Download as pdf or txt
Download as pdf or txt
You are on page 1of 31

Aerosol Science 32 (2001) 1}31

Kinetic models for particle resuspension in turbulent #ows:


theory and measurement
M.W. Reeks *, D. Hall
European Commision, Joint Research Centre, I-21020 Ispra (Va), Italy
Magnox Electric, Berkeley, Gloucestershire, GL13 9PB, UK
Received 22 March 2000; accepted 25 May 2000

Abstract

Measurements are reported of the short-term resuspension of nominal 10 and 20 lm alumina spheres and
graphite particles from a polished stainless-steel #at plate in fully developed turbulent channel #ow. Prelimi-
nary measurements were made of the normal and tangential forces holding the particles onto the surface.
Whilst both forces had a broad spread and were on average much reduced in value compared to that for
smooth contact, the average tangential force was typically of order 1/100th of the average normal adhesive
force, suggesting as has been reported previously that drag forces can play a more important role in
esuspending a particle than lift forces. The resuspension measurements are compared with predictions of the
RRH (1988) kinetic model based exclusively on lift/normal forces and those of a rock'n roll kinetic model that
involves the rocking of a particle about an asperity (the motion being dominated by the drag force). The RRH
model consistently under predicted the amount of resuspension in contrast to that of the rock'n roll model
which gave values much closer to the measured resuspension. As in the RRH model, the rock'n roll model
admits the possibility of removal of particles by resonant energy transfer. However the results indicate, at least
for the cases considered, that this contribution is generally small, in which case the resuspension rate constant
reduces to the form appropriate for a balance of moments due to adhesion and aerodynamic forces about the
surface asperities at the points of contact. Under these &quasi-static' conditions a simpler and more exact model
for resuspension can be constructed: however in practice for a Gaussian distribution of removal forces this gives
very similar results to the original rock'n roll model. The formula for the resuspension rate constant under
&quasi-static' conditions has similarities with the empirical formulae proposed by Wen and Kasper (1989) on the
kinetics of particle re-entrainment from surfaces. Journal of Aerosol Science, 20, 483}498, although there are
important di!erences in interpretation.  2000 Elsevier Science Ltd. All rights reserved.

Keywords: Resuspension; Adhesion; Turbulent #ows

* Correspondence address: University of Bristol, Interface Analysis Centre, Oldbury House, 121 St. Michael's Hill,
Bristol BS2 8BS, UK. Tel.: #44-117-925-5666; fax: #33-117-925-5646.
E-mail address: mike.reeks@bristol.ac.uk (M. W. Reeks).

0021-8502/00/$ - see front matter  2000 Elsevier Science Ltd. All rights reserved.
PII: S 0 0 2 1 - 8 5 0 2 ( 0 0 ) 0 0 0 6 3 - X
2 M.W. Reeks, D. Hall / Aerosol Science 32 (2001) 1}31

1. Introduction

In their comprehensive state of the art review of particle resuspension, Ziskind, Fichman and
Gut"nger (ZFG) (1995) analysed "ve di!erent theoretical models for particle resuspension, point-
ing out their advantages and shortcomings. They categorised the models into two types: models
based on a force balance and models based on energy accumulation. Both types of model were
essentially statistical based on either a Monte-Carlo simulation or on formulae for the rate
constant for resuspension. The simplest force balance model assumes that once a threshold for
removal is exceeded, the rate of removal is determined by the frequency of the turbulent bursts
(characteristic of the random #uid motion close to the surface). This threshold is based on a balance
between the instantaneous aerodynamic lift and particle surface adhesive forces. No account is
taken of the possible distribution of both aerodynamic and adhesive forces. Attempts to improve
this situation were made by Braaten, Shaw and Paw (1990), who developed a particle resuspension
model based on a Monte-Carlo simulation. The model assumed that aerodynamic forces were
applied to the particle on the surface in discrete bursts. The value of these forces were derived from
some prescribed probability distribution, resuspension occurring when the forces exceeded the
local adhesive forces which were given a log-normal distribution.
ZFG also classed the kinetic model of Wen and Kasper (1989) as a force balance model though
the basic formula for the resuspension rate constant is analogous to the Arrhenius formula for the
desorption rate of molecules from a surface, i.e.
p"n exp (!q), (1)
where n is taken to be the maximum resuspension rate (the bursting frequency) and q the ratio of
the adhesive force to the instantaneous aerodynamic removal force. Wen and Kasper assumed that
the initial resuspension rate is given by the value of n appropriate for lightly bound particles and
derived expressions for the longer-term resuspension rate based on a distribution of the adhesive
forces which accounts for situations where q*1. In an alternative development of this approach,
Jurcik and Wang (1991) accounted for the in#uence of the distribution of removal forces by
de"ning a rate constant for a given adhesive force by multiplying the maximum resuspension rate
by the probability that the removal force exceeds the force of adhesion f i.e.



p"n P(F) dF, (2)
D
where P(F) is the probability density distribution for a removal force F.
The energy accumulation model analysed by ZFG was the model of Reeks, Reed and Hall
(RRH) (1988). This model deals exclusively with the in#uence of aerodynamic lift and surface
adhesion on resuspension: a particle is resuspended when it has accumulated enough vibrational
energy from the turbulent #ow to detach itself from the surface. The model allowed for the
possibility of resonant energy transfer when the forcing frequency of the lift force #uctuations is not
too di!erent from the natural frequency of vibration of the particle-surface deformation. Under
such circumstances a particle is removed much easier from a surface than applying the same lift
force quasi-statically (which is the basis of the simple force balance approach). It is a more general
model than the so-called force balance model because it admits the possibility of resonant energy
transfer. When this transfer is zero, then the formula for the resuspension rate constant reduces to the
M.W. Reeks, D. Hall / Aerosol Science 32 (2001) 1}31 3

