You are on page 1of 40

Index

1. Introduction ...................................................................................................... 2
2. Model for electrostatic precipitators ................................................................ 4
2.1. Gas properties................................................................................................................. 4
2.2. Estimating the corona current ........................................................................................ 5
2.3. Electrical properties of the electrostatic precipitator...................................................... 5
2.4. Particle charging ............................................................................................................. 6
2.4.1. Field charging .............................................................................................................. 6
2.4.2. Diffusion charging .................................................................................................. 7
2.5. Drift velocity and collection efficiency ..................................................................... 8

3. Centrifugal forces in a flow field ................................................................... 11


4. Process identification using simultaneous heat and mass transfer ................ 14
5. Inertial separation based on turbulent flows .................................................. 20
6. The influence of a falling water film.............................................................. 26
7. Ionic wind Literature review and experiments ........................................... 29
7.1. Literature review .......................................................................................................... 29
7.2. Experimental detection of ionic wind .......................................................................... 31

8. Particle diffusion ............................................................................................ 34


9. Particle agglomeration.................................................................................... 36
10. Properties of dust in the paper industry ....................................................... 39
11. Nonspherical particles .................................................................................. 41
12. CFD Simulations .......................................................................................... 43
13. Experiments .................................................................................................. 52
13.1. Electrostatic collection efficiency .............................................................................. 52
13.2. Determination of a regression model for the new equipment .................................... 53
13.3. Experiments validating the collection efficiency for the new geometry .................... 56

14. Model for combined electrical and centrifugal air cleaning in turbulent
flows ................................................................................................................... 57
14.1. MODEL 1 for the air cleaner...................................................................................... 57
14.2. MODEL 2 for the air cleaner...................................................................................... 59
14.3. MODEL 3 for the air cleaner...................................................................................... 59
14.4. Experiments and validation of models ....................................................................... 60

15. Conclusions .................................................................................................. 62


References .......................................................................................................... 64
Appendix 1 ......................................................................................................... 66

1. Introduction
Air cleaning in the paper industry has its specific problems requiring some extraordinary precautions and equipment. For instance, since the dust is inflammable and even explosive, a normal bag
filter has to be placed outdoors, leading to excessive piping etc. A wet scrubber on the other hand
consumes significant amounts of water and power.
This is why a concept based on an electrostatic precipitator with a wetted collection surface, where
the charging electrode consists of rings with needles on, was developed (see figure 1). A project at
the VTT Manufacturing Technology Center was carried out in order to investigate the electrostatic
cleaning properties in such an air cleaner. From the beginning the concept was like a normal
electrostatic air cleaner, where the air is lead through the cleaner as calmly as possible. In the paper
industry, however, the air can contain large pieces of paper which eventually lead to the clogging of
the charging electrode at the centre of the air cleaner. To avoid this clogging, the inlet was changed
to a tangential position; see figure 1 for a better description.
This project was started in order to establish whether the conventional model for electrostatic collection can be used under such highly turbulent conditions, and how forces other than electrostatic
ones should be considered. A mathematical model to determine the collection efficiency for different air cleaner geometries in varying paper production or conversion industries was also set as a
goal. A Masters thesis was produced within the project, to determine the properties and distribution functions for the dust present in different industrial environments (Dahlbacka 1999). The project
lasted from 1.2.1998 to 31.5.1999.

Figure 1: Full scale electrostatic air cleaner. The reactor height is 5 m (see figure 6). The
charging electrode with its rings can be seen at the center, with the reactor surface acting
as the collection surface. The equipment is situated at the Mets-Tissues factory in Mntt,
Finland.

The experiments described in chapters 4 and 7 are all conducted on the apparatus in Mntt figure
1. Later on during the project a new, slightly different apparatus was built, to be used for experimental purposes, by Valmet Corp. Air Systems for the project. The new design was based on measurements from the first apparatus, preliminary calculations from this project and the CFD calculations described in chapter 12. The new air cleaner (see figure 36) was built lower, which was one
fundamental goal for the development of the cleaner.
In chapter 13 some different experiments are described. One important part is the one certifying the
electrostatic collection efficiency model for a full scale air cleaner.

2. Model for electrostatic precipitators


Electrostatic precipitators are well covered by literature and therefore it should not be necessary to
discuss these things so thoroughly in this context. A mathematical model based on a wire-in-tube
electrostatic precipitator is presented together with other equations necessary for solving a practical
problem numerically. For more detailed information see e.g. (Lehtimki 1998a).
2.1. Gas properties
The mobility of a gas ion, Zi , is defined as the ratio of ionic drift velocity i to the electric field E,
i.e.

i = Zi E

(2.1)

The theory dealing with the mobility of gas ions is complicated. A reasonable approximation for the
ion mobility can be expressed as follows:

Zi = (

3e
1
1 1/2
1
)(
+
) (2 k T )1/2

16 p m g mi

(2.2)

where p is the gas pressure, mg is the mass of the gas molecule, mi is the ionic mass, e is the
elementary charge (e = 1.6021 10-19 As) and is the mean impact cross section. The mean impact
cross section can be determined from the following polynomial fit curve:

= a mi 2 + b mi + c

(2.3)

where the coefficients a, b and c have the following values: a = 1.94 10 m/amu, b = 3.62 10-21 m/
amu and c = 1.01 10-18 m.
The diffusional motion of gas ions is characterised with the diffusion coefficient Di which is related to the ion mobility Zi by

Di =

kT
Zi
e

(2.4)

where k is the Boltzmann constant (k = l.3807 10-23 J/K) and T is the absolute temperature.
The mean thermal speed of gas ions is calculated from a formula describing the mean thermal speed
in the case of an ideal gas. The mean thermal speed is given by
4

ci = (

8 k T 1/2
)
mi

(2.5)

