You are on page 1of 11

Pergamon

PII: SOO21-8502(97)00037-2

J. Aerosol SCI. Vol. 29, No. l/2, pp. 129-139, 1998


p 1998 Elsewr Science Ltd. All rights reserved
Printed in Great Brmm
0021-8502/98 $19.00 + 0.00

AN EXPERIMENTAL
STUDY OF THE DRY DEPOSITION
MECHANISM FOR AIRBORNE DUST
Sergey Biryukov
Ben-Gurion

National Solar Energy Center, The Jacob Blaustein Institute for Desert
Ben-Gurion
University of the Negev, Sede Boqer Campus, 84990, Israel
(First received 12 March

1996; and infinalform

Research,

29 May 1997)

Abstract-A
study of dry deposition of the coarse mode of airborne dust on mirror surfaces has been
performed at Sede Boqer, Israel, as a part of a research program to protect the reflective surfaces of
solar concentrating
collectors. In particular,
the angular dependence
of the deposition
rate was
measured in order to separate between its diffusive (isotropic) and vertical components.
Dust was
collected on a flat sampler with changeable
slope, mounted on a wind vane. No deviation from
a cosine-type
angular dependence
was found for dust particles larger than 4pm. This result is
interpreted as domination,
in the vicinity of the collecting surface, of vertically directed particle flux
over diffusive components.
This conclusion was compared to the results of existing dry-deposition
models. The size distribution
of particles, collected on the flat horizontal surface, was converted to
a volume-mass
concentration
using calculated
sedimentation
velocity as the only component
of
deposition
velocity. The resulting value was confirmed
by our control measurements
of the
volume-mass
concentration.
Our results are also compared
with recent data on deposition
velocities, published by Lin et al. (1994). 0 1998 Elsevier Science Ltd. All rights reserved

1.

INTRODUCTION

Interest in the investigation


of atmospheric
dust deposition on surfaces arises in numerous
branches of science and industry. Theoretical
and experimental
results of such investigations are, e.g. summarized
in the reviews of Pye (1989) and Williams and Loyalka (1991).
Applied solar energy also turns out to be one of the fields where such knowledge would be
useful. The reason for this is the fact that contamination,
by airborne dust, of the reflecting
surfaces of concentrating
solar collectors may lead to considerable
revenue losses for solar
power stations.
Regular washing with water is the usual method employed for the reduction
of such
power losses. But for desert and semi-arid regions, which are the most attractive locations
for solar energy production,
this way of cleaning may not be feasible and, in any case, can
become very expensive.
For this reason we have developed
a program
to study the
possibility of suppressing the deposition of dust on the collectors of solar power stations. As
an alternative
to washing we have examined various approaches
to surface dust cleaning/
protection which are based on repulsive physical interactions
between the surface and the
dust particle.
The degradation
rate r = Aq/r] of optical efJiciency y of a solar collector due to accumulation of particles with projected area diameter d, on its surface can be expressed on its initial
stage in the following way:

44) = Cl - $(d,)ldd,)S(d,) J(d,).

(1)

(See Hinds, 1982, for definition of the projected urea diameter, and Stine and Harrigan, 1985,
for optical efficiency of solar collectors.)
Here J(d,) = V(d,)n(d,)
is the flux of particles with diameter d,, where V(d,) is the
deposition velocity (see Williams and Loyalka,
1991, p. 327, for definition) and n(d,) the
numerical volume concentration
of particles, q(d,) S(d,) the reduction of mirror reflectivity
or of glass transmissivity
due to a single particle, where S(d,) is the particle projected area
and coefficient q(d,) depends on surface properties
(e.g. it is different for the first- or
129