form appropriate to a force balance. That is the motion of the particle/surface deformation is
approximated by a balance between the instantaneous aerodynamic lift force and the adhesive
force at each location/deformation in the adhesive surface potential well. This implies a functional
relationship between the particle deformation and the removal force at each instant of time. The
form for p in these circumstances is similar to the form suggested by Wen and Kasper (1989) though
the form for the exponent q is not related to the instantaneous value of the lift force but using the
equation of motion of the particle surface deformation is formally related to the ratio of the average
potential energy of the particle to the height of the adhesive potential barrier (hence the term energy
balance rather than force balance). This dependence on average quantities is re#ected also in the
value of n which is dependent upon the ratio of the rms of the #uctuating lift to that of its time
derivative. That is n is not an arbitrary quantity separate from the model itself but an intrinsic part of
the model and the motion of the particle-surface deformation.
The important feature of resuspension models (with or without resonant energy transfer) is that
the fraction of particles resuspended is greater than that based on a simple force balance approach:
by this we mean an approach whereby particles are said to resuspend when the mean removal force
is greater than the force of adhesion. (Note that the Wen}Kasper model implies that particles can
be removed when the instantaneous removal force is less than the adhesive force). A very important
conclusion of the ZFG analysis was that a comparison of estimates of the force of adhesion and the
mean and rms lift forces in a number of resuspension experiments, indicated that the lift force was
not su$cient to have caused the observed resuspension whether by resonant energy transfer or
simple force balance (even admitting the possibility that the mean lift force could be less than the
force of adhesion). ZFG suggested that rolling provided an alternative and more realistic mecha-
nism by which the particles could be detached from the surface. This conclusion is very much in line
with some earlier work by Wang (1990), who considered the e!ects of inceptive motion on particle
detachment from a surfaces and showed that the critical removal force depending on either a
force or moment balance could vary over three orders of magnitude for a given particle-surface
combination by simply varying the direction of the applied force. Wang concluded on the basis of
a simple calculation of drag in a linear velocity pro"le in a boundary layer, that incipient rolling
was the most dominant mechanism for particle removal. In addition ZFG by making plausible
estimates of the moments of the drag force and the adhesive force based on current theories also
showed that a balance of adhesive and drag moments was more consistent with the observed
resuspension in these experiments than a balance of adhesive and lift forces. In keeping with
Wang's analysis, Ziskind, Fichman and Gut"nger (1997) re"ned this approach by examining in
detail the moments arising from the contact of a smooth particle on a smooth surface as well as
a particle in contact with two and three asperities. As a result they obtained a condition for particle
detachment from a rough surface. Though this approach and that of Wang represented an
important step forward in identifying a more realistic mechanism for resuspension, the approach
ignores the essential statistical nature of the process. This feature was then added to the rolling
mechanism when ZFG developed a kinetic model which considered rocking about the contact
zone of a single asperity (Vainshtein, Ziskind, Fichman and Gut"nger, 1997).
Following this approach, we present here a model for particle resuspension in a turbulent
boundary layer which is an extension of the RRH model. We also report the results of a series of
experiments which support ZFG's conclusions, namely that the resuspension rates based on the
RRH lift force model signi"cantly underestimates the amount of resuspension from rough surfaces
4 M.W. Reeks, D. Hall / Aerosol Science 32 (2001) 1}31

Fig. 1. Particle-surface geometry for rock'n roll model.

where there is a spread in adhesive forces. In the revised model we have accounted for drag as well
as lift through the rocking of the particle about the asperities in the contact zone (not about the
contact zone of a single asperity as in Vainshtein et al. (1997)). So rather than oscillating vertically
in a potential well, a particle will oscillate about a pivot P (see Fig. 1). The basic formula for the
resuspension rate involving the average potential energy is basically the same as in the original
RRH model except that this potential energy now depends upon the torque acting on the particle
(including therefore the in#uence of drag as well as lift) and the moment of inertia of the particle
about the pivot.
In Section 2 we present the formulae for the resuspension rate for smooth and rough surfaces
based on the RRH model and then show how these are changed in the light of the in#uence of
couples rather than forces acting on the particle. We also compare the formula for the resuspension
rate with the formula in the absence of resonant energy transfer when the typical frequency of the
aerodynamic forces is much less than the natural frequency of the particle surface deformation at
the point of minimum potential. In this situation the resuspension rate then depends only on the
distribution of adhesive and aerodynamic forces (it does not depend upon the size and shape of the
surface potential well).
In Section 3 we compare predictions of the RRH and revised models with measured values of the
fraction of particles resuspended from a polished stainless steel #at plate in a fully developed
channel #ow boundary layer. In all cases the distribution of adhesive forces was predetermined
from a set of centrifuge experiments. Signi"cantly the revised model based on rocking and rolling
about a point of contact gave much better agreement with measured values of the resuspension
than the original RRH model.
M.W. Reeks, D. Hall / Aerosol Science 32 (2001) 1}31 5

2. The RRH and revised models for resuspension

2.1. The RRH model

In this model the surface adhesive potential well is derived exclusively from the aerodynamic
mean lift and the surface forces using the Johnson, Kendall and Roberts (1971) (JKR) model for
adhesion. A detailed consideration of the random motion within the surface adhesive potential well
(RRH, 1988) leads to a formula for the resuspension rate constant p similar to the desorption rate of
molecules from a surface, namely

 
Q
p"n exp ! , (3)
21PE2
where: n is the typical frequency of the particle-surface deformation within the adhesive potential
well; Q is the height of the potential barrier which depends upon the di!erence between the
adhesive force and the mean lift force; and 1PE2 the average potential energy of particles in the
well.
Both n and 1PE2 contain contributions from the resonant energy transfer. So if g is the ratio of
the near resonance contribution to 1PE2 to the o!-resonance contribution, then for a lightly
damped harmonic oscillator,
p
g" uEK (u), (4)
2b *

where u is the natural frequency of the particle-surface deformation (radians/s), b is the combined
#uid and mechanical damping of the particle-surface motion. EK (u) refers to the value of the energy
*
spectrum (normalised to unity) of the lift force #uctuations at the natural frequency. To be more
precise, if the lift force F (t) at time t is represented by a mean component 1F 2 and a #uctuating
* *
(zero mean) component f (t), whose energy spectrum is E (n) at a frequency n, then by de"nition
* *



1 f 2" E (n) dn, (5)
* *

where 1 f 2 is the covariance of the lift force #uctuations. Then the normalised energy spectrum is
*
given by
E (n)
EK (n)" * . (6)
* 1 f 2
*
In terms of g and u, the particular value of n appropriate for the formula for the resuspension rate
constant p in Eq. (3) is given by

g#(1 fQ 2/1 f 2u) 


  
u
n" * * , (7)
2p g#1
where 1 fQ 2 is the covariance of the time derivative of the lift force #uctuations.
*
6 M.W. Reeks, D. Hall / Aerosol Science 32 (2001) 1}31

We note that when g"0 (zero resonant energy transfer), the value of n reduces to

 
1 
n" (1 fQ 2/1 f 2). (8)
2p * *

This quantity represents the typical forcing frequency of particle motion within the potential well,
and, without resonant energy transfer within the well, this quantity is the maximum removal
frequency/resuspension rate. Of course with the in#uence of resonant energy transfer this value can
be much increased.
Using explicit expressions for Q and 1PE2 involving the force of adhesion f and the mean 1F 2
*
and covariance 1 f 2 of the lift force gives an expression for p of the form
*

 
k( f !1F 2)
p"n exp ! * , (9)
1 f 2(1#g)
*
where k is a numerical constant dependent upon the shape of the surface potential. For the case of
the JKR potential k"1, in contrast to that for a harmonic potential where k". The value for p is

appropriate for a discrete value of the adhesive force f (based on the radius of curvature of the
asperity at the point of contact). In the case of a real surface the contact geometry is characterized
by a wide distribution of surface roughness at the contact which will produce both a reduction and
spread in the force of adhesion compared to that for a perfectly smooth contact. Based on the
observation by several authors, the RRH model assumes that the distribution of adhesive forces
f is a log-normal distribution with the adhesive forces scaling on the adhesive force F for
perfectly smooth contact which in the JKR model for a sphere of radius r is given by
F "pcr, (10)

where c is the adhesive surface energy of the particle and substrate. Thus if the adhesive forces are
normalised on F to give a normalised adhesive force f  , the distribution u( f ) of f  is assumed to
have a log-normal distribution:

    
1 1 1 ln( f  / f  ) 
u( f  )" exp ! , (11)
(2p f  ln p 2 ln p

where: f  is the geometric mean of the normalised adhesive forces f  , and is a measure of the
reduction in adhesion due to surface roughness (1/f  is referred to here as the adhesion reduction
factor); p is a measure of the spread in adhesive forces (referred to here as the adhesive spread
factor).
So if the force of adhesion scales on the particle radius, u( f  ) will remain invariant to particle
size. Signi"cantly even for a nominally smooth surface the distribution can have a large spread
p&10 and a reduction greater than 100 : this in turn leads to a wide range of probabilities of
resuspension for nominally the same particle. In such circumstances the resuspension of particles
from rough surfaces is characterised by two features:

E short term resuspension in which a large proportion of those resuspending do so over a very
short time, typically of the order of 10 ms;
M.W. Reeks, D. Hall / Aerosol Science 32 (2001) 1}31 7

E long-term resuspension for the remainder which resuspend at a rate almost inversely propor-
tional to the time of exposure to the #ow.

