bodies. More recently, Rao et al. (2002) have measured wake properties with a trailer that<br />

was drawn by a cargo van and equipped with multiple sonic anemometers. Hider et al.<br />

(1997) have examined a modeling approach for particles dispersing in a vehicle wake.<br />

Though preliminary and based only on a single vehicle type, their data suggest that the<br />

turbulent kinetic energy in the wake of the vehicle is greatest at approximately the vehicle<br />

height and approaches background levels at twice the vehicle height. This is qualitatively<br />

consistent with our assumption of a wake plume that is completely mixed from the<br />

ground up to 1.7 times the vehicle height.<br />

5.2.3.2 The effect of injection height on removal of particles downwind of an<br />

unpaved road<br />

The injection height is least important under unstable atmospheric conditions, and<br />

most important under stable atmospheric conditions. This is because under unstable<br />

conditions, the dust plume is lofted up high quickly anyway, more or less regardless of<br />

the starting height. Under stable conditions, the extent of vertical mixing is retarded by<br />

buoyancy, and a lower injection height allows the particles to be closer to the ground for<br />

a longer period of time, thereby enhancing deposition.<br />

According to the ADE model, overall, the effect of injection height on particle<br />

removal is quite small. The reason for this was discussed earlier in section 5.2.1, namely<br />

that the dispersion near the source (nominally less than 1,000 m downwind) is greatly<br />

5-22


overestimated by the ADE since some of the basic model assumptions are violated in that<br />

region.<br />

The ISC model more accurately reflects the degree of dispersion that occurs in the<br />

first 1,000 or so meters downwind of the unpaved road source. Recall that the injection<br />

height of the dust plume can be accounted for in the model through the initial dispersion<br />

parameter σ z , assumed to equal one-half of the injection height, and a “virtual” upwind<br />

distance (see section 4.1.2). Comparison of downwind removal rates (Figure 5-16)<br />

indicates that for unstable conditions, the injection height has a non-trivial, but small<br />

effect. For very stable conditions, the injection height has a large effect, approaching a<br />

factor of two difference in estimated removal rates at 1,000 m downwind of the source.<br />

The difference between the removal of particles for two plumes with different injection<br />

heights is confined to the near-source region. That is, the difference does not continue to<br />

grow past about 1,000 meters downwind (Figure 5-16). With this in mind, and noting as<br />

before that most unpaved road dust emissions are likely to occur in the daytime, when<br />

conditions are neutral to unstable, the effect of the injection height is secondary to the<br />

uncertainties associated with deposition velocities and dispersion parameters.<br />

1.2<br />

1.2<br />

Fraction of particles remaining in<br />

suspension<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

IH=1<br />

IH=2<br />

IH=4<br />

IH=6<br />

Fraction of particles remaining in<br />

suspension<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

IH=1 m<br />

IH=2 m<br />

IH=4 m<br />

IH=6 m<br />

0<br />

0<br />

0 200 400 600 800 1000<br />

0 200 400 600 800 1000<br />

Distance Downwind (m)<br />

Distance Downwind (m)<br />

a. Moderately unstable (Class B) b. Neutral (Class D)<br />

1.2<br />

1.2<br />

Fraction of particles remaining in<br />

suspension<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

IH=1<br />

IH=2<br />

IH=4<br />

IH=6<br />

Fraction of particles remaining in<br />

suspension<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

IH=1<br />

IH=2<br />

IH=4<br />

IH=6<br />

0<br />

0 200 400 600 800 1000<br />

Distance Downwind (m)<br />

c. Very stable (Class F)<br />

0<br />

0 2000 4000 6000 8000 10000<br />

Distance Downwind (m)<br />

c. Same as c. except x-axis extends to 10<br />

km<br />

Figure 5-16. Comparison of fraction of particles removed according to the ISC3 model for various<br />

values of the injection height under a. moderately unstable conditions; b. neutral conditions; c. very<br />

stable conditions; and d. same as c. but x-axis extends to 10 km. The injection height is manifested in<br />

the model by setting the initial value of σ z to 0.5×IH.<br />

Another aspect of the injection height that is not examined here is its effect on the<br />

removal efficiency for particles in the “impact” region. This region (labeled Region A in<br />

Figure 2-1) may play an important role in the removal of particles, especially if the<br />

injection height is lower than or comparable to the vegetative canopy height. None of the<br />

5-23


models analyzed in this study is suited for estimating the interactions of particles with<br />

obstructions in the immediate vicinity of the source (i.e. at the junction of the unpaved<br />

road shoulder and the first row of vegetation downwind of the road). For estimating<br />

particle deposition in this region, it is preferable to use a technique similar to the one<br />

presented by Raupach and Leys (1999), who have developed a method to estimate the<br />

removal of particles by windbreaks. Their analysis is applicable to windbreaks of finite<br />

depth, but it may be possible to extend it to entire canopies with some modifications.<br />

5.3 Summary<br />

Field experiments at the Ft, Bliss military facility indicated that for particles<br />

covered by the PM 10 size range, removal by deposition was not measurable at a distance<br />

of 100 m. This was confirmed by comparison with predictions by the ISC and the ADE<br />

models. Based on these findings, it was concluded that under neutral to unstable<br />

conditions in sparsely vegetated arid shrubland, which may comprise a large portion of<br />

the southwestern United States, removal of particles in the near source region by<br />

deposition is not an important consideration.<br />

The ADE model was found to over predict dispersion very close to the source (<<br />

100 m), but the ISC concentration profiles fit the data reasonably well. One shortcoming<br />

of the ISC algorithm is the lack of ability to represent varying friction velocities.<br />

The turbulence behind moving vehicles was measured to quantify the height of<br />

the wake, termed the “injection height”. Using very simple analysis, it was determined<br />

that the size of the wake varies roughly linearly, by a multiple of 1.7, with the height of<br />

the vehicle. The injection height was found to have a large effect on the removal of<br />

particles downwind of an unpaved road, though only for stable atmospheric conditions.<br />

Comparison of emission factors measured at Ft. Bliss and those calculated<br />

according to the AP-42 guidance document (USEPA, 1999) indicated that by not<br />

accounting for vehicle speed, the AP-42 estimates are likely to be too high for vehicles<br />

traveling at low speeds. When only passenger-sized vehicles were considered, the Ft.<br />

Bliss data suggest that at 20 to 25 mph (a speed that is commonly assumed for unpaved<br />

road travel in emission inventories), AP-42 over predicts PM 10 emissions by 50%-100%<br />

compared to the measurements. When larger vehicles were included in the analysis, the<br />

AP-42 factors were in line with those measured for travel speeds of 20 to 25 mph. One<br />

reason cited for this apparent dependence on the distribution of vehicle size was the use<br />

of an aggregate-average weight in the AP-42 equation. The “lumped” weight approach is<br />

not self-consistent and is likely to result in inventories that overestimate the contribution<br />

of unpaved road dust to ambient PM 10 concentrations. This source of error should be<br />

addressed as part of the effort to reconcile emission inventories for geologically-derived<br />

PM 10 with the amount of crustal material found on filters measuring ambient<br />

concentrations.<br />

5-24


6. DETERMINATION <strong>OF</strong> PARTICLE REMOVAL RATES UNDER<br />

STABLE CONDITIONS IN A MOCK URBAN SETTING<br />

As part of a DoD-funded study, the University of Utah measured vehiclegenerated<br />

road dust under stable atmospheric conditions at a flat site in the Utah west<br />

desert, and the horizontal flux of dust with the aid of interpolating functions that were fit<br />

to the wind speed and dust concentration measurements at discrete heights. The<br />

measured horizontal flux of dust and particle removal by deposition were compared to<br />

predictions by the one-dimensional Atmospheric Diffusion Equation (ADE) and the<br />