130

Serpey Biryukov

second-surface
reflectors) and on physical properties of dust, and $(d,) the efficiency of
surface cleaning or protection
against dust accumulation.
Our research program obviously
requires a twofold starting point: determination
of
(a) the relevant particle-size distributions
and (b) the principal mechanisms for their deposition. The first is important
because their size is the main factor which determines
the
physical interaction
of particles. The second, because knowledge of the actual deposition
mechanisms
provides-via
deposition
velocity-a
quantitative
basis for measuring
the
efficiency of any resulting cleaning or surface protecting techniques. This knowledge also
results in a relationship
between data on surface deposition
rates and the corresponding
particle-volume
concentrations.
In some cases, moreover, it renders possible the use of
a reference sample in order to normalize other surface data. At the practical level, of course.
such knowledge constitutes the required basis for designing solar collectors with better antisoiling properties.
The atmospheric
dust accumulation
on solar collectors has previously been studied by
Sayigh et al. (1987), Guddihy (1988) and El-Nashar
(1994). These works present results
which are integrated over long time periods and which include a wide spread of meteorological conditions.
They provide no quantitative
information
about the relative contribution of particles with different sizes to surface contamination,
or about the mechanisms
of
dust deposition on the solar collector surfaces from the ambient atmosphere.
Quantitative
information
about the influence of artificially produced and deposited dust particles of
different sizes (optically determined
mean diameters
5, 10, 50, 60 and 80,~m) on the
performance
of photovoltaic
panels can be extracted from the laboratory
measurements
of
El-Shobokshy
and Hussein (1993).
In the present article we report data on the size distribution
of natural airborne particle
flux J to glass-coated
mirror samples for particles having projected area diameter larger
than 4pm. As previously
demonstrated
by our measurements
of specular refectivity
degradation,
accompanied
with microscopic studies (Biryukov et NI., 1994) particles in this
size range play the dominating
role in the kind of optical degradation
under investigation
here.
The presently reported size distributions
of deposition
rates, on glass and glass-coated
mirror samples of different orientations,
demonstrate
high stability of their shapes. They
peak between 10 and 20,um in the projected area diameter. This shape repeats itself under
similar meteorological
conditions,
which will be discussed later. This fact points to the
existence of a definite mechanism for particle accumulation
(at least for the coarse particle
region, of interest to us) and provides the possibility to study it using both long-term (days)
and short-term
(hours) measurements.
In the present study we have placed a high level of importance
on using natural rather
than simulated dust-deposition
conditions. On the other hand, the conditions had to be well
defined in order to permit a clear-cut theoretical interpretation
of the results. To these ends,
the measurements
reported here were all performed under conditions, corresponding
to the
dry-deposition
process. Common definitions of this process can be found in Twomey (1977)
and Brimblecombe
(1986). This ensured the absence of complicating
water condensation
effects. Furthermore,
we were also able to ensure that all particle deposition
data would
correspond to conditions of wind parallel to flat surfaces of known slope. In this manner we
could determine the angular dependence of particle fluxes to glass surfaces, for particles of
different sizes.
As will be shown below, one can use the measured angular dependence
in order to
estimate the relative contribution
of the isotropic (diffusive) and vertically directed components in the particle flux. One must, however, ensure the absence of particle deposition by
impact on the obstacle from the disturbed air stream. In order to eliminate this mechanism,
glass samples with a variety of fixed slopes relative to the horizontal
were kept parallel to
the wind direction by attachment
to a wind vane. In order to take into account effects of
resuspension,
measurements
were performed with the samples half-coated with an adhesive
layer. Particle-size
distributions
for the collected samples were then determined
using
a computerized
microscope.

An experlmental

study of the dry deposition

131

The common
predictions
of the model of dry deposition
from turbulent
flow under
gravity are compared with our measured angular dependence of particle flux. We conclude
from our results that, for the particle sizes of interest, gravitational
settling effectively
dominates over the totality of all diffusive mechanisms. That means that air motion parallel
to a Aat collecting plate is slow enough to preserve laminar boundary layer near its surface
(see the discussion of conditions for turbulent deposition from the flow parallel to surface in
Friedlander,
1977, p.115).
Using this conclusion together with our data on particle fluxes to horizontal surfaces we
are able to estimate the corresponding
volume densities of dust particles with sizes in the
range 4-64 pm. These results are compared with some direct control measurements
of our
own and with published
data for the Negev region. We also compare our results with
measurements
of dry-deposition
velocities, reported by Lin et al. (1994).