2.2. Estimating the corona current


The electrostatic cleaning efficiency, calculated as described in the following sections, requires as
an input parameter the corona current. If the calculations are performed for an existing air cleaner,
a measured value for the corona current is normally used, but in the design phase, at least an approximate value for the corona current has to be calculated. The corona current depends, naturally,
on several factors e.g. the design of the discharge electrode and high voltage. The expression presented here can be used for the electrode configuration as seen in figure 1 (Lehtimki 1998b):

r L 2
2
U
I = K 0 Zi
(r r )3
2 x

(2.6)

In equation (2.6) r2 is the radius of the air cleaner, rx is the radius of the discharge electrode and U
is the onset voltage. The factor K is based on experiments . A set of data points for K is graphically
presented by Lehtimki, and to these points a curve can be fitted. Here the following expression for
K is suggested:

L / nr

K = 1.6278
r2 rx

0.468

(2.7)

In this equation L is the length of the electrostatic collection section and nr is the number of electrode rings. At high voltages, or with a small distance between the electrode rings, it would seem
that (2.6) would give somewhat too low values for the corona current but the expression is anyhow
used to estimate the corona current.
2.3. Electrical properties of the electrostatic precipitator
The electric field between the discharge electrode and the collection electrode plays an important
role in the particle charging and collection processes. The electric field depends on the voltage
applied between the electrodes. The electric field and ion concentration are modelled by assuming
a simple wire and pipe geometry, i.e. a thin corona wire on the centre line of the cylindrical collection tube.
The corona current per unit length of the corona electrode, jL ,is given by

jL =

I
= 2 e Z i n(r) E(r) r
L
5

(2.8)

For high corona currents and for r >> r1 , the electric field can be estimated by

jL
E0
2 0 Z i

1/ 2

(2.9)

An approximate solution for the average ion concentration, n , between the electrodes can be calculated:

0 j
L
n2
2
2 e Z
i

1/2

1
r2

(2.10)

This solution can be used in the idealised case, i.e. the space charge due to aerosol particles is
assumed to be zero. It has also been assumed that there is no dust layer on the collection electrode
that could affect the electrical behavior of the system. This, naturally, is always the case when the
walls are constantly washed with water.
2.4. Particle charging
Particle charging has been studied for several decades and therefore relatively good models are
available. There are two basic mechanisms responsible for the charging of aerosol particles. These
are referred to as field and diffusion charging. Particle charging in an electric field is assumed to be
due to the ordered motion of ions under the influence of an electric field. This approach is applicable predominantly for large particles (dp > 0.5 m). Diffusion charging is due to ion attachment to
particles caused by the random motion of the ions. Particle charging in an electrostatic precipitator
takes place in a strong electric field and at high ion concentration. Thus both charging mechanisms
must be taken into account.
2.4.1. Field charging
Particle charging in a strong electric field can be described by the classical field-charging theory
(Hinds 1982). According to the field charging theory the mean charge number of a particle is given
by

t
s F ( t ) = s0
t + ts
where the saturation charge number is given by

(2.11)

s0 =

3 0
e

r 2
d p E0

r + 2

(2.12)

and the time constant ts is given by

ts =

4 0
n e Zi

(2.13)

and t is the particle charging time calculated from the geometry and the mean flow-velocity. The
constant r is the dielectric constant for the particle material.
2.4.2. Diffusion charging
Ion attachment to aerosol particles can also be caused by the thermal motion of ions (Brownian
motion). This is called diffusion charging and it is the dominating process for particles smaller than
0.2 0.5 m (depending on source). Due to diffusion charging, the maximum charging predicted
by (2.12) can be slightly exceeded.
The so-called White equation gives the particle charge as a function of charging time t:

sW ( t ) =

ln 1 + d p d 0 ci n t
d0
4

(2.14)

e2
d0 =
2 0 k T

(2.15)

dp

where the coefficient d0 is given by

When modeling particle charging by corona discharge, both field and diffusion charging have to be
taken into account. The simplest way of doing this is by combining the two mechanisms:
s = sW ( t ) + s F ( t )

(2.16)

This equation can be used to estimate the total particle charge in an electrostatic precipitator.

2.5. Drift velocity and collection efficiency


Electrostatic force and aerodynamic forces govern the motion of a charged aerosol particle in a gas.
The electrostatic force FE caused by the electric field E0 is given by
(2.17)

FE = s e E0

where s is the particle charge number and e is the elementary charge. This force causes a particle
drift in the direction of the electric field. The electrostatic force is balanced by the gas resistance
force or drag force, which is given by

F D=CD

d p 2 w 2 = 3 w d p
8 g

(2.18)

(the middle part of which, is the general form of Newtons resistance equation and the right hand
part is known as Stokes law) where CD is the drag coefficient, g is the gas density, dp is the particle
diameter and w is the velocity of the particle (Hinds l982). The drag coefficient can be approximated by

24
CD
Re p

2/3

1 + Re p

(2.19)

where Rep is the particle Reynolds number given by

Re p =

g wd p

(2.20)

At mechanical equilibrium the forces FE = FD and the drifting velocity w for a charged particle in an
electric field can be calculated by

Re p
s e E C c
w=
1+
3 d p
6

2/3

(2.21)

where Cc is the Cunningham slip correction factor (Hinds l982) the value of which is given by