Industrial Source Complex (ISC) models.<br />

6.1 Experimental Methods<br />

The field measurements were conducted at the Dugway Proving Ground, Tooele<br />

County, Utah in collaboration with the Mock Urban Setting Test (MUST), an<br />

atmospheric dispersion experiment funded by the Defense Threat Reduction Agency<br />

(Biltoft, 2001; Yee and Biltoft, 2002). The site is located at 40° 12.6' N, 113° 10.6' W<br />

and is 1310 m above sea level. The configuration, shown schematically in Figure 6-1,<br />

consisted of a 10 by 12 array of 2.5 m high by 2.4 m wide by 12.2 m long rectangular<br />

cargo shipping containers simulating buildings. The area surrounding the test site has a<br />

slope of 0.5 m / km rising to the south. Sand dunes 4-6 m high are 1 km to the north.<br />

Vegetation is a thin cover of brush (shadscale and black greasewood) 0.5 to 1 m high.<br />

The soil is classified by the Natural Resources Conservation Service (2000) as Skumpah<br />

silt loam and has 15% silt content. Screen analysis of a sample from the test road is<br />

provided in Table 6-1.<br />

180 m<br />

9 8 7 6 5 4 3 2 1 0<br />

A<br />

B<br />

C<br />

D<br />

176 m<br />

4<br />

E<br />

F<br />

G<br />

H<br />

I<br />

J<br />

95 m<br />

30 m<br />

1<br />

2 3<br />

Y<br />

Unpaved Test Road<br />

X<br />

True<br />

North<br />

K<br />

3 m<br />

L<br />

Figure 6-1. Plan view of the MUST site showing the array of shipping containers representing<br />

buildings and the location of the measurements. (1) 2-D sonic anemometers at 4, 8, 16 m; (2)<br />

DustTraks at 0.9, 1.7, and 3.7 m; (3) 3-D sonic anemometer at 1.6 m; (4) DustTraks at 1.8, 4.6, 9.1,<br />

and 18.3 m and 2-D sonic anemometers at 4, 8, 16, 32 m.<br />

6-1


Table 6-1. Screen analysis of test road surface material. Average of two soil samples.<br />

Cumulative<br />

+4 mesh 1.00<br />

-4 mesh 0.88<br />

-18 mesh 0.81<br />

-30 mesh 0.70<br />

-50 mesh 0.52<br />

-100 mesh 0.40<br />

-200 mesh 0.16<br />

Silt Content 15.7%<br />

The vehicle activity was designed to approximate a line source where discrete<br />

puff releases create a uniform cloud in the crosswind (y) direction. A 1994 Ford pickup<br />

truck (3900 kg maximum GVW) was driven on a graded, native soil road running parallel<br />

to the upwind edge of the container array. At 1-1.5 minute intervals, the vehicle was<br />

driven at 9 m s -1 (20 mph) along a test section of road that extended beyond the end of<br />

the container array in each direction.<br />

Dust concentration was measured using seven portable DustTrak analyzers (TSI,<br />

Inc. St Paul, MN) with PM 10 inlets. The DustTrak uses a 90° light scattering laser diode<br />

sensor and has a range of 0.001 - 100 mg/m 3 . Reported mass concentration was based on<br />

the factory calibration. Flow rate, zero on filtered air, and the instrument clock were<br />

checked daily.<br />

Instrument locations are shown in Figure 6-1. The near-source dust concentration<br />

measurement consisted of DustTraks at 0.9, 1.7, and 3.7 m above grade located 3 m from<br />

the edge of the road or about 4.5 m from centerline of vehicle travel. Two DustTraks<br />

were mounted on a stepladder placed in the open aisle between containers and the highest<br />

DustTrak was offset parallel to the road and located 1.2 m above the top of the first row<br />

of shipping containers. The downwind dust concentration was measured at a 32 m high<br />

tower located at the center of the container array. The DustTraks were 95 m from the<br />

centerline of vehicle travel at 1.8, 4.6, 9.1, and 18.3 m above grade. The dust<br />

concentration versus time data were treated as a series of 44 measurements of the puff<br />

generated by a discrete vehicle trip using both the peak concentration value for each trip<br />

(mg m -3 ) and the time integrated area under the concentration curve (mg sec m -3 per<br />

trip).<br />

Sonic anemometers provided time-resolved wind data upwind of the container<br />

array, between the test road and the first container, and in the center of the array.<br />

Ambient conditions were monitored at a permanent meteorological station (DPG08)<br />

located about 1 km from the test site.<br />

6.2 Data Analysis Methods<br />

The horizontal flux of dust is the product of the dust concentration times the wind<br />

speed integrated from ground level to the top of the dust cloud. For a line source, the<br />

6-2


units are particulate matter mass per length of road per vehicle trip. Defining the<br />

coordinate axes as x perpendicular to the road, y parallel to the road and z vertical, and<br />

assuming dust concentration and wind profile do not vary parallel to the road, the mass<br />

flux per length of road passing through a plane at constant distance from the road is:<br />

F dust =<br />

z=∞<br />

<br />

t=tmax<br />

<br />

z=0 t=0<br />

C(x, z,t) u(x, z,t) dt dz<br />

Equation 6-1<br />

where C is the dust concentration (mg/m 3 ), u (m s -1 ) is the wind component<br />

perpendicular to the y, z plane, and tmax is the trip interval or other averaging time.<br />

The dust flux calculation approach used in this portion of the study was to fit<br />

plausible interpolation functions to the measurements and then numerically integrate<br />

using the interpolation function values evaluated at each height step. This algorithm has<br />

the advantage of being mathematically well defined and providing insight into the<br />

sensitivity of the results to wind speed and concentration vertical profiles.<br />

6.2.1 Wind Speed Models<br />

An empirical power law equation can give a reasonable approximation to the<br />

wind profile over a limited range of height. This gives<br />

<br />

u = u ref<br />

<br />

<br />

z<br />

z ref<br />

<br />

<br />

<br />

P<br />

Equation 6-2<br />

where u ref is the wind speed measured at height z ref . The power law interpolation<br />

function has only one adjustable parameter, extrapolates to zero velocity at ground level,<br />

and is mathematically convenient.<br />

6.2.2 Vertical Dust Concentration Models<br />

A first-order exponential decay model was used for dust concentration<br />

C(z) = A exp( −Bz)<br />

Equation 6-3<br />

where A and B are empirical fitting parameters. An exponential equation for dust<br />

concentration was derived by Goosens (1985) by assuming that the change in eddy<br />

diffusivity with height is a power law function. Coefficients were obtained by a leastsquares<br />

fit to the time-averaged field data.<br />

Stepwise expressions for wind speed and concentration were compared to the<br />

interpolation function method. For the stepwise method, the value of the bottom<br />

measurement was used for the interval from ground level to the midpoint between the<br />

bottom and second measurements where there was a discontinuous change to the new<br />