2.

2.1. Measurements

EXPERIMENTAL

of particle-size

distributions

TECHNIQUE

on a surface

As previously
stated, the particles of interest for our research have projected
area
diameters larger than 1 pm. This size is large enough to permit the use of optical microscopy
for analysis purposes. For the study of particle-size
distributions
on mirror surfaces we
developed a computerized
set up, based on the MonoZoomVideo Microscope System of
Cambridge Instruments Corp. The optical image of the microscope is viewed by a Javelin
Corp. JE 7442X B/W video camera, which has a sensitive CCD-array
of 756 x 512 pixels.
The output of the CCD-camera
was recorded by a DT-2855 Quick Capture frame
grabber of Data Trunslation Corp.
Image processing of grey-scale frames was performed using a 486-based micro-computer,
with the help of specially designed software for particle recognition.
The latter included
some standard image processing operations,
which mostly are presented in Russ (1990).
The grey-scale image was converted to a binary (black and white) one with the help of
brightness discrimination
before sizing (see Russ, 1990 for a discussion of the thresholding
operation). This procedure was used for background
subtraction.
Each particle was characterized by its projected area in pixels. The last one was recalculated
to the diameter of
equivalent
circle, i.e. the projected area diameter in pm, using the beforehand
measured
scale. Special corrections
had to be applied to the effective diameters of the projected
areas of irregular particles. In particular,
it was necessary to take into account the fact
that most particles are to some extent elongated and tend to settle on a surface so as to
have minimum
potential
energy. This phenomenon
makes the average projected area
diameter for an ensemble of particles larger than it would be for the same particles if
they were randomly
oriented.
Depending
on particle and surface properties
(electric
charge, etc.) this correction,
in the most common case, depends on particle size and, if
neglected, can result in substantial
errors when calculating mass distributions,
aerodynamic
diameters, etc.
Particles were grouped into predetermined
size ranges according to their projected area
diameters. All particles in each frame were counted and sorted into their size ranges. The
histogram
over size ranges was incremented
using successive frames until at least 10
particles were counted in each size range (Hinds, 1982).

2.2. Dust sampling


Exposure of the glass samples to dust was performed in the field in the following manner.
The sample holder (Fig. 1) consisted of a pair of light rectangular
plastic plates, 10 x 20 cm2
in size, with rectangular
hollows. These hollows were 2 or 3 mm deep and of different sizes
(2 x 5 cm2, 3 x 3 cm2, 3 x 6 cm2) in order to permit the flush-type mounting of 2-3 mm thick
samples of glass or glass-coated mirrors. The sample holder was mounted on a wind vane, in
order to maintain its surface parallel to the wind direction.

Sergey Biryukov

132

Fig. 1. The sampler

on wind vane with variable slope angle.


(3) sample.

(1) wind vane, (2) sample

holder,

For each exposure both holder plates, with two samples, were installed symmetrically
relative to the vertical wind vane, each of them having the same angle 0 to the horizontal, as
shown in Fig. 1.
After each exposure the homogeneity
of particle density on the sample surface was
checked and the number of particles was calculated according to their sizes, for several
random positions of each of the two samples (using equal areas of integration for each pair).
The measurement
precision for a single sample, mounted close to the vertical position, is
limited by imperfections
in the determination
of the wind vanes vertical position. Equation
(2) gives the value of the relative error E in cos8, caused by an absolute error + 6 in the
value of H for such a sampling geometry
c=

-(l-cos6ftanQsin6).