The main parameters required for the calculation of resuspension using the RRH model are the
spread and reduction in adhesion and the distribution and time scale of the aerodynamic forces
acting upon the particle. There is very little sensitivity to other material properties. The parameters
required from the characteristics of the aerodynamic lift forces contained in the formula for the
p are taken from measurements of lift forces on particles in a fully developed turbulent boundary
layer on a #at plate which are brie#y described in Section 2.3.

2.2. Dynamic rock'n roll model with resonant energy transfer

The geometry of the particle-surface contact in the revised model is shown in Fig. 1 in which the
distribution of asperity contacts is reduced to a two-dimensional model of two-point asperity
contact. Thus rather than the centre of the particle oscillating vertically, it will oscillate about the
pivot P until contact with the other asperity is broken. When this happens it is assumed that the lift
force is either su$cient to break the contact at P and the particle resuspends or it rolls until the
adhesion at single-point contact is su$ciently low for the particle to resuspend. In either situation
the rate of resuspension is controlled by the rate at which contacts are initially broken. The basic
formula for the resuspension rate constant p is still given by Eq. (3) but now 1PE2 and n depend
upon the torque acting on the particle and the moment inertia of the particle both about P. We
note that the torque includes the in#uence of both drag and lift but the drag (although of similar
magnitude to the lift) notes the greater contribution.
If h represents the angle of de#ection for small oscillations about P, then the equation of motion
for h is
h$ #b hQ #uh"I\!(t), (12)
F F
where I is the moment of inertia of the particle of mass m about P, u is the natural frequency of
F
oscillation, b is the damping and ! is the #uctuating component of the couple ! about P caused
F
by the drag and lift. Referring to Fig. 1 and taking moments about P, gives
a
!& F #rF . (13)
2 * "
For a sphere of mass m,
I" mr. (14)

The displacement/deformation at O about P is given by ah so that the restoring couple is given by
muah. Equating this with Iuh, gives the relationship between u and u the natural frequency
F F
for vertical oscillations as
5a
u" u. (15)
F 7r
To calculate the damping constant b we assume the drag mb caused by small oscillations h about
F
O is the same for a sphere of radius r oscillating in position with velocity rhQ .
8 M.W. Reeks, D. Hall / Aerosol Science 32 (2001) 1}31

So equating the moment of this force about P with Ib hQ gives


F
b "b. (16)
F 
The formula for p is the same as in Eq. (3) except that 1PE2 refers to the average potential energy
due to rotations h about the equilibrium con"guration: thus we have

 
!k!
p"n exp 2 , (17)
F a1 f 2(1#g)

where ! is the net couple acting at P due to the static forces i.e. from the weight of the particle, the
2
force of adhesion and the average aerodynamic forces; f is the #uctuating component (zero mean) of
a force F(t) de"ned as

r
F(t)"F # F . (18)
 * a "

Expressing the couple ! in terms of the relevant forces acting on the particle gives
2

 
k( f #1/2mg !(r/a)mg !1F2)
p"n exp !   , (19)
F 1 f 2(1#g)

where n is related to u by the same relationship as for n and u with f replacing f . For
F F *
completeness we have included the in#uence of gravity (see Fig. 1) so that in Eq. (19) g and g are
 
the gravitational components of acceleration acting normal and tangential to the surface. In
practice the contribution from gravity is much smaller than that from the drag force although it
may well be important in re-establishing contact with the surface once the contacts have been
broken. The contribution g to the potential energy from the resonant energy transfer is similarly
given by Eq. (4) with b and u replacing b and u, respectively.
F F
There are two features of this revised model that are worth noting:

1. Because r/a<1, the e!ective force F(t) and hence the resuspension rate constant will be
dominated by the contribution from the drag force.
2. Since u is much lower than the value of u we would correspondingly expect the value of the
F
energy spectrum at u to be much greater than the corresponding value at u. This will
F
correspondingly increase the value of the resonant energy contribution g in the revised model
compared with that in the RRH model. However this increase will be compensated by
a reduction in g by the reduction in the value of the ratio of u/b . See Eq. (4).
F F

2.3. Formulae for the contributions of the aerodynamic forces and damping to the
resuspension rate constant

To obtain values for the resuspension rate constants from both models we have used Hall's
measurements of both the mean lift and #uctuating lift force (Hall, 1988, 1998) in fully developed
turbulent channel #ow.
M.W. Reeks, D. Hall / Aerosol Science 32 (2001) 1}31 9

(a) Mean lift and drag force


Hall demonstrated that the mean lift force scaled on the #uid density o, kinematic viscosity l and
friction velocity u as follows:
s

 
1F 2 ru  
* +20.9 s . (20)
ol l

The #ows and particle sizes used in these measurements covered the range 1.8(ru /l(70.
s
The in#uence of surface roughness was also considered. However for the geometries used in the
experiments we report here, the surfaces can be treated as #at. Similar measurements of the mean
drag are not available and values in the rock'n roll model are based on the mean drag on a sphere
near a surface in simple shear #ow being 1.7 times the Stokes drag (O'Neill, 1968). The result is used
here for the mean turbulent #ow in the predominantly viscous sub-layer. Thus

 
1F 2 ru 
" +32 s . (21)
ol l

As the same #ow structures responsible for lift are responsible for the drag force it would seem
reasonable to assume that lift and drag are well correlated.
(b) RMS lift and drag
Hall's (1998) measurements show that the ratio of the rms to the mean lift is &0.2 for a sphere
in a fully developed turbulent boundary layer in an aerodynamically smooth channel. These
measurements also show that this ratio is roughly 0.2, independent of surface roughness:
however mean values can increase by a factor of 3. In the absence of similar measurements for the
#uctuating drag force, the ratio of the rms to mean drag is assumed to be the same as that for
the lift.
(c) Geometric factor G

The geometric factor here refers to the ratio of r/a in Fig. 1 and in general is related to the surface
roughness on the region of contact of the particle with the substrate. An estimate of this value can
be obtained from the ratio of the tangential force to the normal force required to dislodge identical
particles from a surface: details of these measurements are given in Section 3.1 and suggest a
value &100.
(d) Normalised energy spectra
EK (n) as in Eq. (4) for the resonant energy transfer contribution g is calculated with reference to the
energy spectrum E>(n>) in units of the wall frequency u/l, so that n> is nl/u. So,
O O

 
l
EK (n)" E>(n>). (22)
u
O
Only measurements of the energy spectrum E>(n#) of the lift force are currently available so we
*
assume here that this is the same for the energy spectrum of the drag force and hence also for the
energy spectrum of the e!ective zero mean #uctuating force f (t). We use here the empirical formulae
for an aerodynamically smooth surface obtained from Hall's measurements of the energy spectra of
10 M.W. Reeks, D. Hall / Aerosol Science 32 (2001) 1}31

the #uctuating lift force (Hall, 1998). Explicitly


E>(n>)"58.06, n>)0.0054,

E>(n>)"0.0812(n>)\ , 0.0054(n>(0.104, (23)