Cc = 1 +

d p

A + B exp - C

d p

(2.22)

where is the mean free path of the gas molecules (for air = 6.53 10-8 m). The values of the
coefficients A, B and C are A = 2.5l4, B = 0.8 and C = 0.55 (values for A, B and C vary slightly
depending on the source). The Cunningham slip correction factor extends the range of Stokes law
to below 0.01 m (without correction it is valid only down to 1 m). Thus small particles, using the
slip correction, settle faster by a factor of Cc than predicted by Stokes law. This correction is
needed when the size of a particle approaches the mean free path of the gas molecules because the
particle is so small that it starts slipping between the gas molecules.
In order to predict the collection (or cleaning) efficiency of the electrostatic precipitator the assumption of the complete mixing of air in the collection system is made, i.e. that no gradients exist
in the particle concentration. This should, for a totally turbulent flow, be rather a good assumption.
The collection efficiency is then given by
w A
= 1 - exp

V&

(2.23)

where A is the collecting area (i.e. the area of the collecting electrode at the length of the charging
electrode) and V& is the air flow. The equation is known as the Deutsch formula (Hinds 1982). The
equation (2.23) is based on the assumption that no particle re-entrainment takes place. This assumption is valid for liquid particles or if a falling water film wets the collection surface. As the
drifting velocity w depends on particle size, the collection efficiency is of course also different for
different particle sizes. As several phenomena are examined in this context, the collection efficiency given by (2.23) is hereafter referred to as e and the corresponding velocity as we. A typical
collection efficiency curve is shown below:

Collection efficiency (%) as a function of particle size


100 %

Efficiency [%]

80 %
60 %
40 %
20 %
0%
0.001

0.010

0.100

1.000

10.000

100.000

Particle diameter [m]

Figure 2: Collection efficiency for an electrostatic precipitator as a function of particle


size. The calculations have been made for a system with the following dimensions:
Flow rate V& =3.0 m/s
Length of collection section L=2.6 m
Diameter of the collector tube d=1.6 m
Corona current I=3.2 mA

10

3. Centrifugal forces in a flow field


A rotational air flow will invoke a centrifugal force throwing airborne particles outwards in the
flow field. This property is commonly utilised in, for example, cyclones. The airflow is brought
into rotation by, for instance, bringing the air tangentially into a larger vessel. The vessel studied in
this particular case could well be described as a cylindrical reactor (for cyclones it would then be a
conical reactor). A tangential inlet, as described earlier, is assumed in this study (see figure 1).
In order to determine the centrifugal collection efficiency, the centrifugal particle velocity wc has to
be determined. Assuming that the centrifugal force acting on a particle is equal to the drag force
acting on the same particle (see equation 2.18) the radial velocity of the particle is calculated by
(Ogawa 1984)

d p 2 p g r 2
wc =
18

(3.1)

where is the angular velocity, g and p are the gas and particle densities respectively and r is the
radius of the particle path in the flow. This radius is growing with time since the particle is slung
outwards by the centrifugal force. Assuming that no other forces tend to drag the particle towards
the center of the reactor, a simplification is done by using an average value for the particle path
radius, r, calculated from the centerline of the reactor to the centerline of the tangential inlet. Since
the measure of the inlet is rather small compared to the reactor radius, using an average value for r
does not drastically affect the result of equation (3.1).
Equation (3.1) should be completed with the Cunningham slip correction factor

d p 2 p g r 2 C c
wc =
18

(3.2)

The meaning of this correction factor was discussed in chapter 2.


Since the inlet and outlet not are necessarily equal in size and shape, it is therefore justifiable to
assume that the angular velocity varies with the length (or height) of the reactor. Calculating the
angular velocities in and out for the inlet and outlet from the velocities in the connecting pipes
(depending on geometry it might be necessary to use only the tangential velocity component) from

in =

win
rr

wout
and out = r
r

gives the angular velocity as a function of reactor height:


11

(3.3)


in
h
( h ) = in + out
hr

(3.4)

where rr marks the total reactor radius (3.3) and hr the reactor height (3.4). In a later section these
assumptions can be compared to the results from CFD simulations. Equations (3.l) and (3.2) hold,
only if the Reynolds particle number is small (equation 2.20), i.e. if the relative velocity between
the particle and the fluid is small enough, whilst only then the drag force CD can be given by the
simpler form

CD

24
Re p

(3.5)

Considering that Rep might be significantly higher than one, equation (2.19) should be used. Thus
equation (3.2) has to be modified to:

h2

d p 2 p g r 2 C c
dh

Re 2/3
h1
p

18 1 +
6

wc =
h2 h1

(3.6)

A mean value for the velocity wc is obtained from equation (3.6). In equation (3.6) the equation for
(3.4) is inserted. An expression similar to the Deutsch formula can be used to calculate the
centrifugal collection efficiency

w A
c = 1 - exp c
V&

(3.7)

Since the centrifugal particle velocity wc depends on the square of the particle diameter, the result of
equation (3.7) will depend heavily on the particle size. A typical collection-efficiency curve is
shown below (figure 4).
These calculations, as well as the equations presented above, are made under the assumption that
the angular velocity changes linearly through the collector. The angular velocity at the inlet is
calculated from the mean velocity win in the inlet pipe ( wmax win 1.1K1.2 ), see figure 3, using
the total radius of the collector rr instead of the actual radius of the particle path r (assuming wmax at
the centre line of the inlet). The same principle goes for the outlet. The value for win obtained this
way is smaller than it would be using wmax and the radius r. Thus a compensation for wall losses is
automatically obtained, giving a flow profile more according to the somewhat hypothetical one
inserted in Figure 3. The profile can be compared with the ones given by the CFD calculations
(chapter 12).
12

wmax

win

rr

Figure 3: A drawing describing the assumptions made when calculating


the collection efficiency for the centrifugal flow.