6-3


value which was used up to the midpoint between the second and third measurement, and<br />

so on.<br />

6.2.3 Calculation of Horizontal Flux<br />

The power law wind profile is appealing since analytical solutions exist for the<br />

integral of a power law for wind with either an exponential or a Gaussian function for<br />

dust concentration. For the exponential dust concentration case:<br />

F dust =<br />

<br />

C(z) u(z) dz =<br />

∞<br />

( Ae − Bz)<br />

<br />

<br />

z<br />

u ref<br />

<br />

<br />

0<br />

z ref<br />

<br />

<br />

<br />

P<br />

<br />

− P Γ( P +1)<br />

dz = Au ref z ref<br />

<br />

B ( P+1)<br />

Equation 6-4<br />

where Γ is the gamma function. Using a mathematical model that results in a converging<br />

definite integral avoids any uncertainty about the upper limit of integration, z max . The<br />

integral of a power law for wind, times a power law for dust concentration has an analytic<br />

solution, but does not converge. For this study, the horizontal dust flux was also<br />

calculated by trapezoidal rule numerical integration of Equation 6-1 using alternative<br />

interpolation equations to model the time averaged wind speed and concentration as a<br />

function of height. Variable step size (starting at ∆z = 0.1 m) was used to provide higher<br />

resolution near the ground where both concentration and wind speed change rapidly with<br />

height.<br />

6.3 Results<br />

The data were collected during two vehicle activity periods between 01:00 and<br />

02:30 Mountain Daylight Time on September 26, 2001. The clear sky, nighttime, desert<br />

conditions resulted in a stable atmosphere during the sampling period. Wind speed was<br />

generally between 1 and 5 m s -1 , which was sufficient to move the dust cloud, and the<br />

wind direction relative remained within 45° of perpendicular to the container array and<br />

road. General meteorological data are summarized in Table 6-2. Table 6-3 gives the<br />

wind speed and dust concentration mean and standard deviation.<br />

Table 6-2 Meteorological data from station DPG08, approximately 1 km south of test site, for<br />

September 26, 2002 01:00-03:00 MDT.<br />

Average Range<br />

Temperature 294 K 293 - 297 K<br />

Relative Humidity 23% 20-26%<br />

Barometric Pressure<br />

866.54 mbar<br />

24 hr- Precipitation 0<br />

Wind at 10 m 3.0 m s -1 15 m s -1 gust<br />

Wind Direction from True North 98-186°<br />

6-4


Table 6-3. Mean ± one standard deviation wind and concentration data for the entire 1.5 hr test<br />

period with 44 vehicle trips. Wind standard deviation based on variation between 15-minute<br />

averages. Concentration standard deviation based on variation between individual vehicle trips.<br />

Pulse area is the time-integrated concentration for the interval corresponding to one vehicle pass.<br />

Wind<br />

Height<br />

m<br />

Upwind<br />

m s -1<br />

4 3.42 ± 0.51 2.22 ± 0.20<br />

8 4.07 ± 0.50 3.44 ± 0.32<br />

16 4.72 ± 0.52 4.59 ± 0.25<br />

32 4.89 ± 0.31<br />

Concentration<br />

Near Measurement<br />

3 m from road<br />

Height<br />

m<br />

Pulse Area<br />

mg s m -3 per trip<br />

32 m Tower<br />

m s -1<br />

Peak Height<br />

mg m -3<br />

Far Measurement<br />

95 m from road<br />

Pulse Area Peak Height<br />

mg s m -3 per trip mg m -3<br />

0.9 302 ± 171 38.9 ± 21<br />

1.7 144 ± 92 19.9 ± 11<br />

1.8 24.3 ± 17 0.72 ± 0.52<br />

3.7 77.8 ± 56 10.3 ± 7.2<br />

4.6 10.7 ± 7.2 0.41 ± 0.31<br />

9.1 3.18 ± 3.1 0.21 ± 0.17<br />

18.3 0.35 ± 0.1 0.019 ± 0.016<br />

6.3.1 Wind Profile<br />

Figure 6-2 shows the wind speed versus height with the logarithmic and power<br />

law interpolation equations superposed on the 15-minute average data points. Both<br />

equations fit the upwind data within the variation of the measurements. Based on the<br />

field measurements, the fitted coefficients for Equation 6-2 upwind and at the center of<br />

the array are:<br />

Upwind<br />

Center of Array<br />

U ref at z ref, m/s 3.42 2.22<br />

Z ref, m 4 4<br />

P 0.234 0.414<br />

6.3.2 Dust Concentration<br />

Typical dust concentrations versus time data are shown in Figure 6-3.<br />

Measurement locations are identified by the x, z coordinates in meters where x is<br />

downwind from the road and z is height above grade. Near-source dust clouds from<br />

individual vehicle trips appear as clearly defined spikes in the particle concentration<br />

6-5


measured at 3,1. Downwind the dust clouds were lower in magnitude and had longer<br />

duration. It was still possible to resolve peaks corresponding to the individual vehicle<br />

trips as shown by the time trend for the 95, 4.6 monitor. The trend for the highest<br />

measurement, 18.3 m above grade, shows that this instrument was above the top of the<br />

dust cloud for most trips. A floodlight allowed observation of the dust, which confirmed<br />

that the visible cloud was less than half the height of the 32 m tower. The background<br />

dust concentration was approximately 10µg m -3 during periods with no vehicle activity.<br />

Based on field measurements, the fitted coefficients Equation 6-3 near the road and at the<br />

center of the array are:<br />

Near (3 m)<br />

Center of Array<br />

A, mg/m3 594.9 41.15<br />

B 0.725 0.289<br />

Upwind<br />

Array Center<br />

35<br />

35<br />

30<br />

30<br />

25<br />

25<br />

20<br />

20<br />

15<br />

15<br />

10<br />

10<br />

5<br />

5<br />

0<br />

0 2 4 6 8<br />

m/s<br />

0<br />

0 2 4 6 8<br />

m/s<br />

Figure 6-2. Wind Equation Fit. Both the logarithmic (dashed line) and power law (solid line)<br />

equations give a reasonable fit to the 15-minute average wind speed measurements (circles).<br />

6.3.3 Horizontal Dust Flux<br />

The horizontal dust flux decreased to less than 15% of the near-source value as<br />

the cloud passed over and through the container array. The horizontal dust flux passing<br />

through vertical planes 3 and 95 m from the vehicle activity was calculated using the<br />

power law interpolation function for the wind speed and the exponential equation for the<br />

dust concentration. The analytical integration of the flux, Equation 6-4, provides the<br />

following:<br />

6-6


Near (3 m) Center of Array Ratio<br />

Horizontal dust flux, mg/m trip 1990 261 13.6%<br />

Numerical integration of the flux using alternative models for wind speed, (power<br />

law, logarithmic, and step), and for dust concentration, (exponential, power law, and<br />

step), gave similar results. The ratio of downwind to upwind dust flux ranged from 9% to<br />

15% for the different cases. The numerical integration shows that most of the dust flux is<br />

below 5 m. The observation of a significant decrease in the flux ratio at 3 and 95 m<br />

appears robust, but the calculated absolute values vary between the different assumptions<br />

of how the wind speed and dust concentration vary near the ground. The calculated nearroad<br />

dust flux ranges from 1.1 to 4.4 g per vehicle meter traveled which can be compared<br />

to 1.8 g per vehicle meter traveled obtained using the correlation for silt fraction and<br />

vehicle weight from AP-42.<br />

x = 95, z = 18.3<br />

0.1<br />

mg / m 3<br />

0.075<br />

0.05<br />

1<br />

0.025<br />

0<br />

2:14 2:16 2:18 2:20 2:22 2:24<br />

2<br />

x = 95, z = 4.6<br />

mg / m 3<br />

1.5<br />

3<br />

1<br />

4<br />

0.5 1 2<br />

5 6<br />

7<br />

0<br />

2:14 2:16 2:18 2:20 2:22 2:24<br />

x = 3, z = 1<br />

mg / m 3<br />

100<br />

75<br />

50<br />

25<br />

1<br />

2<br />

3<br />

0<br />

2:14 2:16 2:18 2:20 2:22 2:24<br />

Local Time<br />

4<br />

5<br />

6<br />

7<br />

Figure 6-3. Dust concentration versus time as measured by DustTraks. Vehicle passes generated<br />

well-defined spikes at 3 m horizontal and 1 m vertical from the road (bottom). The spikes were<br />

broader but still well defined 95 m downwind and 4.6 m above grade (middle). Most vehicle trips did<br />

not cause a noticeable spike at 95 m downwind and 18.3 m above grade (top). Note that full scale for<br />

the top graph and middle graph are 1/1000 and 1/50 of full scale for the near road measurements<br />

respectively.<br />

Figure 6-3 shows large variations in peak dust cloud concentrations even though the<br />

upwind conditions and vehicle travel were relatively constant. A large part of this<br />

variation is due to non-uniformity in the vehicle dust cloud.<br />

Figure 6-4 shows the results of a separate experiment where 5 DustTraks were<br />

collocated parallel to a road (Seshadri, 2002). Over the entire experiment, 18 vehicle<br />