E>(n>)"0.0000173(n>)\ , n>*0.104.
(e) Natural and forcing frequencies
The value of u is obtained from Eq. (15) using the expression for u given in RRH (1988) which
F
gives its dependence on the elastic restoring forces as well as the adhesive forces; explicitly

u"(s/m, (24)
where for a particle of weight mg, the sti!ness is given in terms of an applied load P !1F 2
 *
namely,

 
P #P
s(P )"KrP   , (25)
   5P #P
 
where R is the asperity radius at the point of contact (the other asperity from P in Fig. 1 for the

rock'n roll model); the elastic constant K is de"ned in terms of particle and substrate with Young's
moduli E and E and Poisson's ratios l and l as
   

 
4 1!l 1!l \
K" #  , (26)
3 E E
 
P is given as

P "P #3pcr #[6pcr P #(3pcr )], (27)
  
where r is the radius of curvature of the asperity at the point of contact.
The value for the typical frequencies n and n which includes the in#uence of both the forcing
F
frequencies (o! resonance) and natural frequency (near resonance) is obtained from Eq. (7) using
the formula


(1 fQ 2/1 f 2) u
"0.00658 O , (28)
2p l
derived from Hall's measurements of the energy spectrum of the lift force (which is assumed here to
be the same as that for the e!ective #uctuating force f (t)). We recall that in calculations of the
resuspension rate in Reeks et al. (1988) where the RRH model was "rst described, this quantity was
set equal to the so-called bursting frequency l which represented the typical frequency of
turbulent velocity #uctuations in a fully developed turbulent boundary layer in channel #ow (since
no measurements of the energy spectrum of the lift force were available at the time). The value
l used was taken from the measurements of Blackwelder and Haritonidis (1983), namely,

 
u
l +0.0033 O . (29)
l
M.W. Reeks, D. Hall / Aerosol Science 32 (2001) 1}31 11

Whilst we would not expect the typical frequency of the random lift force #uctuations to be the
same as those for velocity #uctuations we might expect them to be of the same order and
a comparison of the formulae in Eqs. (28) and (29) con"rms that this is indeed the case.
( f ) Damping constant
The formula for calculating the damping constant in the rock'n roll model is based on that for
the #uid and mechanical damping used in the RRH model. We assume that the #uid damping is the
same as for a sphere of mass m oscillating in an unbounded #ow with small amplitude of oscillation.
We recall that the mechanical damping in the RRH model is based on energy loss from elastic wave
propagation.

2.4. &Quasi-static' rock'n roll model

If the motion of particles on the surface is driven by turbulent removal forces which are
o!-resonance (no resonant energy transfer) then the particle equation of motion can be approxi-
mated by a force balance between the aerodynamic removal forces and the adhesive forces or
a balance of couples as in Fig. 1. Taking the force balance as an example,

F (y)#F"0, (30)

where F is the adhesive restoring force as a function of the deformation of the particle/surface and

F(t) is the aerodynamic removal force as a function of time t. When dealing with a balance of
couples F(t) is the force given in Eq. (18). Splitting this force into an average component 1F2 and
#uctuating component of zero mean f (t) we have from Eq. (30)

f (t)"!F (y)!1F2. (31)



At the point of detachment y"y we can write

f "!F (y ), (32)

so that we have formally at the detachment point

f"f !1F2. (33)

In general then, Eq. (31) de"nes a relationship between f and y so that

y(t)"t( f ), y (t)"fQ t( f ). (34)

The resuspension rate constant p can then be shown to be

  
 D 
p" fQ P( f !1F2, fQ ) d fQ P( f !1F2, fQ ) d fQ d f, (35)
 \ \
where P( f, fQ ) is the joint probability distribution for f and fQ . The formula used in RRH (1988) is
consistent with the joint distribution P being jointly Gaussian with zero correlation between f and fQ ,
12 M.W. Reeks, D. Hall / Aerosol Science 32 (2001) 1}31

i.e. the two variables are statistically independent of one another. Thus

P( f, fQ )"(fQ  f )G(1)G(m), (36)

where G( ) is a Gaussian distribution normalised to unity

G(1)"(2p)\ exp (!1/2) (37)

and 1 and m are two statistically independent random variables of unit variance.
Substituting the analytic form for G( ) into the formula for p and evaluating the integral gives
explicitly

   
1 fQ 2  !( f !1F2) 1
p"(1/2p) exp +1#erf (( f !1F2)/(21 f 2),. (38)
1 f 2 21 f 2 2

We recall that the maximum resuspension rate in the dynamic rock'n roll model with no resonant
energy transfer (which is typical of the frequency of the forcing motion) is given by the formula

 
1 fQ 2 
p"n "(1/2p) . (39)
F 1 f 2

So for consistency with this model we use the formula in Eq. (38) so long as p/n (1, otherwise we
F
set the value of p"n . Evaluating the ratio p/n from Eq. (38) as a function of ( f !1F2)/1 f 2
F
indicates that p/n&1 when ( f !1F2)1 f 2"0.75. So in the implementation of this model and
for comparison with the dynamic rock'n roll model (with no resonant energy transfer) therefore

p"n for ( f !1F2)/1 f 2)0.75,


F (40)
p is given by Eq. (38) for ( f !1F2)/1 f 2'0.75.

We note that the formula in Eq. (38) does not depend upon the shape of the potential, only upon the
magnitude and distribution of aerodynamic forces/couples acting on the particle: we have chosen
a Gaussian distribution here for comparison with the RRH and revised models. In this respect we
note that the RRH and revised models in the absence of resonant energy transfer agree with the
formula above when the aerodynamic forces are much less than the adhesive forces in which case
the denominator in Eq. (38) is unity and the potential well is harmonic (in which case the value of
the numerical constant in Eq. (9) is ). However the analysis implies that in the case of zero resonant

energy transfer the value of this numerical constant should be still  even when the potential well is

not harmonic as in the case of the JKR potential.
In Fig. 2 we compare predictions using the formula in Eq. (38) with those obtained using the
Jurcik and Wang (J&W) formula for a Gaussian distribution of removal forces, namely,

p"n  erfc (( f !1F2)/(21 f 2), (41)


F
where for the purposes of comparison we have used the same value of n in either case. In other
F
words the formulae apply to the same system: however the J&W formula has no rigorous
M.W. Reeks, D. Hall / Aerosol Science 32 (2001) 1}31 13

Fig. 2. Comparison of resuspension rate constants based on Jurcik and Wang model (Eq. (41)) with quasi}static rock'n
roll model (Eq. (40)).

foundation and is strictly empirical. It is clear from Fig. 2. that the J&W formula signi"cantly
under-estimates the resuspension rate constant compared to that calculated using the &quasi-static/
force balance' model formula in Eq. (38). This is especially so at large values of the adhesive force
compared to the rms or the mean of the removal force. This has an important bearing also on the
long-term resuspension rate from rough surfaces.

3. Measurements of the adhesion and resuspension of particles

We report here on the measurements that were made of the adhesive forces of alumina and
graphite particles on a polished stainless-steel substrate and the resuspension of those particles
when the same particles & substrate were exposed to a fully developed turbulent channel #ow.