Collection efficiency (%) as a function of particle size


100 %

Efficiency [%]

80 %
60 %
40 %
20 %
0%
0.001

0.010

0.100

1.000

10.000

100.000

Particle diameter [m]

Figure 4: Collection efficiency in a rotating flow field as a function of particle size.


The calculations have been made using the following values:

Length of collection section L=5 m

Diameter of collection section d=1.6 m

win=12.3 m/s in a 0.60.4 m duct

wout=0.5 win

Particle density rp=1000 kg/m

Air properties at 25C

13

4. Process identification using simultaneous heat and mass transfer


This chapter explains how a regression model giving heat- and mass-transfer coefficients for the
studied air cleaner is obtained. The coefficients are needed later when calculating the collection
efficiency for the air cleaner. A more detailed description can be found in (Berg and Altner 1999).
The studied process is described with the following differential energy balance for the control
volume a, b, c and d in figure 5, and gives:
dq& c + m& a i + m& w c wt0 + dm& v c wt0 = m& a ( i + di ) + m& w c w t1

(4.1)

The numerical differentiation of equation (4.1) requires governing equations for both turbulent heat
and mass transfer.

i x i1 x1

i0 x0

1.
6.

i+di
. .
x+dx dmv+mv
t0
t+dt

t
b

.
dq
. 2. t
dm

6.
2.

dqc

3. h0 +tcv

1.

ma

i
x
t

mv t1

d
A

5. di/dx

4. i/ x

Figure 5: Heat and mass transfer through the boundary layer in the cylinder with a falling water
film. The air state change during evaporation is also described by including a principal picture of
a perspectively transformed humidity chart (Soininen 1994), often named the i,x-chart. The control
volume and control surface are described by the dotted lines connecting a,b,c and d.

Heat transfer
Heat-flow through a boundary layer is usually defined by the following expression, see figure 5:

dq& = (tt )dA

dq&
(tt )dA

(4.2)

We may, in principle, define an average mass flow mn A perpendicular to the surface and consider
that the measured heat flow is the difference between the enthalpies of mass flow away from the
surface, at surface temperature t, and mass flow of the same magnitude towards the surface, at the
main flow temperature t. This gives:
&

14

C&
m&
m&
dq& m& n
c p t n c p t = n c p (tt ) = n (tt )
=
A
A
A
dA A

(4.3)

From a comparison of equations (4.2) and (4.3), we obtain:

m& n
C&
cp = n
A
A

(4.4)

Mass transfer
If the surface in contact with air is the surface of a drying material, the humidity of the flow away
from the surface is higher than that of the flow towards the surface. Denoting the temperature of the
surface with t, the humidity of saturated air at that temperature with x and the humidity of the
main flow with x and replacing mn A combined with a more accurate definition of the specific heat cp
one may write (Berg and Karlsson 1998):
&

(x x)
dm& v m& n
=
( x x) =
dA
A
ca + xcv

(4.5)

where the content of moisture in the air, also known as air humidity x, is defined as:

x=

M v pv
M a p pv

(4.6)

and the air humidity x (kgH2O/kgda) is the saturation moisture content defined by the local control
volumes mass content of water vapour divided by the mass content of dry air.
The content of moisture in the air is defined using Daltons law for ideal gas combined with Antoines
water vapour equilibrium curve. Here the notation is used for the mass-transfer coefficient instead of , that is, the heat-transfer coefficient. The stress mark indicates that we now refer to the
mass transfer which occurs simultaneously to heat transfer. Equation (4.5) is known in the literature
as the Lewis equation when k1 in equation (4.7) equals one.

= k1 when k1 1

(4.7)

Let us now examine the experimental process by looking at figures 5 and 6. Convective heat transfer is the product of the heat capacity mass flow rate of air and the temperature drop of the air.
Whereas mass transfer takes place at the same time, we write that this convective heat flow is
(t t)dA . Further, when we consider that m& a dx represents evaporation dm& v , equation (4.1) can
15

be rewritten as:

(h0 + tcv tc w )

m&
dm& v 1
dm& v
)dt
+ (t t) k c (t c t) = c w ( w +
dA dA 2
dA

(4.8)

It is now possible to solve equation (4.8) numerically when assuming k1=1 and solving an approximate value kc 4 W/m2C from the free convection theory (Jakob 1949).
Surface temperature t according to the Lewis analogy
During heat and mass transfer, the state of the air is changed from point 1 to point 6 shown as the di/
dx line B in figure 5. The Lewis analogy surface temperature t is obtained, if the difference between curve A and curve B in figure 5 is neglected. This is the case when k1=1 and thus along this
line the direction of the air state change is indicated by the ratio i/x and can be expressed with the
following statement.

i di
i
ii
=

=
x dx
x x x

(4.9)

The Lewis analogy surface temperature t, is therefore, always different from the measured surface
temperature when k1 1 . The assumptions that are related to equations (4.8 and 4.9) are considered, at this stage of the work, as the method basis when determining . It is often very
informative to critically analyse the process by drawing the experimental wet and dry bulb temperatures in an i,x-diagram, and thereby obtain a figurative description showing if the assumed air state
changes differ from the experimental ones.
The experiments were carried out on a full-scale apparatus, shown in figure 6. In order to obtain the
heat- and mass-transfer coefficients ( and ) for the apparatus using the theory presented, some
process variables were measured (see table 1 below): dry and wet bulb temperatures of both the
inlet and outlet air flows, as well as flow velocities. The flow velocities were measured with a Pitot
tube and a digital micromanometer. The inlet temperatures were measured in the vertical duct (see
figure 6) about 2 metres before the duct entered the cylinder body and outlet temperatures at a point
about 2 metres from the cylinder exit. The temperature of the surrounding hall was measured on
several occasions, and it could be considered constant (tc=25 C). The heat-transfer coefficient
from the surrounding hall air to the water film was calculated as 4 W/mC. Table 1 also shows the
air humidity, calculated from dry and wet temperatures.
The experiments were carried out during the cold season, with very dry air. In this range of dry and
wet temperatures, the influence of daily variations in total air pressure can, in some cases, be considered negligible when calculating differences in air humidities, x. The influence of the absolute
pressure p0 on the experimental results should, however, be checked before continuing with or
without correction.