(2)

This expression shows that the measurement


error becomes infinite as 0 approaches 90,
thus, limiting the possibility of precise flux measurements
for samples with relatively sleep
slopes.
On the other hand, it is relatively easy to guarantee the rigidity of an isosceles triangle,
having a high degree of symmetry, relative to the vane plane, for pairs of samples. We
therefore used, for our data analysis, the sum of particle numbers
for two symmetric
samples. In this case the relative error g in cos0 becomes independent
of the slope angle
0 since equation (2) reduces to
c=cos6-1.
For reasonably

small values of 6 the resulting

2.3. Bounce-@blow-ofS

(3)
relative

error now becomes

negligible.

correction

Once particle has touched the surface of deposition, it may remain on it infinitely long, or
returned to air stream due to bounce-off (rebound), blow-off (reentrainment),
or due to
combined effect of these two processes.
In order to avoid experimental
errors in flux measurements,
caused by particle resuspension, the technique of coating the sample with an adhesive substance (greasing) is widely
used. But in optical measurements
the adhesive layer changes the optical properties of the
sample surface and may contribute to the error in particle diameter determination.
We used

An experimental

133

study of the dry deposition

a modification of this technique, in which we prepared samples, half-coated with an


adhesive layer. In this way about 50% of the sample surface area remains free of adhesive
coating. This technique gives us the possibility of measuring both total and effective fluxes
to the surface (J and Jeff, respectively) and of extracting from these measurements data
about the probability for particles of different sizes not to be resuspended after their contact
with the surface. It can be expressed as a ratio A = Jeff/J which we shall call below in the
text the coejicient of attachment to surface, or attachment coefficient.
Our adhesive layer was prepared from a solution of white Vaseline in xylol according to
recommendations in the manual for the SM16706 Konimeter of Sartorius-Membranjilter
GmbH. Its quality was checked in an experiment on the re-entrainment of loess particles
with d = 2&1OOpm from the surface. These measurements demonstrated zero probability
of reentrainment from the layer.
According to our measurements the value of the attachment coefficient was close to 0.9
without any meaningful dependence on particle size. These measurements are illustrated by
Fig. 2, where the average attachment coefficient over sizes 4-64pm equals 0.86 k 0.02.
Much efforts have been devoted to experimental and theoretical investigation of particle
resuspension, especially in connection with a problem of dust samplers efficiency (Vincent
and Gibson, 1981, Ingham and Yan, 1994). A review of models of particle reentrainment can
be found in the article of Ziskind et al. (1995). Pau U and Braaten (1995) gave a review of
rebound-reentrainment
theories and made numerical simulations of combined effect of
these processes. The last two authors conclude, that rebound is a dominant process in
limiting net deposition of particles from flows parallel to collecting surfaces.

2000

1500

1.5

z
3
k
2

I
3
1;

1000

%
.$
f

E
8
u
0.5

500

IL

16

32

64

Particle size, micrometer


+ A: glass surface
+

coefficient

Fig. 2. Measurement

B: glass with adhesive

coating

of attachment

of attachment

coefficient

on a sample,

half-coated

with a adhesive

layer.

Sergey Biryukov

134

Relatively small resuspension


from the non-greased
part of our samples can be explained
by low wind speeds during the reported measurements
(see Section 3.2 of this article).

3. EXPERIMENTAL

3.1. Frequenqs

distribution

DATA

qfparticle

AND

DISCUSSION

sizes on suyfaces

We started by measuring the size distributions


of dust particles which had accumulated
on the surfaces of solar collectors. The reported results are limited to the case of dry
deposition. This means that for all reported cases the samples were exposed to airborne dust
under ambient conditions
(of temperature
and relative humidity) that were relatively far
removed from the dew condensation
point. As a result, the weight of the individual particles
was conserved during deposition, not being changed by accumulated
water. In addition, the
collected dust was not washed out from the sample surface by dew and the picture of
randomly deposited particles on the surface was not disturbed by water drops.
The average wind speed for different data sets during all reported particle-size distribution measurements
was usually 2.5-3.5 m s- and did not exceed 5.8 m s- . The highest
registered average value for a lo-minute
interval during the same period was 8.5 m s- .
It was found that the typical size distribution
has its maximum at 15-20/~rn, its shape
being very stable from day to day and very similar for different glass-coated mirror surfaces
with various orientations.
This statement
relates to the average flux, corresponding
to
a relatively long exposure time: in our measurements
this varied between 4 and 100 h.
We illustrate the stability and similarity of particle-size distribution
shapes for particle
flux to surfaces of interest by Figs 3 and 4. Figure 3 shows the size distribution
of particle