3.1. Measurement of particle size and adhesive forces

The alumina particles used in the resuspension experiment were nominally mono-dispersed 10
and 20 lm spheres. The graphite particles were angular in shape with a broader spread in size.
14 M.W. Reeks, D. Hall / Aerosol Science 32 (2001) 1}31

Table 1
The particle size distributions

Particle type Minimum particle Maximum particle Geometric mean Geometric standard
diameter (lm) diameter (lm) particle diameter deviation
(lm)

10 lm alumina 6 18 12.2 1.15


20 lm alumina 16 30 23 1.17
Graphite 6 30 13 1.85

Fig. 3. Centrifuge data for normal and tangential forces for nominal 10 lm alumina particles.

Measurement of cumulative size distribution showed all three sets of particles to have truncated
log-normal size distributions, details of which are given in Table 1 below.
The adhesion forces for a mono layer of these three particle types distributed uniformly over
a polished type 316 stainless-steel surface were carried out using an MSE Superspeed 75 centrifuge
in the manner described by Reed and Rochowiak (1988). In all three cases measurements of the
M.W. Reeks, D. Hall / Aerosol Science 32 (2001) 1}31 15

Fig. 4. Centrifuge data for normal force for nominal 20 lm alumina particles.

fraction removed were made for forces applied both normal and tangential to the substrate upon
which the particles were deposited. For the normal forces the e!ects of humidity were examined.
Relative humidity was found to have a signi"cant e!ect on the adhesion. The results we report here
are for dry conditions. Figs. 3}5 show the fraction of particles removed versus centrifuge speed for
each particle type and nominal size for both normal and tangential forces applied to the particle-
substrate. Also shown are the earlier reported measurements of Reed and Rochowiak (1988) for
nominally the same particles. Note the absence of data on tangential force for the nominal 20 lm
alumina particles: the particles came o! the tangential centrifuge cell so easily that no measure-
ments were possible.
The conversion of fraction removed versus centrifuge speed to distribution of normal and
tangential forces follows the approach of Reed and Rochowiak (1988). Thus for a particle of mass
m, the normal force f balances the centrifugal force when

f "m¸), (42)

where ) is the angular frequency in radians s\ of the centrifuge rotor and ¸ is the distance that the
particles are located from the centre of rotation. A similar equation exists for the tangential
adhesive force f : however in terms of the rock'n roll model this balance should be interpreted as

a balance of couples due to the normal force of adhesion and the centrifuge force. That is with
16 M.W. Reeks, D. Hall / Aerosol Science 32 (2001) 1}31

Fig. 5. Centrifuge data for normal and tangential forces for graphite particles.

reference to Fig. 1:

f "(a/r) f (43)

(We note that the location of point P in Fig. 1 is not exact.)
So that measuring simultaneously the normal and tangential forces gives a measure of the
geometric factor G &a/r. If (u f  ) is the distribution of normalised adhesive forces f  "f /F

where F is the adhesive force for perfectly smooth contact for a particle of radius r with a #at
surface, then the fraction remaining of particles of size r is



F (r)" u( f  ) d f  . (44)

D pAP
Integrating over the particle size distribution P(r), the total fraction adhering, F is
0


B  
F " P(r)F (r) dr (45)
0 
B 


with a similar equation for the fraction removed for tangential forces ( f replacing f ). Referring

to the log-normal distribution in Eq. (11), values for spread and reduction in adhesion were
M.W. Reeks, D. Hall / Aerosol Science 32 (2001) 1}31 17

Table 2
Fitted values for the adhesive force distribution

Particle Force Measurement Spread p Reduction

10 lm alumina Normal Set 1 49 592


10 lm alumina Normal Set 2 208 56
10 lm alumina Normal Reed and Rochowiak (1988) 2.55 37
10 lm alumina Tangential Set 1 10.4 848
10 lm alumina Tangential Set 2 2.95 1053
20 lm alumina Normal 78 56
Graphite Normal 489 1.55
Graphite Tangential 19 16

Table 3
Material properties required to calculate particle resuspension using the RRH model

Material Graphite Alumina

Interfacial surface energy, Jm\ 0.15 0.56


Substrate density (steel), Kg m\ 7830 7830
Substrate Young's modulus, Pa 2.1;10 2.1;10
Particle Young's modulus, Pa 2.0;10 3.5;10
Substrate Poissons ratio 0.29 0.29
Particle Poissons ratio 0.3 0.3
Particle density, Kg m\ 2300 1600

"tted to minimise the measured and predicted values of F . Table 2 gives the best "t for the
0
various cases. Values for the surface energies are taken from Table 3 containing the material
parameters.
We note that the "tted values show large variations between various nominally identical sets of
data. Fig. 6 shows the "ts for various values of p for the 10 lm normal adhesion data (set 1) using
the Reed & Rochowiak data which gave much better "ts to the data than the more recent
measurements shown also in Fig. 2. It is evident that the Reed and Rochowiak data follow the
log-normal distribution quite well, unlike the other data. One possible explanation for the
di!erence is that not enough care may have been taken to ensure that the surfaces used in the most
recent measurements were su$ciently well aligned i.e. the tangential component of the force
applied may well have exceeded the tangential removal force for relatively small misalignments,
given that the tangential removal forces are typically 100 times less than the normal removal forces.
The small values of the adhesion reduction factor for graphite is probably a re#ection of the fact
that the graphite particles have a greater number of contacts with the substrate than the alumina
particles.
18 M.W. Reeks, D. Hall / Aerosol Science 32 (2001) 1}31

Fig. 6. A comparison of the Reed and Rochowiak nominal 10 lm alumina centrifuge data with predictions for di!erent
adhesive spread factors with adhesion reduction factor"37.

3.2. Measurements of resuspension

3.2.1. Experimental set-up


Fig. 7 shows a schematic diagram of the air rig in which resuspension was observed. Flow,
provided by a Gri$n blower, passes through a series of ducts into a plenum chamber to damp out
any #ow #uctuations. The air passes through an absolute "lter to ensure that no particulate is
present in the air stream. To establish fully developed #ow, the air then passes through a 5 m long
0.2 m;0.02 m rectangular duct in which the test section is situated 3.5 m downstream along the
duct. At the end of the duct a sample of the air passes through a Royco counter which has the
facility to record the particle concentration as a function of time in four size intervals. The Royco is
used to check that the resuspension coincides with the #ow changes and not some other in#uence.
Finally the air is exhausted to the atmosphere. Care was taken in constructing the wind tunnel to
ensure that there were no steps upstream of the test section that could disturb the #ow. The
test section is shown in Fig. 8. It comprises a polished type 316 stainless-steel surface which
incorporates a centrifuge cell mounted #ush with the surface. Windows are situated on the sides

 A Royco is the trade name of an instrument commonly used for counting aerosol particles based on re#ected light
from gas-borne aerosol particles.
M.W. Reeks, D. Hall / Aerosol Science 32 (2001) 1}31 19

Fig. 7. Air rig and test section.

of the duct to allow the particles to be illuminated and photographed. A 120 mm Hasselblad
camera body "tted with a 135 mm lens and a bellows unit was used for the photography.
This allows a 1 : 1 magni"cation to be used. Using Ilford Pan F50 ASA it was possible to resolve
10 lm particles.
The rig #ow and humidity were monitored as follows:
E the test section pressure was monitored using a KDG series 3000 absolute pressure transducer;
E the test section temperature was monitored by a platinum resistance thermometer;
E the humidity was monitored using a MCM dewlux moisture meter, which measures the vapour
pressure of water in the gas; the pressure at the sensor was also measured using a KDG series
3000 pressure transducer to allow the relative humidity to be calculated;
E the velocity of the #uid a few mm above the surface was measured using a pitot static tube; the
pressure di!erence was recorded using a Furness FC012 di!erential pressure transducer.
The electrical outputs of the above transducers were recorded by a Christie CD248 data logger.
The camera shutter state was also recorded by the logger to enable the #ow to be determined when
the photographs were taken. The Royco output was transmitted via an RS232 link to an IBM PC.
20 M.W. Reeks, D. Hall / Aerosol Science 32 (2001) 1}31

Fig. 8. The test section.