16

wout
dout=0.4 m

Starting point of
falling water film

dc=1.6 m
See figure 1 for
enlargement
win
din=0.5 m

h4.4 m

Tangential inlet

Figure 6: Drawing showing the principles of the apparatus used for


experiments.

Measured inlet
Velocity [m/s].

Measured Dry
Temperature at
inlet [C]

Measured Wet
Temperature at
inlet [C]

Measured Dry
Temperature at
outlet [C]

Measured Wet
Temperature at
outlet [C]

Psychrometric
Air Humidity at
inlet
[kgH2O/kgda]

Psychrometric
Air Humidity at
outlet
[kgH2O/kgda]

Table 1: Experimental data.

15.1
7.5
10.6
14.6
7.3
6.4
14.8
10.8
6.9
14.7

26.2
26.5
26.3
26.1
26.6
24.5
23.6
24.5
24.6
24.5

14.4
14.2
13.8
13.5
14
12.1
11.1
12.2
12.5
12.9

21.9
22.2
22.5
22
22.3
20.1
20.2
20.1
20.6
20.6

14.2
14
13.9
13.8
13.8
11.8
12.1
11.8
12.9
13.1

0.00538
0.00504
0.0047
0.00447
0.00479
0.0037
0.0031
0.0038
0.00406
0.00451

0.00693
0.00659
0.00636
0.00646
0.00634
0.0052
0.00546
0.0052
0.0061
0.00631

Using the numerical data, together with the equations presented, gives the possibility of calculating
. Iterative calculations are carried out as follows: is given a value, after which the outlet dry
temperature and the outlet air humidity are calculated. The calculated values are then compared
with the measured ones, and the -value adjusted until a satisfactory result is obtained. The main
goal is to match the dry temperatures, as the measured dry temperature values can be expected to be
the most reliable ones. As one can see from the results in table 2, the agreement of the measurements is good, and even the calculated air humidity values (which were mentioned as being of
secondary interest at this point) give good agreement. In addition, the inlet temperature of the
falling water film had to be adjusted (note: water inlet in the upper region of the cylinder).

17

Measured Dry
Temperature at inlet [C]

Measured Dry
Temperature at outlet [C]

Calculated Dry
Temperature at outlet [C]

Difference [%]

Psychrometric Air
Humidity at outlet
[kgH2O/kgda]

Calculated Air Humidity at


outlet [kgH2O/kgda]

Difference [%]

Resulting Transfer
coefficient [W/mK]

Table 2: Experimental and calculated values. The transfer coefficients are adjusted
to an accuracy of 0.5 W/mK.

26.2
26.5
26.3
26.1
26.6
24.5
23.6
24.5
24.6
24.5

21.9
22.2
22.5
22
22.3
20.1
20.2
20.1
20.6
20.6

21.86
22.18
22.5
21.95
22.31
20.13
20.19
20.12
20.61
20.57

-0.19
-0.1
0
-0.21
0.04
0.15
-0.05
0.08
0.05
-0.16

0.00693
0.00659
0.00636
0.00646
0.00634
0.0052
0.00546
0.0052
0.0061
0.00631

0.00691
0.00656
0.00635
0.00643
0.00631
0.00519
0.00544
0.0052
0.00608
0.00627

-0.34
-0.55
-0.24
-0.59
-0.57
-0.22
-0.33
0.01
-0.43
-0.57

69.5
32.5
42.5
67.5
30.5
27.5
69
45.5
33.5
70

When looking at the small differences between the measured and the calculated values for the dry
temperatures and air humidity, the conclusion is that the agreement seems to be acceptable. The
relation between heat and mass transfer can be expressed as = k1 where, in this particular case,
k11. From the resulting heat- and mass-transfer coefficients presented in table 2 the Nusselt and
Reynolds numbers are calculated from their respective characteristic equation:

(4.10)

wm d

(4.11)

Nu =

Re =

For separate flow geometries, the characteristic flow velocity (w) and the characteristic length unit
(d) will be defined differently. Here wm is defined as the mean axial velocity in the cylinder and d is
the cylinder diameter (dc in figure 6). The experimentally obtained Nusselt numbers, as a function
of Reynolds numbers, can be seen in figure 7. A curve is, thereafter, fitted to the experimentally
obtained Nusselt numbers. As can be seen, a fairly good agreement is reached using a straight line.
This would give the relation:
Nu = 0.027 Re

(4.12a)

which can be transformed into a more familiar form:

Nu = 0.032 Re 1 Pr 0.45

18

(4.12b)

Equation (4.12a) is presented graphically in figure 7. The tangential flow pattern seems to give high
Nusselt numbers compared to Reynolds numbers.

Nusselt number as a function of Reynolds number


5000
4000
3000
Nu
2000
1000
0
0

50000

100000

150000

200000

Re
Figure 7: The experimentally obtained Nusselt numbers as a function of
Reynolds numbers. The dotted line marks the equation (4.12a).