6-7


trips, the average readings for the five instruments differed by only 0.06 mg/m 3 which<br />

was not statistically significant (P>0.77), and which is small compared to peak readings<br />

of 20-50 mg/m 3 . The time trend shows differences in peak heights and in the arrival time<br />

of the peak. The normalized range for sets of simultaneous readings, defined as<br />

(maximum - minimum)/average, had a mean of 0.96 with a maximum of 3.5, as shown in<br />

the distribution plot. This type of short duration random variation within the dust cloud is<br />

consistent with turbulent mixing.<br />

3.5<br />

3.0<br />

2.5<br />

Range<br />

Average<br />

0.7<br />

0.6<br />

0.5<br />

2.0<br />

1.5<br />

1.0<br />

0.5<br />

0.0<br />

mg/ m 3<br />

0.2<br />

0.1<br />

0.0<br />

14:09 14:10 14:11<br />

Figure 6-4 Collocated DustTraks. Five collocated DustTraks showed large differences in individual<br />

readings but little difference between the instruments when averaged over the entire experiment.<br />

Left: Histogram of maximum minus minimum divided by the average for sets of five coincident<br />

readings. Middle: outlier box plot of the same data. Right: a typical time trend showing poor<br />

correlation between individual readings.<br />

6.4 Discussion<br />

This study supports the hypothesis proposed by Watson, Chow et al. (2000) that<br />

not all dust initially suspended by vehicles is transported far downwind. The<br />

measurements reported in this study were made under stable atmospheric conditions at a<br />

flat site with large solid roughness elements between the test road and the downwind<br />

measurement location. Atmospheric stability and the terrain irregularities have not been<br />

variables in most studies of unpaved road dust. Additional field studies under a range of<br />

atmospheric and terrain conditions will be needed to determine when deposition is<br />

significant and to develop models for predicting the magnitude.<br />

More accurate measurements of dust flux will require better characterization of<br />

dust concentration and wind speed near the ground. The analytical interpolation<br />

equations and the empirical results of this study can be used to determine the optimum<br />

locations of wind speed and dust concentration instruments for future fieldwork.<br />

6.4.1 Comparison of measurements with model results<br />

The experiment at the Dugway Proving Ground, UT, was performed late at night<br />

when the atmosphere was stable. The wind speed varied between 1 and 5 m/s, resulting<br />

in an average friction velocity of 0.23 m/s (z 0 =0.1 m) upwind of the container array, and<br />

0.39 m/s (z 0 =0.71 m) at the center of the container array. This situation presents a unique<br />

set of circumstances where although the atmosphere is stable to very stable, a high<br />

6-8


friction velocity is made possible by the abrupt change in land use from the desert shrub<br />

surroundings to an array of cargo containers. Under these conditions, high stability<br />

concurrent with fairly high friction velocity (therefore also deposition velocity), the<br />

removal rate for particles is expected to be significant, even over the short distance of 95<br />

m between the unpaved road and the instrumented tower.<br />

Figure 6-5 compares the measured PM 10 removal rate from the Dugway<br />

experiments with the estimated removal rate from the ISC and the ADE models. The<br />

PM 10 DustTrak instruments used in the field do not provide an estimate of the particle<br />

size distribution. Therefore, removal rates based on measurements with those<br />

instruments are applicable to the entire fraction of particles with diameters less than about<br />

10 microns. Since the models use specific deposition velocities for each of the particle<br />

sizes, comparison of measured and modeled values requires some assumptions about the<br />

particle size distribution during the Dugway experiments. For the purposes of<br />

illustration, the average particle size distributions obtained during the Ft. Bliss<br />

experiments (Chapter 5) was used. The white circles in Figure 6-5 represent the fraction<br />

of the total PM 10 mass that is associated with particles in the stated size bin (sum is equal<br />

to unity). The gray circle is the measured PM 10 removal rate 95 meters downwind of the<br />

unpaved road. The black symbols correspond to the removal rates for the stated particle<br />

sizes estimated from the ISC and ADE models. The modeled overall removal rate for<br />

PM 10 (“Total” in the figure) was calculated by multiplying the fraction of the mass<br />

associated with each size by the removal rate predicted for that size and then summing<br />

the values for all four sizes shown.<br />

The previously noted trend of higher removal rates predicted by the ISC than the<br />

ADE is readily apparent in the figure. According to the ISC, 56% of the 9.9 micron<br />

particles and 22% of 7.2 micron particles deposit within the first 95 meters of the road.<br />

Removal rates for smaller particle sizes are negligible. The model predicts that 30% is<br />

deposited over the same distance when the entire PM 10 size fraction is considered. While<br />

significant, and clearly larger than the removal rates for the neutral to unstable conditions<br />

at Ft. bliss, the modeled removal of PM 10 falls short of the measured values (87%<br />

removal) by a considerable amount. The reason for this disparity is not known. The large<br />

roughness elements in the Dugway container array were intended to simulate an urban<br />

area. Perhaps the parameterization of the deposition velocity (Equation 2-26) used for<br />

the case of vegetative canopies does not readily extend to other land uses such as urban<br />

settings. In particular, the air flow around bluff, impermeable bodies may deviate<br />

significantly from what occurs in a comparatively porous vegetative canopy. Related to<br />

this point, it is possible that the dispersion parameter (σ z in the case of the ISC) does not<br />

adequately represent the channeling of the airflow through the spaces between the<br />

containers or the stagnation regions in their wake. Despite the models substantial under<br />

prediction of PM 10 removal, some solace can be found in noting that most unpaved roads<br />

are not located in urban areas and that it is likely that the majority of traffic on unpaved<br />

roads occurs during the daytime when the atmosphere is either neutral or unstable.<br />

6-9


Fraction of particles of stated size remaining in<br />

suspension 95 m downwind of unpaved road<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

ISC prediction (u*=0.4 m/s, very stable)<br />

ADE prediction (u*=0.4 m/s, very stable)<br />

Mass Fraction Associated with Particle Size (based<br />

on Ft. Bliss size distribution)<br />

Measured suspended fraction remaining<br />

3.9 5.6 7.2 9.9 Total<br />

Aerodynamic Diameter (microns)<br />

Figure 6-5. Comparison of fraction of PM 10 particles remaining in suspension 95 m downwind of an<br />

unpaved road during a nighttime test on Dugway Proving Grounds, UT. The white circles represent<br />

the fraction of PM 10 associated with each size bin. Mass fractions associated with each size bin were<br />

estimated from Ft. Bliss data from a different experiment. The gray circle is the fraction remaining<br />

in suspension according to the methods described in the previous section. The black squares and<br />

triangles correspond to modeled removal rates under very stable conditions.<br />