3.2.2. Procedure
Prior to each resuspension run, the particles used were sub-divided using a spinning ri%er
to ensure that the size distributions of particles used in the various runs were as similar as possible.
The test section was polished with diamond paste, then degreased using a 5% solution of Decon 90
in distilled water. The particles were deposited on the surface by transporting them by compressed
air through two impinging jets which broke up any agglomerates. The suspension was then
injected into still air and allowed to settle on the test surface and on the centrifuge cells for
adhesion measurements. Samples of the particles deposited both upstream and downstream
of the test section were taken for subsequent size analysis using an IBAS image analyser. The
test section was then placed in the air rig. A photograph of the particles was then taken when
the air rig was started and the velocity increased in steps with photographs taken of the test
surface at each step. After the run the photographs taken were compared and the fraction
remaining on the surface at each stage was determined by observing which particles had left the
surface.
M.W. Reeks, D. Hall / Aerosol Science 32 (2001) 1}31 21

Fig. 9. Comparison of resuspension measurements with model predictions for the nominal 10 micron alumina particles.

3.2.3. Resuspension measurements


A series of 20 resuspension runs for both graphite and alumina particles were performed. The
results for the nominal 10 and 20 lm alumina are shown in Figs. 9 and 10 and those for graphite in
Fig. 11. In all runs the Royco recorded particles present in the gas stream when the #ow was
increasing. At no other times were particles counted. In comparison with the adhesion measure-
ments, the resuspension measurements from run to run were far more repeatable.

4. Comparison of resuspension measurements with model predictions

4.1. Comparison with model predictions

Predictions of the resuspension using the models described are based on the formulae for the
relevant properties of the aerodynamic forces given in Section 2.4, the values of the material
properties given in Table 2 and the adhesion measurements described in Section 3.1. The material
properties of graphite are taken from gas-cooled reactor core data (Reed, 1983) and those of
alumina (including the surface energy) from Kendall, Alford and Birchill (1987). The values chosen
for the reduction and spread in adhesion used in the models are given in Table 4 for the RRH and
the rock'n roll models (with and without resonant energy transfer) and need some explanation.
22 M.W. Reeks, D. Hall / Aerosol Science 32 (2001) 1}31

Fig. 10. Comparison of resuspension measurements with model predictions for the nominal 20 lm alumina spheres.

For the RRH comparison the distribution of normal forces is required and for the 10 lm particles
the values obtained from the Reed & Rochowiak data have been chosen since for the reasons stated
they appear to be the most reliable. The spread for the 10 lm particles are used for the 20 lm
particles as it seems likely that the data are suspect as described above. Examination of Figs. 3}5
indicate that the centrifuge had to spin 10 times faster to dislodge a proportion of particles
normally than by removing the same proportion tangentially. As the applied force is proportional
to the square of the spin speed this implies that the net value of r/a&100. The fact that on any
surface there will be a distribution of a means that we would expect a broader distribution of
tangential forces than normal forces: this is true in the case of alumina particles but not for the
graphite particles. Since r<a, the drag force will dominate the couple, so consequently the removal
force will be tangential to the substrate. Hence the adhesive spreads for the tangential data have
been used in predictions using the rock'n roll model.
Figs. 9}11 give the measured values of the fraction remaining as a function of the friction velocity
for the alumina and graphite particles. Also shown are the corresponding predictions of the RRH
and rock'n roll models. Clearly the rock'n roll model gives much better agreement with measured
values of the resuspension than the original RRH model, the agreement being quite good
considering the large uncertainty in the adhesive force measurement. Also shown are the predic-
tions of the rock'n roll model without resonant energy transfer showing that resonant energy
transfer has very little contribution. This means that a quasi-static rock'n roll model is appropriate
M.W. Reeks, D. Hall / Aerosol Science 32 (2001) 1}31 23

Fig. 11. Comparison of resuspension measurements with model predictions for the graphite particles.

Table 4
Fitted values for the adhesive force log-normal distributions

RRH lift force model Rock'n roll model

Data set Adhesive Adhesion Adhesive Adhesion Geometric


spread p reduction spread p reduction factor r/a

10 lm alumina 2.55 37 10.4 37 100


20 lm alumina 2.55 2.55 10.4 56 100
Graphite 489 1.55 19 1.55 100

and the comparison with these predictions are shown for a Gaussian distribution of aerodynamic
forces (strictly speaking couples): for the case of the longer-term resuspension particles resuspend-
ing in a time &1 s, the di!erence between the two formulae is re#ected in the value of k in the
formula for p in Eq. (9). However in practice because of the broad spread of adhesive forces for
a rough surface the di!erence between the fraction remaining is only slight, see Fig. 12 for the
24 M.W. Reeks, D. Hall / Aerosol Science 32 (2001) 1}31

Fig. 12. Comparison of rock'n roll model predictions with no resonant energy transfer for the nominal 10 lm alumina
spheres: model 1 Eq. (19) with k"1; model 2 Eq. (40) for a Gaussian distribution of aerodynamic forces.

nominal 10 lm alumina spheres. Indeed despite the di!erences between the rock'n roll model and
the Jurcik and Wang models for the value of the primary resuspension rate p (See Fig. 2), again
because of the broad spread in the adhesive forces, there is little di!erence in the predicted values
for the integrated resuspension (see Fig. 13). However there is a marked di!erence in the behaviour
of the long-term resuspension rates (see Fig. 14).

4.2. Sensitivity to material parameters

It is instructive to determine the actual sensitivity of the model to the material/adhesive


parameters. The base case is for the nominal 10 lm. alumina spheres in a #ow with a friction
velocity of 1m s\. The results are presented in Table 5 below. The lack of sensitivity to particle
density and K is evidence that resonant energy transfer is extremely small. Note that the spread in
adhesion p is so large that the sensitivity to the material parameters in minimal. Consequently, the
predictions of the model will be robust to possible variations in the values of the various material
parameters. The di!erences between the adhesive reduction factor for the 10 and 20 lm is so small
that particles of di!erent sizes can be characterised by the same spread and reduction in adhesion
cannot be disproved.
M.W. Reeks, D. Hall / Aerosol Science 32 (2001) 1}31 25

Fig. 13. Comparison of Jurcik and Wang model (- - -) and rock'n roll model (solid line) predictions of short-term
resuspension for the nominal 10 lm alumina spheres.