16

32

64

Particle size, micrometer


Fig. 3. Solar-tracking

mirror:

particle

deposition

rates

An experimental

0.35

study of the dry deposition

_r

135

4
i
-

0.3

-t
-

-~

--t----+
j

-I

-_

32

16

64

-.

Particle size, micrometer


1 horizontal sample
Fig. 4. Deposition

rates horizontal

vertical sample

and close to vertical

samples.

Wind vane, h = 6 m

deposition rates measured on the mirrors of a tracking solar-thermal


system at Sede Boqer.
In Fig. 4 two size distributions
are shown. These correspond to samples, which were mounted
on a wind vane at a height 6 m from the ground, having inclinations 85 (close to vertical) and
0 (horizontal),
respectively. The size distributions
in Fig. 4 are normalized
to unit flux.
The stability of the particle-size distribution
shape is an important observation,
because it
indicates the stability of the principal deposition mechanisms and makes it possible to study
the latter over time scales of various numbers of hours.

3.2. Dry deposition controlled by gravity and diffusion: outdoor measurements


dependence of particle flux to surface

of the angular

A large variety of atmospheric


processes, united by the term dry deposition,
are
discussed in the literature (see, e.g. Pruppacher
et al., 1983). A few works discuss theoretical
models for the turbulent
deposition
of airborne
dust under gravity, e.g. Sehmel and
Hodgson (1978) Crump and Seinfeld (1981). These considerations
lead qualitatively
to
results, similar to those that can be derived from considerations
of particle transport,
by
means of gravity and diffusion, through a thin layer of stationary
gas above a flat plate.
Such considerations
were applied by Friedlander
(1977) to the case of a horizontal plate in
the presence of Brownian diffusion.
Two main outdoor mechanisms exist for the deposition of dust on a surface, in the case of
particles larger than 1 ,nm: settling in viscous air under gravity (sedimentation)
and inertial
impact from eddies, (eddy diffusion). See, e.g. Williams and Loyalka (1991).

Sergey Biryukok

136

In order to derive an expression for the deposition


velocity I/ via turbulent flow under
gravity one may consider a stationary gas above a flat plate with slope 0 to the horizontal
and particles that are transported,
by means of gravity (V, cos 0) and diffusion, through
a thin layer of the gas. For the particle sizes in our range of interest Brownian diffusion is
negligible and we can describe by a diffusion coefficient D other diffusive mechanisms
(in
our case, turbulent
diffusion).
Far enough from the leading edge of a thin flat plate 2 (Fig. 1) parallel to the wind, for
a flush-type flat sample 3 (Fig. 1) in the middle of it which is small relative to the plate, and
for reasonably low wind speed, which also enters the expressions for Reynolds number and
for the thickness of the laminar boundary
layer (Friedlander,
1977), there is practically no
dependence
of deposition
rate on the coordinate
in the wind direction.
Our control
measurements
of dependence of the numerical surface density of deposited particles on the
coordinate in the wind direction confirmed this assumption.
The one-dimensional
approximation can be used for calculation of particle flux to the sample surface in this case and the
linear equation for flux to the surface is
J = D$

+ V,cosCln.