6-10


7. SUMMARY <strong>AND</strong> CONCLUSIONS<br />

Six major conclusions were formed from the models and field data examined in<br />

this report. We summarize these conclusions below and discuss their ramifications with<br />

respect to recent and ongoing work in the next subsection.<br />

First, the hypothesis that fugitive dust particles emitted from unpaved roads can<br />

deposit to an appreciable extent within several hundred meters downwind of the road<br />

(Watson and Chow , 2000; Countess, 2001) is well-founded, though the fraction of<br />

particles removed varies greatly based on setting and atmospheric stability. Two recent<br />

studies with robust datasets were examined. The field study at Ft. Bliss provided<br />

information on dust particle removal in a sparsely vegetated terrain, under neutral to<br />

unstable atmospheric conditions. The experiments at the Dugway Proving Ground in<br />

Utah occurred under nighttime stable conditions with large roughness elements<br />

downwind of the unpaved road. These two studies provide upper and lower bounds for<br />

the transportable fraction of PM 10 dust emissions.<br />

A field study set at the Ft. Bliss military installation in El Paso, TX took place in<br />

April 2002. An unpaved road, located in a desert open range, sparsely peppered with<br />

small shrubs, was used to conduct a series of emissions tests. Three towers, 9 to 12 m in<br />

height, were placed downwind of the road at distances of 7, 50, and 100 m. A variety of<br />

vehicles were driven back and forth along the road. DustTrak particle monitors located at<br />

several heights along the towers were used to capture the PM 10 concentrations due to the<br />

passing dust plumes. Wind speed and direction were recorded at the far downwind tower<br />

(100 m). The horizontal flux of PM 10 particles was calculated for each of the three<br />

towers. No measurable differences were found between the horizontal flux at the road<br />

and the flux 100 m downwind. This indicated that the removal of PM 10 by deposition<br />

was not occurring to an appreciable extent. Modeling studies based on individual particle<br />

sizes confirmed this finding and further suggested that even at 1,000 m downwind, the<br />

fraction of particles removed is less than 25% at high wind speeds (u * > 0.5 m/s). At<br />

lower, more common wind speeds (u * < 0.5 m/s), the fraction of particles removed at<br />

1,000 m was less than 20% for neutral conditions and less than 10% for very unstable<br />

conditions.<br />

Vehicle generated fugitive dust transport was examined in a field study at the<br />

Dugway Proving Ground, UT. A 10 by 12 array of 2.5 m high by 2.4 m wide by 12.2 m<br />

long cargo shipping containers was used to simulate urban terrain. The containers were<br />

arranged with their widest dimension against the prevailing winds and were spaced 14<br />

meters apart. An unpaved road upwind of the container array was traversed with a<br />

vehicle under nighttime stable atmospheric conditions. Two tower arrays, one near the<br />

road and one 95 m downwind were instrumented with DustTrak PM 10 monitors.<br />

Calculation of the horizontal dust flux using a variety of mathematical fits to the wind<br />

and concentration profiles indicated that by the time the plume from the road reached the<br />

95 m tower, approximately 85% of the PM 10 had deposited.<br />

The two field tests described span a large range of conditions with regard to<br />

particle removal. For the Ft. Bliss tests, dispersion of the fugitive dust plume was<br />

efficient because of the neutral to unstable atmosphere. In contrast, the stable atmosphere<br />

during the Dugway tests retarded near-source dispersion, keeping the dust plume close to<br />

7-1


the ground, where particles have a better chance to deposit. In addition, the areas of<br />

stagnation on both the windward and leeward sides of the containers may have enhanced<br />

the removal rate of PM 10 particles.<br />

The conditions during the Ft. Bliss test may be more true to everyday reality than<br />

the Dugway tests. This assertion is supported by two observations. 1.) Unpaved road<br />

dust is more likely to be an air quality concern in arid, open-range regions characterized<br />

by much of the American Southwest, and 2.) It is likely that the majority of travel on<br />

unpaved roads occurs during the day when the atmosphere is either neutral or unstable.<br />

With these observations in mind, it seems that in areas were unpaved road dust is of<br />

greatest concern, the removal of particles by deposition is smallest in magnitude. This<br />

goes against a recent movement to divide fugitive dust emissions by a factor of 2-4 in<br />

order to account for deposition within several hundred meters of the road. We note<br />

however that the conditions at Ft. Bliss are not representative of every area in the United<br />

States that is concerned with road dust. There may be some instances where there is<br />

thick vegetative cover very close to an unpaved road (e.g. farms). This possibility is<br />

revisited later in the text.<br />

Second, for the purposes of modeling the transport and removal of dust particles<br />

near an unpaved road source, the Gaussian-style models, such as EPA’s ISC3, provide an<br />

imperfect, but reasonable preliminary approach. Dispersion and deposition are the two<br />

main processes that must be accounted for in any model used. The ISC model was<br />

compared to the one-dimensional Atmospheric Diffusion Equation (ADE) and to data<br />

obtained at Ft. Bliss. The ADE relies on similarity theory for the parameterization of<br />

dispersion. Based on vertical concentration profiles measured as part of the Ft. Bliss<br />

experiments, it was determined that ADE over predicts dispersion in the region very close<br />

to the source (i.e. 500 m downwind or less) and that the ADE is not suited for modeling<br />

over short downwind distances. The ISC model uses a distance-based dispersion<br />

parameter that depends on atmospheric stability (σ). Comparison of predicted<br />

concentration profiles with those observed at Ft. Bliss indicated that the ISC captures the<br />

approximate shape of the dust plume. However, whereas the measured profiles varied in<br />

steepness at different values of the friction velocity u * , the ISC profile was invariant.<br />

This is because the friction velocity is not included in the parameterization of σ.<br />

Several different models for dry deposition were compared. They produced<br />

deposition velocities that were in agreement within a factor of two or so. The<br />

formulation for deposition velocity in the ISC model has the advantage that it only<br />

requires knowledge of three parameters: The friction velocity u * , the roughness height z 0 ,<br />

and the Monin-Obukhov length L. U * and L may be estimated from commonly-measured<br />

meteorological parameters. Z 0 can be inferred from existing land use databases (e.g.<br />

EPA’s BELD database). In contrast, more complex formulations, including the widely<br />

cited Slinn (1982) model, require many parameters as input, some of which are difficult<br />

to obtain.<br />

It is difficult to determine how well models for deposition velocity reflect realworld<br />

values. This is largely because these models have not been widely-tested with<br />

field data. In most cases, adjustable parameters are made to fit Chamberlain’s (1967)<br />

wind tunnel tests for deposition on grass surfaces. As part of this study, the deposition of<br />

particles to two different types of surrogate surfaces was measured. Comparison with the<br />

7-2


predictions of Slinn’s (1982) model and the one used in the ISC3 showed that both<br />

models tend to under predict the measured deposition by at least a factor of ten. This<br />

discrepancy was attributed, partly, to the difficulty of relating measurable field deposition<br />

velocities to parameterized model results. The accurate measurement and modeling of<br />

particle deposition velocities continues to pose a challenge to scientists, even after four<br />

decades of research. For the purposes of this report, it is sufficient to note that there may<br />

be substantial (factor of two or more) uncertainties in the specification of deposition<br />

velocity for use in a model.<br />

Despite the previously-noted shortcomings, both in the deposition and the<br />

dispersion portions of the ISC model (they are not unique to the ISC), it predicted the<br />

results for the Ft. Bliss tests fairly well. Specifically, the ISC predicted that at 100 m<br />

downwind of the road, there are no measurable reductions in the horizontal flux of dust<br />

particles in the size range covered by PM 10 (Aerodynamic diameter of 10 microns or<br />

less). However, the model suggested that for larger particles, D p =19.7 µm, the fraction<br />

remaining in suspension would be reduced to about 90% over the same distance.<br />