4.3. Sensitivity to other parameters

We have made certain assumptions about the aerodynamic forces that are important to re#ect
upon especially when considering the resuspension of sub-micron size particles outside of the range
measured here. Such particles are likely to have radii in wall units ;1.8 and the values for the
mean lift obtained using Eq. (20) are likely to be a signi"cant over-estimate. However since the drag
force is the dominant removal force this overestimate should not a!ect the resuspension to the
same degree: the formula for the drag force is not subject to the same sensitivity as the lift force very
close to the wall.
We recall that we have used a single value for the so-called geometric factor (see Section 2.3c)
which is a measure of the distance between asperities (see Fig. 1). It is clear that for a rough surface,
the geometric factor is not a constant but, like the adhesive forces, a statistical quantity and its
distribution would lead to a further broadening of the distribution of adhesive couples (in addition
to the broadening due to the adhesive forces themselves). In some sense then, the apparent
distribution of the measured adhesive forces contains this contribution: that is the measured
distribution is determined both by the spread in the distance between asperities in contact with the
26 M.W. Reeks, D. Hall / Aerosol Science 32 (2001) 1}31

Fig. 14. Predictions of long-term resuspension rates for the nominal 10 lm alumina spheres based on rock'n roll model
and Jurcik and Wang model. Also shown 2(ln (nt))\t\.

Table 5
The sensitivity of the fraction remaining (after resuspension) to the material parameters

Parameter Fraction remaining (after 1s.)

Parameter name Base value Base case Parameter Parameter

doubled halved
K, Pa 1.9;10 0.541 0.541 0.541
Adhesive surface energy c, J m\ 0.56 0.541 0.655 0.424
Particle density, kg m\ 1600 0.541 0.540 0.541
Adhesion reduction factor 56 0.541 0.423 0.655

particles and the spread in the radii of curvature of the asperities at the points of contact (which
determine the adhesive forces).
Finally the formula we have used to calculate the resuspension rate in practice assumes
a Gaussian distribution of drag forces. In the absence of any measured values for this distribution
M.W. Reeks, D. Hall / Aerosol Science 32 (2001) 1}31 27

this is the only "rst approximation we can reasonably make though the model does not of itself rely
implicitly upon this assumption. It may well be that the major contribution to #uctuating drag
component is associated with the intermittent bursting processes close to the wall which are
strongly non-Gaussian in nature.

5. Longer-term resuspension rates

The experimental set up reported here was not best suited to measuring the long-term resuspen-
sion rates and we refer to an earlier experiment involving the transport of alumina particles down
a tube (Hall & Reed, 1988) and to the measurements and analysis of deposition and long-term
resuspension of particles in recirculating #ows (Reeks & Hall, 1988). However it would seem
appropriate to make some comments about long-term resuspension rates with regard to the rock'n
roll model and also that of earlier models. We recall that it was a basic feature of the original RRH
model that for times <the typical time scale of the removal forces, particles were resuspended
from rough surfaces at rates which varied almost inversely with the time. In fact this behaviour
could extend for many hours, being very robust to changes in #ow, particle size and material
properties. So long as the roughness is associated with a very broad spread in the distribution of
adhesive forces, which is true even for nominally very smooth surfaces, then this inverse decay is
una!ected by the detailed nature of the roughness. The resuspension rate K(t) is given formally by



K(t)" pe\NRu( f ) d f , (46)

where we recall that u( f ) is the distribution of adhesive forces f . The inherent (1/t) dependence
re#ects the fact that the primary resuspension rate pe\NR has a maximum value t\e\. Thus as
time increases, the major contribution to the overall resuspension rate shifts towards greater
adhesive forces so that the greatest contribution to the resuspension rate is around values of the
adhesive forces for which p+1/t. The degree to which the overall resuspension rate follows 1/t
depends upon the range of these values of f (for which p+1/t) and how sensitive it is to the value
of t. In the case of a Gaussian form for p this contribution comes from a range of values that varies
as +ln (nt), about the value for which p is a maximum, where we recall n is the typical frequency of
the aerodynamic removal forces. Clearly when nt<1, this width changes on a much longer time
scale than the value of p associated with that range namely (1/t), suggesting that the resuspension
rate is close to 1/t decay. To illustrate the point suppose we normalise the adhesive force f on the
rms of the removal force and consider the case for which
p"n exp (!f K ), (47)
where fK is the normalized adhesive force. We suppose that the distribution u( fK ) is constant over
the range of values for which p exp(!pt)&1/t. The maximum value of p exp(!pt) occurs when
f "(ln (nt)) furthermore for nt<1,
K(t)+u(+ln (nt),)+ln (nt),\t\. (48)

Fig. 14 gives an indication of how accurate this approximation is compared with the value of K(t)
evaluated numerically. Note that the decay is always greater than t\.
28 M.W. Reeks, D. Hall / Aerosol Science 32 (2001) 1}31

The longer-term resuspension rate behaviour is exactly the same in the new rock'n roll model
as in the original RRH since the basic form for K(t) in Eq. (46) remains unchanged and the Gaussian
form for p is also the same. As an example we have calculated the long-term resuspension rate
for the nominal 10 lm alumina particles based on the "tted values for the adhesive
force log-normal distributions in Table 4 and a value of the friction velocity of 10 m s\. The results
are shown in Fig. 14 where they are compared with the values based on the Jurcik & Wang form
for p given in Eq. (2). It is evident that the longer-term resuspension rate using the J&W form
gives signi"cantly greater less long-term resuspension and the proximity of the decay to 1/t decay
persists for signi"cantly less time than that based on the Gaussian form for p in the rock'n roll
model here.

6. Some 5nal comments on kinetic models for resuspension

We recall that Ziskind et al. (1995) classi"ed kinetic models for particle resuspension into
two types: force/moment balance models and energy accumulation models. We would prefer to
classify these models as either dynamic models (energy accumulation) or quasi-static models
(force/moment balance). The quasi-static model then becomes a special case of the dynamic model
as indeed we have demonstrated. More importantly a dynamic model allows for a continual input
of turbulent energy (via the turbulent aerodynamic forces and their moments) until this rate of
input is balanced by the rate of energy dissipation either by #uid/mechanical damping. It is well
known that such a process of energy transfer takes place most e$ciently when the driving
forces/couples are applied at frequencies around the natural frequency of the particle-surface
deformation. It means that depending upon the e$ciency of the resonant energy transfer process,
particles can be removed by forces/moments which are less than the force/moment of adhesion. The
three models that fall perfectly into that category are the original RRH model (appropriate for
removal forces applied normal to the surface) and the rock'n model described here and the model of
Vainshtein et al. (1997). The latter two models are extensions of the original RRH approach but
recognise the contribution from drag forces (tangential forces) through the onset of rolling: so both
these models refer to a &balance' of rotational energy through the action of moments rather than
forces. Both models complement one another: the Vainshtein model (1997) is appropriate
for perfectly smooth contact between particle and surface and the rock'n roll model appro-
riate for rough surface contact. Together the two models give a comprehensive view of particle
resuspension.
For the resuspension measured in these experiments the contribution from resonant energy
transfer is generally small. Under these circumstances the rock'n roll model and the Vainshtein
et al. model are then consistent with a quasi-static model in which the removal forces and their time
derivatives are Gaussian and statistically independent of one another. Under quasi-static condi-
tions we have shown that a more general expression can be used for the resuspension rate p for
which the distribution of the removal forces and the time derivatives is entirely arbitrary. We note
that a Gaussian distribution of removal forces leads to a Gaussian form for p as a function of the
adhesive force and re#ects the fact that the concentration of particles at the detachment point is
itself Gaussian. In this respect we note also that the dependence on material/#uid properties is
much reduced, depending only on the distribution of adhesive forces and the distribution of the
M.W. Reeks, D. Hall / Aerosol Science 32 (2001) 1}31 29