(4)

Here V, is the sedimentation


velocity, D the turbulent
diffusivity. and n the particle
concentration.
The z-axis is normal to the surface.
Let n = N at the edge of the gas layer (at z = h) and at z = 0 let n = 0.
The flux J is obtained from the integral of equation (4) under these boundary
conditions as
J=

vs cos (IN
1 - exp(-

V,cos0h/D)

If Vs cos Gh/D >>1 then J z Vs cos UN, or I 2 I/, cos 0 (a situation in which gravity controls
deposition, the diffusion term being negligible). This gives similar particle-size distributions
on a surface for different slopes and zero flux (deposition
rate) on a vertical surface.
If, on the other hand, Vscos Bh/D<< 1 then J k DNjh, or V c D/h (gravity being now
negligible). In this case the deposition velocity I/ and deposition rate do not depend on the
surface slope.
We performed outdoor measurements
in order to estimate the relative contribution
of
sedimentation
and diffusion processes in the deposition rates for particles of different sizes.
The measurements
were performed at Sede Boqer, in the Negev highlands.
We measured the value of Je (flux on the inclined glass sample, kept parallel to the wind
direction) relative to J (flux on a horizontal sample) using the previously described sampling
technique
and particle sizing by means of computerized
optical microscopy.
We also
determined
the coefficient of attachment
for particles of different sizes d, in the manner
described above. Our measurements
with different slopes (for (1 = 45 and 60) revealed
a dependence of the form J,(d)/ J (II) = cos (1. Even for 0 > 80 we did not find a meaningful
deviation from this cos 0 dependence.
The results of these experiments
are shown in Fig. 5. Experiments
on horizontal
but
down-facing samples at heights of 2 and 6 m indicated a flux at least an order of magnitude
lower than on vertical samples.
The average wind speed for different data sets during the measurements
with variable
sample inclinations
never exceeded 3.6 m s- . During measurements
the highest individual
value of this parameter approached
4.9 m s -I (10 min average). This range of wind speeds
applies also to the measured values of the attachment
coefficient.
3.3. Discussion

of the model,fov

We can summarize

the main

dry deposition
properties

(1) the particle-size distributions


deposition
and moderate average

qf ambient dust

of the measured

deposition

rates as follows:

demonstrate
a stability of shape under conditions
wind speeds of O-6 m s- ;

of dry

An experimental

137

study of the dry deposition

JW 1J(c) vse
for particles less than 8 micron

for partides of all sizes

1.1

1.1
1

0.9

0.9

0.8

0.8

0.7

0.7

0.6

0.6

0.5

0.5

0.4

0.4

03

0.3

0.2

02

0.1

0.1
0

0
0

0 10 20 30 40 50 60 70 -80 90

10 20 30 40 50 60 70 80 90

The slope angle, 8 (degrees)

The slope angle, 8 (degrees)

for particles larger than 16 micron

for pat-tides between 8 and 16 micron


1.1

1.1
1

0.9

0.9

0.8

0.8

9.7

0.7

0.6

0.6

OS

0.5

0.4

0.4

0.3
0.2 -

0.1
0

0.3
0.2

\\

2.
0 10 20 30 40 50 60 70 80 90
The slope angle, 0 (degrees)

Fig. 5. The measured

deposition

\\
\

0.1
o-

10 20 30 40 so 60 70 80 90

The slope angle, 8 (degrees)

rates compared

with the cos O-dependence.

(2) the shape of the particle-size distribution does not depend on the surface slope angle
for particle sizes in the range 4-64 pm;
(3) the deposition rate is proportional to cos6 for 0 varying from 0 to at least 85.

According to our discussion of the relative role of gravitational and diffusive types of
deposition we conclude that the former mechanism dominates for particles with sizes
4-64pm under conditions of dry deposition and moderate wind speed (less than 4 m s- ), at
least for quantities, averaged over periods of several hours.
This being the case, we can make an attempt to derive data about volume dust
concentration versus particle size from data on the deposition rate on horizontal surfaces:

J,(d)

44 = I/,o.
The sedimentation

velocity can be calculated as


(7)

where p is the particle density, g the acceleration due to gravity, p the airs dynamic
viscosity, C the Cunningham correction factor (x 1 for d > 1 pm), and K the dynamic shape
factor (see: Williams and Loyalka, 1991, p. 52-66).
K depends on the nonsphericity of the particles and on their surface area. We estimate
this factor, to be 1.0 < K < 1.2 for our case.
.