The ISC did not perform as well for the tests at the Dugway Proving Ground.<br />

While the model indicated that the fraction of PM 10 remaining in suspension 95 m<br />

downwind was 70%, the measured values were closer to 15%. It was hypothesized that<br />

the disagreement between the model and the measurements was due to the unusual terrain<br />

consisting of large containers. Perhaps deposition in the stagnation zones around those<br />

containers is not adequately represented in the modeled deposition velocity.<br />

Alternatively, it is possible that the channeling of particles between the containers under<br />

stable conditions is not adequately represented by the ISC dispersion parameter.<br />

The similarity between the concentration profiles measured at Ft. Bliss and those<br />

predicted by the ISC is encouraging. The further agreement between the modeled and<br />

measured values of near-field particle removal support the use of the ISC. The same<br />

cannot be said for the experiments at the Dugway Proving Ground. Nevertheless,<br />

considering the uncertainties associated with deposition velocities and the paucity of field<br />

data, we cannot expect any model to provide better than an approximate estimate of the<br />

transportable fraction. Therefore, in the foreseeable future, the ISC provides a<br />

reasonable, though imperfect approach. An additional consideration is that the ISC has<br />

been standardized by the US EPA, making it facile to disseminate to the air quality<br />

community if so needed.<br />

Third, the Box Model approach is elegant in its simplicity, but the assumptions<br />

behind the model, specifically those relating to dispersion, are probably not valid for<br />

unpaved road dust emissions. Gillette (Countess, 2001) proposed a Box Model to<br />

estimate the transportable fraction of unpaved road dust PM 10 emissions. In the model,<br />

dust particles deposit at the ground and disperse out of the top of the box at a rate that is<br />

proportional to the concentration in the box (assumed to be uniform with height). For<br />

deposition, this is a reasonable assumption, especially since the deposition velocity is not<br />

too variable with height for neutral to unstable conditions (i.e. drawing on the resistance<br />

analogy, r a


particles through the top of the box musts always be in the upwards direction. These<br />

conditions are closely met during a dust storm when the ground is a strong source, and<br />

the vertical concentration profile does not vary over a large distance in the horizontal.<br />

They are only met under neutral to unstable conditions for the case of emissions from<br />

unpaved roads, and even then, only over very short downwind distances.<br />

Fourth, the height of the wake generated behind a moving vehicle has an effect on<br />

the fraction of particles removed as the plume travels downwind, though this effect is<br />

only significant for stable conditions. Preliminary measurements suggested that the<br />

vertical extent of the turbulence induced in the lee of a passing vehicle scales<br />

approximately linearly with vehicle height. According to the ISC model, a larger fraction<br />

of particles is removed as the dust plume travels downwind for small values of the<br />

“injection height” than for larger values. This effect is more pronounced under stable<br />

conditions when the dust plume is not rapidly expanded in the vertical direction. Overall,<br />

in view of the uncertainties in quantifying deposition velocities, the “injection height” is<br />

probably a secondary consideration.<br />

Fifth, measurement of emission factors for seven different vehicles with gross<br />

weights varying from 1,622 kg to 23,636 kg were at odds with the silt-based emission<br />

factors suggested in AP-42 (USEPA). The Ft. Bliss measurements indicated that<br />

emission factors were highly dependent on vehicle speed. Linear regression of emission<br />

factor vs. speed for each of the seven vehicles resulted in R 2 values ranging from 0.37 to<br />

0.95, with six of the seven vehicles giving values greater than 0.70. In contrast, the AP-<br />

42 emission factor formulation does not consider vehicle speed. As a result, the AP-42<br />

tends to over predict PM 10 dust emitted from unpaved roads when the vehicle speed is<br />

less than about 20 to 25 mph. In addition, the use of an “average” vehicle weight in the<br />

AP-42 equation tends to introduce a positive bias for small vehicles. According to the<br />

measurements at Ft. Bliss, for passenger-sized vehicles (large, heavy trucks excluded),<br />

the AP-42 appears to overestimate emissions for vehicle speeds under 35 mph. Emission<br />

inventories often assume an average unpaved road travel speed of 20 or 25 mph. Under<br />

those conditions, use of AP-42 would result in an overestimate of PM 10 emissions by<br />

50% to 100% based on the field study results. This discrepancy, if it can be shown to<br />

hold at locations other than Ft Bliss, would explain some of the incongruencies between<br />

PM 10 dust emission inventories and the amount of crustal material found on filter samples<br />

(Watson and Chow, 2000).<br />

Sixth, prior to applying a correction to existing emissions inventories, it is<br />

important to verify if a substantial amount of PM 10 is removed near the source and if it is<br />

possible that there are other sources of error in the emissions estimates for unpaved road<br />

dust. The conclusions of this report are based, in part, on two field tests that represent the<br />

opposite extremes of atmospheric and land use conditions. In one case, the terrain<br />

consisted of small roughness elements and the atmosphere was unstable. In the other<br />

case, the roughness elements were building-sized and the atmosphere was stably<br />

stratified. It is important to compare model data, ISC or other models, to measurements<br />

performed under a range of conditions. Considering the uncertainties in deposition<br />

velocities and dispersion parameters, model predictions must be taken with a grain of salt<br />

until field data are available for comparison. The two field studies summarized in this<br />

report utilized highly time resolved particle monitors operating at multiple heights. This<br />

7-4


allowed for the comparison of downwind concentration profiles and horizontal fluxes of<br />

PM 10 dust with model predictions. A similar approach is recommended for future studies<br />

with the addition of a LIDAR instrument to quantify dispersion at long downwind<br />

distances.<br />

It is particularly important to obtain data for multiple types of terrains. The ISC<br />

model performed well for sparsely vegetated desert terrain; for a mock urban area, the<br />

model under estimated particle removal by a factor of 2.5. It is unclear from these results<br />

how well the ISC model performs when the downwind fetch is comprised of, for<br />

example, corn crops. Related to this, the models considered in this report only account<br />

for the deposition of particles after the emitted dust plume has passed the first several<br />

rows of vegetative cover. That is, we have not considered what happens at the interface<br />

between the unpaved road and the first row of vegetation encountered downwind of the<br />

road. Classic dispersion models are not suited for representing this “impact region”. The<br />

efficiency of particle removal in this region is unknown, largely due to a paucity of data.<br />

Preliminary work performed in this area (Raupach and Leys, 1999) indicates that the<br />

filtering of particles by windbreaks may be significant. This hypothesis should be<br />

examined experimentally.<br />

Futhermore, at the time of the writing of this report, it is still unclear if the<br />

discrepancy between PM 10 emission inventories and the fraction of PM 10 found in<br />

ambient samples can be attributed solely to the near field deposition of particles. There<br />

are a number of other possible sources of error that may result in inflated emission<br />

inventories (Watson and Chow, 2000; Countess, 2001). One finding of the present study<br />

is that the AP-42 tends to over predict emissions of PM 10 dust for smaller, passengersized<br />

vehicles, especially when the vehicles are traveling at speeds less than 35 mph.<br />

This may frequently be the case on small rural unpaved roads or driveways.<br />

Additionally, the accuracy of the activity levels that are applied to unpaved road dust<br />

inventories is a great source of uncertainty. There is a paucity of traffic counts and<br />

accurate vehicle speed estimates for unpaved roads. This shortage of data makes intuitive<br />

sense, since unpaved roads are unpaved, precisely because they are not used often enough<br />

to warrant attention from community planning associations.<br />

7.1 The findings of this study in relation to recent work<br />

Watson and Chow (2000) documented the discrepancy between emission<br />

inventories for PM 10 fugitive dust and the source attribution of ambient filter samples.<br />