aerodynamic forces (principally the drag force) and their time derivatives. In these circumstances
the quasi-static forms for p (proposed here and in Vainshtein et al.) have certain similarities to the
models originally proposed by Wen and Kasper (Eq. (1)) and Jurcik and Wang (Eq. (2)). However
there are important di!erences in interpretation.
In the Wen and Kasper (1989) model, whilst drawing upon the analogy with molecular
desorption in using an Arrhenius form for the rate constant p in Eq. (1), the exponent q is de"ned to
be the ratio of the adhesive force over the instantaneous removal force. This is at variance with
a statistical approach to resuspension as presented here and elsewhere where p is a statistical
average over the distribution of removal forces and their time scales. So p in a formally statistical
approach can only contain the statistical moments of the removal force F represented as a stochas-
tic process [F(t); 0)t)R]. This feature is also evident in the Arrhenius formula which contains
the temperature which is of course a statistical property of the random motion of the molecules.
It is also very much apparent in the formula for the release rate of Brownian particles
from a potential well which is a direct analogue of the resuspension of particles by a turbulent
#ow (see e.g. Chandrasekhar, 1943). Furthermore in choosing q as the instantaneous ratio of
removal over adhesive force, the formula for p admits the possibility that resuspension will
occur with values of the removal force less than the adhesive force. This is at variance with
a quasi-static model. However this inconsistency can easily be resolved by rede"ning q as the ratio
of the adhesive force to the rms of the removal force or perhaps more precisely and in line with
the formula in Eq. (38),

q"+ f !1F2,/1 f 2 (49)

cf the de"nition of f) in Eq. (47).


In both the Wen & Kasper and Jurcik and Wang models there remains still further the
interpretation of n, the maximum frequency of removal in the formula for p. In either model it
is treated as an arbitrary independent variable whereas in reality it is intimately bound up in
the motion of the particle surface deformation and therefore to the average properties of the
process of [F(t)]. That is formally evaluating p in the way we have done in Section 2.4 (see Eq. (38))
determines both n and q in terms of the statistical properties of [F(t)]: in fact, we have been able to
determine n precisely from a measurement of the energy spectrum of [F(t)]. Finally whilst
the Jurcik and Wang formula for p is consistent with the view that p must depend only on average
properties of [F(t)], the form suggested for a Gaussian distribution for F(t) whilst being plausible is
however inconsistent with the form given in Eq. (38) using the same assumptions but rigorously
derived.

7. Summary and conclusions

We have presented a series of measurements of normal and tangential adhesive forces for solid
particles on a nominally smooth polished surface. The results were consistent with a log-normal
distribution of adhesive forces with a very broad spread and were much reduced on average
compared to the normal adhesive force for perfect contact of a sphere with a smooth surface
(absence of asperities). Signi"cantly the pull-o! force tangential to the surface (tangential adhesive
30 M.W. Reeks, D. Hall / Aerosol Science 32 (2001) 1}31

force) was about 1/100th of the normal pull-o! force. This latter result has important implications
for the way particles are resuspended from surfaces and highlights the dominance of the drag force
over the lift force even though both forces are of similar magnitude for the particles considered
here Accordingly we compared our resuspension measurements with predictions of our earlier
model for resuspension (RRH model) in which the aerodynamic forces were exclusively lift
forces and a rock'n roll model which included the in#uence of drag as well as lift on resuspension.
The latter model considered the in#uence of couples about a single asperity in breaking
the contacts with the remaining asperities, the drag having a much greater e!ect by a factor of
100 over the lift force consistent with our tangential and normal adhesive force measurements. We
found that using measured values for the spread and reduction in adhesion as input parameters,
the original RRH model consistently underpredicted the resuspension compared with the rock'n
roll model which gave valves much closer to the experimental measurements. Both models
admit the possibility of resonant energy transfer although in the rock'n roll model this contribution
is much less. In the predicted values of resuspension the contribution of resonant energy transfer
is generally small (in the case of the alumina particles it is virtually zero). Under these circumstances
the resuspension rate dependence on material properties is much reduced, depending only
on the distribution of adhesive forces and the distribution of the aerodynamic forces (principally
the drag force) and their time derivatives. A simpler but more exact model for resuspension
was constructed: however in practice this gives very similar results to the original rock'n roll
model.

References

Braaten, D. A., Shaw, R. H., & Paw, U. K. (1990). Particle Resuspension in a turbulent boundary layer. Journal of Aerosol
Science, 21, 613}628.
Blackwelder, R. F., & Haritonidis, J. H. (1983). Scaling of the bursting frequency in turbulent boundary layers. Journal of
Fluid Mechanics, 132, 87}103.
Chandrasekhar, S. K. (1943). Reviews of Modern Physics, 15, 1}89.
Reed, J. R. (Editor), (1983). Compendium of CAGR Core and Sleeve Data and Methods, CSDMC/P28 (1983).
Hall, D. (1988). Measurements of the mean lift force on a particle near a boundary layer in turbulent #ow. Fluid
Mechanics, 187, 451.
Hall, D. (1998). Measurements of the #uctuating lift force on a particle at a boundary in turbulent #ow. Journal of
Experimental Fluids, submitted.
Hall, D., & Reed, J. (1988). Transport of particles through a pipe. Journal of Physics D, 21, 1481}1485.
Johnson, K. L., Kendall, K., & Roberts, A. D. (1971). Proceedings of the Royal Society of London A, 324, 301.
Jurcik, B., & Wang, H.-C. (1991). The modeling of particle resuspension in turbulent #ow. Journal of Aerosol Science,
22(Suppl 1), S149}S152.
Kendall, M. C., Alford, W. E. T., & Birchill, J. (1987). Proceedings of the Royal Society A, 412, 269}283.
O'Neill, M. E. (1968). Chemical Engineering Science, 23, 1293}1298.
Reeks, M. W., & Hall, D. (1988). Deposition and resuspension of gas-borne particles in recirculating turbulent #ows.
Journal of Fluids Engineering, 110, 165}171.
Reeks, M. W., Reed, J. R., & Hall, D. (1988). The resuspension of small particles by a turbulent #ow. Journal of Physics D,
21, 574}589.
Reed, J. R., & Rochowiak, P. (1988). The adhesion of small particles to a surface. Proceedings of the second conference of
the aerosol society (p. 229). Oxford: Pergamon.
M.W. Reeks, D. Hall / Aerosol Science 32 (2001) 1}31 31

Vainshtein, P., Ziskind, G., Fichman, M., & Gut"nger, C. (1997). Kinetic model of particle resuspension by a drag force.
Physical Review Letters, 78(3), 551}554.
Wang, H. C. (1990). E!ects of inceptive motion on particle detachment from surfaces. Aerosol Science and Technology, 13,
386}393.
Wen, H. Y., & Kasper, G. (1989). On the kinetics of particle re-entrainment from surfaces. Journal of Aerosol Science, 20,
483}498.
Ziskind, G., Fichman, M., & Gut"nger, C. (1995). Resuspension of particulate from surfaces to turbulent #ows } review
and analysis. Journal of Aerosol Science, 26(4), 613}644.
Ziskind, G., Fichman, M., & Gut"nger, C. (1997). Adhesion moment model for estimating particle detachment from
a surface. Journal of Aerosol Science, 28(4), 623}634.

You might also like