The values of in the cylinder might be considered to be higher than expected. One could, however, compare these values with the values calculated for the inlet pipe diameter and the flow
velocity of the air. The expression for Nu in turbulent flow is (VDI-Wrmeatlas 1953):

d 2 / 3
Rem b Prm c
Nu = a 1 +

(4.13)

where a = 0.024, b = 0.786 and c = 0.45 for turbulent flows in straight circular tubes. In equation
(4.13) the physical properties of air should be calculated using arithmetic mean values, i.e. tm=(t1+t2)/
2, xm=(x1+x2)/2. For the cylinder under observation the ratio l/d is 4.2 m/1.6 m 2.6. Noting that the
constant in front of Re in (4.12a and 4.12b) describes the development of the temperature or velocity profile, where in this case thermal entrance effects are significant, the same constant together
with the mentioned l/d, should be used in (4.13) when comparing the levels of heat- and masstransfer coefficients.
Thus the level of turbulence (i.e. the level of heat and mass transfer) in the cylinder, depends both
on the inlet geometry and the kinetic energy of the flow in the inlet pipe. Hence, this is the most
plausible explanation for the remarkably high exponent b in Reb. Values of b=0.80.95 can be
found in the pertinent literature for turbulent flows over rough surfaces and impingement drying
(e.g. Nunner 1956 and Mujumdar 1987, pp. 461-474).
Some more experiments and thoughts around regression models for the studied air cleaner can be
found in chapters 7 (same geometry as here) and in 13 (a new geometry).
19

5. Inertial separation based on turbulent flows


The deposition of particles in turbulent flows has been discussed by many authors (for instance
Ogawa 1984 and Friedlander 1977). One physical interpretation of the deposition is that fine solid
particles with a large inertial force caused by the fluctuating velocity of the turbulent flow, in the
fully turbulent flow region, penetrate the quiescent region near the wall and reach the wall surface.
In other words, the stopping distance of a particle is compared with the thickness of the laminar
sublayer.
Considering a pipe flow, as Friedlander did, gives the stopping distance for a particle, based on the
radial fluctuating velocity of turbulent flow, wr:

wr = 0.9 w

f
2

(5.1)

The equation is the result of experiments (Laufer 1953). Here w is the mean gas-velocity in the pipe
and f is known as the fanning friction factor, which is given by the Blasius equation (Bennett and
Meyers 1982)

f =

0.0791
Re 0.25

(5.2)

Using the fluctuating velocity wr given by equation (5.1) gives the stopping distance for a particle
according to the Stokes equation

S=

p d p 2 wr
18

(5.3)

Friedlander gives two separate equations for the transport (or migration) velocity of particles in the
turbulent flow. When the stopping distance is much smaller than the thickness of the viscous sublayer, the following expression gives the deposition velocity

k=

2 d p 2 p 2U 5 f 5 / 2

6.1 10 5 4 2

(5.4)

where k is the transport velocity, f is the fanning friction factor and U is the average gas velocity.
The upper limit of this equation is given as

20

SU f

2

1/ 2

<5

(5.5)

On the other hand, when the stopping distance is much larger than the thickness of the buffer layer,
the same type of model is applicable. The deposition velocity is given by

k=

1 2S / d
f
U
25 ln (d / S ) 1.73 2

1/ 2

(5.6)

The lower and upper limits of equation (5.6) are given as

Transport velocity as a function of Reynolds number


1000

[cm/min]

B.
100

A.

2 m
5 m
8 m

10

1
1000

10000

100000

1000000

Re

Figure 8: Transport velocity (according to Friedlander) in a turbulent pipe flow as a


function of Reynolds number. Pipe diameter d= 25 mm and particle density r =7800
kg/m (iron). Iron is used to make the comparison with figure 11 easier.

S
SU f
< 0.5 and

d
2

1/ 2

> 30

(5.7)

In figure 8 the group of lines denoted as A are calculated with equation (5.4), while those denoted
as B represent equation (5.6). The lines B are not drawn to the upper limit S/d < 0.5. Already,
before reaching the upper limit the empirical equation (5.6) tends to collapse (i.e. the transport
velocity falls dramatically). The theory, presented above, gives the equations for two extreme situations. The discontinuity of the lines (see figure 8) also causes problems when modelling the cleaning efficiency for turbulent flows.
21

In order to obtain a continuous function and to be able to secure equations for varying geometries,
a slightly different method though based on the same principles as described above is used. The
stopping distance is compared with the thickness of the laminar sublayer (Berg 1998):

k12
g w1

y1 =

(5.8)

where k1=2.5 and w1 gives the flow velocity at the distance y1 from the wall:

0 k12
g

w1 =

(5.9)

which gives a critical diameter, dp,crit

d p ,crit =

18 2 k1 2
wr g p w1

(5.10)

Particle fractions larger than dp,crit are assumed to deposit totally, and consequently no particles
smaller than dp,crit would deposit. Since this is experimentally verified as being false, the function is
made continuous by proportioning the stopping distance of an observed particle to the stopping
distance of the critical particle size. A constant kp is introduced:

kp =

S
S crit

(5.11)

where Scrit is the stopping distance for the particle having the diameter dp,crit. The maximum value of
kp is 1, since the turbulent eddies cannot transport particles faster than the transporting velocity (it
was earlier stated that particles larger than dp,crit are assumed to deposit totally). See also later sections.
The velocity of the particles towards the collection surface is determined by the flows which are
normal to the pipe surface. The heat- and mass-transfer coefficient for a system is, in fact, the
average heat-capacity rate of flows occurring in the direction normal to the surface (Soininen 1994).
For turbulent flows in pipes is given by (VDI-Wrmeatlas 1953):

0.024 Re 0.786 Pr 0.45


d
22

(5.12)

where the physical properties of air (in Re and Pr) should be calculated using arithmetic mean
values, i.e. tm=(t1+t2)/2, xm=(x1+x2)/2. Thus the velocity for the turbulent flow (i.e. eddies) in the
direction normal to the surface is given by (Soininen 1994):

wn =

cp g

(5.13)

The velocity wn can be said to transport particles towards the collection surface in the air cleaner,
whilst the stopping distance is determined by wr (Friedlander 1977). The product k p wn is equivalent to the transport velocity of the particles. This way, the upper limit of kp given above, also
becomes natural.