Sergey Biryukov

138

16

Particle

+ mass concentration
+ deposition
Fig. 6. Mass concentrations

64

32

size, micrometer
(derived)

rates (measured)

of airborne

dust, derived

from deposition

rates

Now, the measured deposition


rate on a horizontal
surface, J?(d), may differ from the
geometrical flux through that surface by some factor A(d), which characterizes the efficiency of particle attachment to the surface. Therefore, the deposition rate due to sedimentation,
J,(d), in equation (6), must be calculated as J:(d)/A(d),
the measured value of A(d) for the
conditions
of our experiment
being equal to 0.86 k 0.02.
Figure 6 shows the mass concentration of dust in my mp3 calculated from experimental
data on deposition rates for a horizontal sample. The initial size distribution
of deposition
rates is plotted on the same graph for reference (in particles/cm2 s , the right scale). The
value p = 2.6 g cm 3 (density of dry loess) was used in this calculation. The integrated value
of mass concentration,
0.3 mgmd3,
agrees, within experimental
error, with our control
measurements
of volume mass concentration
by direct weighing of filters from a dustcollector of the type SM 16711 (Sartorius Membraqjdter).
It also confirms typical values
reported by Offer and Goossens (1990) for the Negev region.
4.

CONCLUSIONS

We have developed
a rapid method for determining
the particle-size
distribution
on
surfaces, which we use for assaying the contamination
by windborne dust of solar collectors.
Using this method we have estimated the range of particle sizes which are the principal
cause of reflectivity losses.
We have performed experiments
to study the process of dry deposition
of dust under
ambient conditions. It was found that this process (which corresponds to clear summer days
and is consequently
of paramount
importance for solar energy production)
is dominated by
the mechanism
of sedimentation
under gravity, for particles larger than 4pm. This conclusion is qualitatively
supported by the previous results of Lin et al. (1994).
This conclusion appears to be of widespread applicability
since the measurements
of Lin
et czl., were performed in Chicago, Illinois, USA, under very different climatic conditions.
Our results also agree with the wind tunnel experiments and dry-deposition
velocity model
calculations
by Sehmel and Hodgson (1978), cited in Lin et al.
This conclusion
enables one to use surface data on particle-size distributions
in order
to derive the corresponding
volume-dust
concentrations
and deposition
velocities, the
calculated sedimentation
velocity being used as a reference in such measurements.

An experimental

study of the dry deposition

139

Using this technique, we derived values for the volume dust concentrations of particles of
various sizes from about 4 pm to about 60 pm, from our outdoor measurements of particle
size distributions on glass surfaces. Furthermore, the absolute values of such concentrations
and their dependence on particle-size agreed with previously published values.
We plan to use the technique, developed here, as a key instrument for experiments aimed
at protecting the surfaces of solar collectors from dust.
Acknowledgements-This
research was partially funded by the Israel Ministry of Science
Israel Ministry of Energy and Infrastructure.
The author is indebted to Prof. D. Faiman
to the practical importance
of this problem and for many valuable discussions. It is also
to express deep appreciation
to Prof. J. Vincent for his instructive recommendations
on
version of this paper.