Their analysis indicated that the amount of geologically-derived PM 10 found in the air is<br />

much smaller than would be expected based on the emission inventories and dispersion<br />

models. Countess (2001) summarized eleven shortcomings in the current treatment of<br />

fugitive dust emissions. Though not all eleven findings are directly relevant to the work<br />

in this report, we address several of them below.<br />

The removal of dust in the area near a source has been hypothesized to be<br />

substantial because concentrations of particles downwind of the source rapidly attenuate<br />

to the background. This is an erroneous deduction. Watson and Chow (2000) and<br />

Countess (2001) both cite an earlier study (Watson et al., 1996) where the concentration<br />

of PM 10 dust was found to decrease by 90% only 50 m downwind from the source.<br />

While this may be true, it is incorrect to assume that the decrease is due solely to<br />

7-5


deposition. Much of the decrease in concentration - indeed for the conditions at Ft. Bliss,<br />

most of the decrease - is due to dispersion in the vertical direction. Therefore, using the<br />

magnitude of the PM 10 concentration downwind of a dust source to estimate the removal<br />

of particles will suggest a much higher deposition rate than is achieved in reality. This<br />

does not invalidate the correct assumption that some of the PM 10 emitted will be removed<br />

as the dust plume travels downwind of the source. However, the removal occurs to a<br />

smaller degree and over larger distances than previously speculated, especially for the<br />

open, sparsely vegetated terrain that much of the American southwest is comprised of.<br />

The results of this study support the first finding put forth by Countess (2001),<br />

“Not all suspendable particles are transported long distances”. However, the Ft. Bliss and<br />

Dugway studies show a large range of near-field removal rates. In the rangeland setting<br />

of Ft. Bliss, little removal of suspended dust was observed at a downwind distance of 100<br />

m. This result was supported with a modeling effort that used a Gaussian approach<br />

similar to the ISC model (USEPA, 1995). Nighttime experiments at the Mock Urban<br />

Setting at the Dugway Proving Grounds suggested that under stable atmospheric<br />

conditions and large roughness elements (cargo containers) 85% of the PM 10 may be<br />

removed by deposition 95 m downwind of the unpaved road. This enormous range of<br />

removal rates emphasizes that it is not appropriate to apply a single correction factor to<br />

all fugitive dust emissions as a means of accounting for near-field particle removal.<br />

Though not documented, the community of scientists and professionals have, in the last<br />

several years, been circulating the idea that if fugitive dust emissions were divided by a<br />

factor of four, then the discrepancy between emissions and ambient measurements of<br />

geological PM 10 would disappear. While it is possible that this is true on an average<br />

basis (i.e. over large spatial domains), it is unlikely that this factor of four is applicable to<br />

every combination of airshed, land use distribution, and atmospheric conditions. Each<br />

combination of setting and meteorological conditions should be considered separately in<br />

a modeling framework that makes use of the known physics of particle dispersion and<br />

deposition.<br />

Related to the last point, the removal of PM 10 dust particles in the “impact zone”<br />

immediately after emission, at the junction of the road edge and the first row of<br />

vegetation, should be examined more closely. The work presented in this report only<br />

considers deposition of particles from “above” the vegetative canopy. It does not<br />

consider that particles may be removed to an appreciable extent by a process similar to<br />

filtration when the dust plume is forced through vegetative elements. Cowherd and Pace<br />

(2002) and He et al. (2002) have initiated work in this area. Those authors suggest that<br />

the fraction of particles removed in the impact zone is proportional to the blockage of the<br />

wind by the vegetation. This is applied to canopies exhibiting up to 50% blockage. At<br />

higher blockages, the impinging dust plume is assumed to be diverted over the<br />

vegetation. This conceptual approach warrants pursuit. However, prior to being used to<br />

correct road dust emission inventories, it requires testing against real data and analysis in<br />

the context of particle physics and previous work (e.g. Raupach and Leys, 1999).<br />

As one of the recommendations from his second finding, “Regional-scale vertical<br />

flux is smaller than the local-scale fugitive dust flux”, Countess (2001) suggests testing<br />

the Gillette box model and modifying the model parameters if appropriate. In this study,<br />

the box model was examined, from a theoretical and practical standpoint. The analysis<br />

7-6


indicated that the model is not practical for wide spread use because the parameterization<br />

of dispersion is reliant on assumptions that cannot always be met in reality.<br />

Related to the third finding from Countess (2001), “Fugitive dust emission factors<br />

need to be appropriate”, the emission factors measured at Ft Bliss with high timeresolution<br />

instruments suggest that AP-42 may overestimate PM 10 emissions from<br />

unpaved roads. This overestimate was due to the lack of accounting for vehicle speed<br />

and to the use of the average weight of all vehicles in calculating the emission factors.<br />

All in all, this report concludes that there is no single, “silver bullet” solution to<br />

the problem of reconciling fugitive dust emission inventories and ambient source<br />

contribution estimates. Progress has been achieved in the quantification of the near-field<br />

removal of dust particles and work in that area should continue. There are still<br />

uncertainties about the AP-42 emission factors for unpaved road dust that have resurfaced<br />

in this study and require resolution.<br />

7-7


7-8


8. REFERENCES<br />

Bache, D.H. (1981). Analyzing Particulate Deposition to Plant Canopies. Atmos Env.,<br />

13: 1881 – 1887.<br />

Biltoft, C., 2001. Customer Report for Mock Urban Setting Test, Dugway, UT,<br />

Meteorology & Obscurants Division, West Desert Test Center. Defense Threat<br />

Reduction Agency, Alexandria, VA.<br />

Byun, D.W., and R. Dennis (1995). Design Artifacts in Eulerian Air Quality Models:<br />

Evaluation of the Effects of Layer Thickness and Vertical Profile Correction on<br />

Surface Ozone Concentrations. Atmos Env. 29: 105-126.<br />

CEPA, 1998a. AP-42, Compilation of Air Pollutant Emission Factors - Vol 1. Stationary,<br />

Point, and Area Sources, Chapter 13.2.2 Unpaved Roads, Washington, D.C., U.S.<br />

Environmental Protection Agency.<br />

Chamberlain, A.C. (1967). Tranport of Lycopedium Spores and Other Small Particles to<br />

Rough Surfaces. Proc. Roy. Soc., 208A: 45-70.<br />

Countess, R. (2001). Methodology for Estimating Fugitive Windblown and Mechanically<br />

Resuspended Road Dust Emissions Applicable for Regional Scale Air Quality<br />

Modeling. Prepared for the Western Governors Association by Countess<br />

Environmental, Westlake Village, CA, April, 2001.<br />

Cowherd, C. (1999). Profiling data for open fugitive dust sources. Prepared for U.S.<br />

Environemntal Protection Agency, Emissions Factor and Inventory Group, Office<br />

of Air Quality Planning and Standards, RTP, NC, by Midwest Research Institute,<br />

Kansas City, MO.<br />

Cowherd, C., and T. Pace (2002). Potential role of vegetative ground cover in the<br />

removal of airborne fugitive dust. Presented at the Annual meeting of the Air and<br />

Waste Management Association, Baltimore, MD, June 2002.<br />

Dabberdt, W.F., Lenschow, D.H., Horst, T.W., Zimmerman, P.R., Oncley, S.P., and A.C.<br />

Delany (1993). Atmosphere-Surface Exchange Measurements. Science, 260:<br />

1472-1481.<br />

Dai, W., C.I. Davidson, V. Etyemezian, and M. Zufall (2001). Wind Tunnel Studies of<br />