Transport velocity as a function of Reynolds number

[cm/min]

1000

100

2 m
5 m
8 m

10

1
1000

10000

100000

1000000

Re

Figure 9: Particle transport velocity in a turbulent pipe flow as a function of


Reynolds number. Pipe diameter d=25 mm and particle density r=7800 kg/m.
The critical particle size is obtained by comparing equations (5.3) and (5.8).

Since the radial fluctuating velocity wr given by equation (5.1) actually gives the velocity of the gas
just outside the buffer layer, the stopping distance should be compared with the thickness of the
laminar and the buffer layer, not just the laminar sublayer as above. The thickness of the laminar
and buffer layer, y2, is given by (Berg 1998)

y2 =

k1k 2
g w1

(k2=22.5) and therefore the expression giving the critical particle size becomes

23

(5.14)

18 2 k1k 2
wr g p w1

d p ,crit =

(5.15)

Figures 9 and 10 can be said to describe two extreme situations, the first with S=y1 and the second
with S= y2. The actual case will probably be somewhere between these two values, which also can
be seen from figure 11 showing some experimental results (Ogawa 1984). Unfortunately these
experimental values are covered by equation (5.4) and never go into the region covered by equation
(5.6).

Transport velocity as a function of Reynolds number

[cm/min]

1000

100

5 m
8 m
2 m

10

1
1000

10000

100000

1000000

Re

Figure 10: Particle transport velocity in a turbulent pipe flow as a function of


Reynolds number. Pipe diameter d=25 mm and particle density r =7800 kg/
m. The critical particle size is obtained by comparing equations (5.3) and (5.14).

Figure 11: The deposition of fine solid


particles. Pipe diameter d = 25 mm.
Experiments presented in (Ogawa 1984).

In agreement with the assumption of no re-entrainment, made in the Deutsch formula, equation
(2.23), the collection efficiency of the turbulent deposition is given by
24

k p wn A
T = 1 exp

V&

(5.16)

Collection efficiency (%) as a function of particle size


100 %

Efficiency [%]

80 %
A

60 %

B
40 %

20 %
0%
0.001

0.01

0.1

10

100

Particle diameter [m]

Figure 12: Collection efficiency in a turbulent flow field as a function of particle


size. Using equation (5.10) for the critical particle size gives (B), using equation
(5.15) gives (A and C). The difference between A and C is explained below.
The calculations have been made with the following values:
Pipe diameter d=0.5 m
Flow velocity w=15.1 m/s
Particle density rp=1000 kg/m
Collection area A=25.1 m

Since the velocity of the fluctuating velocity, wr, is given just outside the buffer layer, the transporting velocity, wn also should be calculated at the same location. Equation (5.13) gives wn as a function of the overall heat- and mass-transfer coefficient. It can easily be shown (Berg 1998) that the
heat- and mass-transfer coefficient for the turbulent core can be expressed as:

3 = 2 K 3 tot

(5.17)

In figure 12, C is calculated using 2.5 tot in equation (5.13).


The maximum level for collection efficiency is consequently determined by the transporting velocity wn, whereas the point at which the maximum level is reached, is determined by the thickness of
the laminar (y1) or laminar and buffer layer (y2). Using 3 for (wn) and y2 is, at this point, considered
the most logical way since wr is used to determine the stopping distance and by definition wr is the
fluctuating velocity at the location where 3 describes the heat and mass transfer rate.
The values used to calculate the lines in figure 12 correspond to the inlet pipe in the studied air
cleaner (Figure 1). The collection area is the area of the air cleaner. This calculation is made just to
get an overall picture of the turbulent collection efficiency. The selection of geometrical parameters
will be discussed more in detail in a later section.
25

6. The influence of a falling water film


The collection surface of the air cleaner is constantly washed with water falling down as a thin film.
The water is applied to the uppermost region of the cleaner (see figure 13) along the whole width of
the wall. To estimate the properties of the falling water film, some equations from (Bird et al l960)
are used.

Figure 13: Flow of liquid film under the influence of gravity with no rippling. Slice
of thickness Dx over which the momentum balance is made. The y-axis is pointing
outward from the plane of the paper.

The maximum velocity of the falling film (Figure 13) is the velocity furthest away from the wall,
and can be calculated by

wmax =

g 2 cos
2

(6.1)

where is the fluid (water) density, is the wall angle and is the film thickness. The integrated
average velocity is given by

wz =

g 2 cos
3

(6.2)

The film thickness may be calculated from the average velocity, the volume rate of flow or from
the mass rate of flow per unit width of the wall. Here, the equation for volume rate of flow only, is
given, for the others please see (Bird et al 1960).
26

=3

3 Q
gW cos

(6.3)

where Q is the volume rate of flow and W is the wall width (perpendicular to the falling direction).

Film velocity [m/s

Film velocity as a function of film temperature


0.7
0.6
0.5
0.4
0.3
0.2
0

20

40

60

80

Film temperature [C]


Q=20 l/min
Q=30 l/min

Figure 14: Velocity of the falling water film for two flow rates. Cylinder
diameter d =1.6 m.