and Arts and parts by the


for drawing his attention
the authors pleasant duty
preparation
of the revised

REFERENCES
Biryukov, S., Faiman, D. and Goldfeld, A. (1994) Characterization
of mirror reflectivity losses caused by dust
deposited under different meteorological
conditions.
In Proc. Int. Conf on Compuratioe Assessments of Solar
Power Technologies SOLCOM I (Edited by Roy, A., Ben Gurion University of the Negev, Beer-Sheva, Israel and
Grasse,W.,
Deutsche
Forschungsanstalt
fur Luft- und Raumfahrt,
e.V., Cologne, Germany.
pp. 373-382.
Jerusalem.
Brimblecombe,
P. (1986) Air Composition and Chemistry. Cambridge
University Press, New York.
Crump, J. G. and Seinfeld, J. H. (1981) Turbulent deposition and gravitational
sedimentation
of an aerosol in
a vessel of arbitrary
shape. J. Aerosol Sci. 12, 405415.
El-Nashar,
A. M. (1994) The effect of dust accumulation
on the performance
of evacuated
tube collectors.
Sol. Energy 1, 105-l 15.
El-Shobokshy,
M. S. and Hussein, F. M. (1993) Effect of dust with different physical properties on the performance
of photovoltaic
cells. Sol. Energy 51, 505-511.
Friedlander,
S. K. (1977) Smoke, Dust and Haze. Wiley, New York.
Guddihy,
E. F. (1988) Soiling mechanisms
and performance
of anti-soiling
coatings. In Particles on Surfaces:
Detection, Adhesion, and Removal, Vol. 1 (Edited by Mittal, K. L.), p. 91. Plenum Press, New York and London.
Hinds, W. C. (1982) Aerosol Technology. Properties, Behurior, and Measurement of Airborne Particles. Wiley,
New York.
Ingham, D. B. and Yan, B. (1994) Re-entrainment
of particles on the outer wall of a cylindrical blunt sampler.
J. Aerosol Sci. 25, 327-340.
Lin, J. J., Noll, K. E. and Holsen, T. M. (1994) Dry deposition velocities as a function of particle size in the ambient
atmosphere.
Aerosol Sci. Technol. 20, 239-252.
Offer, Z. I. and Goossens, D. (1990) Airborne dust in the Northern Negev desert.J. Arid Environ. 18, 1-19.
Paw U. K. T. and Braaten, D. A. (1995) New perspectives on rebound and reentrainment
processes. Aerosol Sci.
Technol. 23, 72279.
Pruppacher,
H. R., Semonin, R. G. and Slinn, W. G. (coordinators)
(1983) Precipitation
scavenging, dry deposition,
and resuspension,
In Vol: 1, 2. Proc. of the 4th Int. Conf Santa Monica, California, 29 November-3
December,
1982, Elsevier, New York.
Pye, K. (1989) Aeolian Dust and Dust Deposits. Academic Press, New York.
Reeks, M. W., Reed, J. and Hall, D. (1988) On the resuspension
of small particles by a turbulent flow. J. Phys. D:
Appl. Phys. 21, 574-589.
Russ, J. S. (1990) Computer-Assisted Microscopy. The Measurement and Analysis of Images. Plenum Press,
New York and London.
Sayigh, A. A. M., Al-Jandal, S. and Ahmed, H. (1987) Dust effect on solar flat surfaces devices in Kuwait. In 1985
Workshop on the Physics of Non-Conventional Energy Sources and Material Sciencefor Energy (Edited by Furlan,
G., Mancini, N. A., Sayigh, A. A. M. and Seraphin, B. 0.) p. 353. World Scientific, Singapore.
Sehmel, G. A. and Hodgson, W. H. (1978) DOE Report PNL-SA-6721.
Pacific Northwest Laboratory,
Richland,
Washington.
Stine, W. B. and Harrigan,
R. W. (1985) Solar Energy Fundamentals and Design. Wiley, Brisbane-TorontoSingapore.
Twomey, S. (1977) Atmospheric Aerosols. Elsevier, New York.
Vincent, J. H. and Gibson, H. (1981) Sampling errors in blunt dust samplers arising from external wall loss effects.
J. Aerosol Sci. 15, 703-712.
Williams, M. M. R. and Loyalka, S. K. (1991) Aerosol Science Theory and Practice. With Special Applications to the
Nuclear Industry. Pergamon
Press, New York.
Ziskind, G., Fichman,
M. and Gutfinger,
C. (1995) Resuspension
of particulates
from surfaces to turbulent
flows-review
and analysis. J. Aerosol Sci. 26, 613-644.

You might also like