Particle Transport and Deposition in Turbulent Boundary Flows. Aerosol Science<br />

& Technology, 35:887-898.<br />

Du, S., and A. Venkatram (1997). A Parametrization of Verical Dispersion of Ground-<br />

Level Releases. J. App. Meteorology, 36: 1004-1015.<br />

Etyemezian, V., C.I. Davidson, S. Finger, M. Striegel, N. Barabas, and J. Chow (1997).<br />

Vertical Gradients of Pollutant Concentrations and Deposition Fluxes on a Tall<br />

Limestone Building. J. Am. Inst. Conservation, 37: 187-210.<br />

Etyemezian V., H. Kuhns, J. Gillies, M. Green, M. Pitchford, and J. Watson (2002).<br />

Vehicle based road dust emissions measurements (I): Methods and Calibration.<br />

(submitted).<br />

8-1


Gifford, F.A. (1961). Use of Routine Meteorological Observations for Estimating<br />

Atmospheric Dispersion. Nucl. Safety, 2: 47-51.<br />

Gillette, D. A. (1974). On the production of soil wind erosion aerosols having the<br />

potential for long range transport., J. de Recherches Atmospheriques, 8: 735-744.<br />

Gillies, J.A.; Watson, J.G.; Rogers, C.F.; Dubois, D.; Chow, J.C.; Langston, R.; and<br />

Sweet, J. (1999). Long term efficiencies of dust suppressants to reduce PM 10<br />

emissions from unpaved roads. JAWMA, 49:3-16.<br />

Gillies, J.A., N. Lancaster, W.G. Nickling, and D. Crawley (2000). Field determination<br />

of drag forces and shear stress partitioning effects for a desert shrub (Sarcobatus<br />

vermiculatus, Greasewood). Journal of Geophysical Research, Atmospheres 105<br />

(D20): 24871-24880.<br />

Goossens, D., 1985. The granulometrical characteristics of a slowly-moving dust cloud,<br />

Earth Surface Processes and Landforms, 10:353-362.<br />

He, S., J. Ching, D. Gillette, W. Benjey, T. Pace, and T. Pierce (2002). 2002 AAAR<br />

annual meeting.<br />

Hider, Z.E., Hibberd, S., and C.J. Baker (1997). Modelling Particulate Dispersion in the<br />

Wake of a Vehicle. Journal of Wind Engineering Vol 67&68 :733-744.<br />

Hosker, R.P. Jr. 1984. Flow and Diffusion near Obstacles. In Atmospheric Science and<br />

Power Production Report for United States Department of Energy. Darryl<br />

Randerson, editor.<br />

Houser, C. A., and W.G. Nickling (2001). The Emission and Vertical Flux of Particulate<br />

Matter


Report prepared for the Great Basin Unified Air Pollution Control District, April<br />

24, 1997.<br />

Peters, K., and R. Eiden (1992). Modelling the Dry Deposition Velocity of Aerosol<br />

Particles to a Spruce Forest. Atmos Env, 26A: 2555-2564.<br />

Pleim, J., A. Venkatram, and R. Yamartino (1984). ADOM/TADAP Model Development<br />

Program. Volume 4. The Dry Deposition Module. Ontario Ministry of the<br />

Environment, Rexdale, Ontario.<br />

Porch, W. and D. A. Gillette (1977). A Comparison of Aerosol and Momentum Mixing<br />

in Dust Storms Using Fast-Response Instruments. J. Appl. Meteor., 16:<br />

1273-1281, 1977.<br />

Rao, K.S., Gunter, R.L., White, J.R., and R.P. Hosker. Turbulence and Dispersion<br />

Modeling near Highways. Atmos Env. 36: 4337-4346.<br />

Raupach, M.R., and F.L. Leys (1999). The Efficacy of Vegetation in Limiting Spray Drift<br />

and Dust Movement. Report Prepared for the Department of Land and Water<br />

Conservation, Gunnedah, Australia by CSIRO, Canberra, Australia.<br />

Schopflocher, T. P., and P.J. Sullivan (2002). A Mixture Model for the PDF of a<br />

Diffusing Scalar in a Turbulent Flow. Atmos Env., 36: 4405-4417.<br />

Sehmel, G.A. (1973). Particle Resuspension from an Asphalt Road Caused by Car and<br />

Truck Traffic. Atm. Env., 7: 291-309.<br />

Seinfeld, J.H., and S.N. Pandis (1998). Atmospheric Chemistry and Physics: From Air<br />

Pollution to Climate Change. Wiley Interscience. New York, USA.<br />

Seshadri, G., 2002. Fugitive Dust Transport from Unpaved Roads in Arid Climates. MS<br />

Thesis, Department of Chemical and Fuels Engineering, University of Utah, Salt<br />

Lake City, Utah.<br />

Shir, C.C. (1973). A preliminary Numerical Study of Atmospheric Turbulent Flows in the<br />

Idealized Planetary Boundary Layer. J. Atmos Sci., 30: 1327-1339.<br />

Slinn, S.A., and W.G.N. Slinn (1980). Prediction for Particle Deposition on Natural<br />

Waters. Atmos Env., 14: 1013-1016.<br />

Slinn, W. G. N. (1982). Predictions for particle depositions to vegetative canopies,<br />

Atmos. Env. 16: 1785-1794.<br />

Turner, D.B. (1970). Workbook of Atmospheric Dispersion Estimates. PHS Publication<br />

No 999-AP-26. U.S. Department of Helath, Education and Welfare, National Air<br />

Pollution Control Administration, Cincinnati, OH.<br />

USEPA, (1995). User’s Guide for the Industrial Source Complex (ISC3) Dispersion<br />

Models Vollume II: Description of Model Algorithms. US Environmental<br />

Protection Agency, Office of Air Quality Planning and Standards: Emissions,<br />

Monitoring, and Analysis Division, RTP, NC.<br />

USEPA (1999). Compilation of Air Pollutant Emission Factors – Vol I, Stationary point<br />

and area sources. Report N0 AP-42, 5 th ed. Prepared by US Environmental<br />

Protection Agency, Research Triangle Park, NC.<br />

8-3


Venkatram, A. (1993). The Parametrization of the Vertical Dispersion of A Scalar in the<br />

Atmospheric Boundary Layer. Atmos Env., 27A: 1963-1966.<br />

Watson, J.G., J.C. Chow, J.A. Gillies, H. Moosmuller, C.F. Rogers, D. Dubois, and J.C.<br />

Derby (1996). Effectiveness demonstration of fugitive dust control methods for<br />

public unpaved roads and unpaved shoulders on paved roads. Report No. 685-<br />

5200.1F. Prepared for San Joaquin Valley Unified Air Pollution Control District,<br />

Fresno, CA, by Desert Research Institute<br />

Watson, J.G., and J. Chow (2000). Reconciling Urban Fugitive Dust Emissions<br />

Inventory and Ambient Source Contribution Estimates: Summary of Current<br />

Knowledge and Needed Research. DRI Document No. 6110.4F. Prepared for the<br />

U.S. Environmental Protection Agency, by Desert Research Institute, Reno, NV.<br />

May, 2000.<br />

Wu, Y.L., C.I. Davidson, D.A. Dolske, and S.I. Sherwood. 1992. Dry Deposition of<br />

Atmospheric Contaminants: The Relative Importance of Aerodynamic, Boundary<br />

Layer, and Surface Resistances. Aerosol Science and Technology 16: 65-81.<br />

Yee, E. and Biltoft, C. A., 2002. Concentration fluctuation measurements in a plume<br />

dispersing through a large rectangular array of obstacles, Boundary Layer<br />

Meteorology In Press.<br />

8-4

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!