For instance, having a wall width W = 5.0 m (i.e. d = 1.6 m) and Q = 20 litres/min would give a
maximum velocity wmax = 0.34 m/s (fluid properties at 12 C), with a film thickness = 0.30 mm. A
higher temperature results in a higher film velocity and a thinner film. Thus it should be considered
which property to keep constant in the air cleaner with seasonal temperature changes the flow
field or the film thickness.
When considering the flow profile of the upward flowing air in the air cleaner, the following question arises: how does a water film with an opposite direction influence the profile? Closest to the
water surface, the air will have the same velocity and direction as the water (the velocity of which
will slightly diminish from the value given by equation (6.1) under the influence of the air. This is,
however, ignored in this context).
Since particles drift towards the collecting surface as they travel through the cleaner, they will, at
some point, enter the region assumed to have a velocity directed downwards. This should, slightly,
influence the cleaning efficiency in a positive manner. Particles entering this region can be assumed
to be separated since their relaxation time reaches infinity. The thickness of the laminar and buffer
layer (y2) can be diminished to represent only the part of y2 being directed upwards. The following
modification is suggested to the theory presented in chapter 5: The velocity of the falling water
film, equation (6.1) is added to the axial velocity component of the main flow. The critical particle
size is calculated by comparing equation (5.3) with the thickness of the reduced laminar and buffer
layer given by

27

w
wmax

y 2+ = y 2 2 max
w2 max

(6.4)

(which is slightly less than y2) where wmax is given by (6.1) and w2max is calculated from the original
w2 increased with wmax.
Changing the thickness of the laminar and buffer layer (from y2 to y2+) will slightly change w2 (and
consequently w2max). The radial velocity component is, however, dominant compared to the axial
component considered here. The result is that minor inaccuracies in the value of w2 will not significantly affect the second factor at the right hand side of equation (6.4). This somewhat simplified
procedure results in a minor error of less than 1% in equation (6.4).

28

7. Ionic wind Literature review and experiments


In an electrostatic precipitator a continuous stream of ions is established, due to the potential difference between the charging and collecting electrodes. Because of this stream of ions, surrounding
gas molecules are also brought into motion in the same direction. This movement of gas (air mostly)
is often called ionic wind (or corona wind, electric wind). In this text a short summary of some
papers on the subject is presented first. Later, some experiments to evaluate ionic wind are reported. A more extensive analysis of the experiments in chapter 7.2 is found in (Berg and Altner
1999).
7.1. Literature review
As a summation of the papers studied, it can be said that they all imply that ionic wind gives a
higher rate of turbulence in the precipitator. Whether this is a positive or a negative factor when
considering collection efficiency, varies from one paper to another.
Zhibin & Guoquan (1994) for instance, reach the conclusion that when increasing the corona current in relation to the flow velocity, the collection efficiency increases, but when increasing the
current over a certain point diminishes the collection efficiency. Furthermore, they state that a
reduction of turbulence results in a higher collection efficiency. An explanation for this maximum in collection efficiency, when varying the corona current, could be that the ionic wind increases the level of turbulence. Increasing the turbulence only a little improves the collection efficiency, but when the corona current is further increased, re-entrainment occurs, resulting in a lowered collection efficiency. Strictly speaking, the statement made by Zhibin & Guoquan, that a reduction of turbulence gives a higher collection efficiency, would mean that ionic wind is a negative
factor for the collection efficiency.
Leonard et al (1982) found experimentally that the intensity of the corona-induced turbulence was
comparable to or less than the background turbulence present, without an electrical field for gas
velocities exceeding 1.5 m/s, but increased rapidly when compared to background turbulence for
decreasing gas velocities. Generally, it is assumed that there will be a progressive deterioration in
the effective migration velocity at increasing flow velocities due to re-entrainment. According to
(Leonard et al 1982), to achieve improved precipitator performance with a negative corona through
turbulence control, the gas velocity should be approximately equal to or larger than 1.5 m/s, i.e. the
ratio of corona-induced turbulence to background turbulence should become lower. This means
that the ionic wind could be seen as a negative factor with regard to collection efficiency. Both
(Zhibin & Guoquan) and (Leonard et al) seems to have made their experiments with dry collection
surfaces. The fact that re-entrainment occurs at higher flow velocities is valid only for dry surfaces
with a wet surface no re-entrainment occurs.
Kercher (1969) made some interesting observations in his experiments. First of all, in quiescent air,
no ionic wind could be observed for a wire discharge electrode. Some needles or points had to be
present at the electrode before any ionic wind could be measured. For a system with a 43 kV voltage
(and needle electrode), increasing the distance, between the charging and collecting electrode, to
not more than 120 mm, decreased the ionic wind to a minor level. For air in motion, a system with
50 kV was used. When the flow velocity was 0.5 m/s the ionic wind was still clearly measurable,
but when the flow velocity was increased to 1.2 m/s no ionic wind could be detected. Overall the
conclusion made by Kercher is that ionic wind has a positive effect on collection efficiency in an
29

electrostatic precipitator because of the corona-induced turbulence pointing towards the collecting
surface. It is especially beneficial when considering small particles that easily form space charge.
Due to the turbulence, the space charge is displaced towards the collecting surface leading to only
a minimal reduction in the corona current. The experiments presented by Laufer were carried out
with clean air only.
Liang and Lin (1994) gave a ratio between the flow velocity and the characteristic velocity of ionic
wind, described as:

Ue =

ew V0

(7.1)

where is the gas density, ew is the charge density and V0 is the applied voltage. They concluded
that when the ratio was less than 0.2, turbulent kinetic energy increased significantly due to ionic
wind, thus diminishing the collecting efficiency. When the ratio was over 0.5 the electric force
acting upon the flow field was not prominent, and the kinetic energy of the flow was near that of a
fully developed flow without an electric field. Furthermore, the authors stated that under normal
operating conditions, the influence of ionic wind on the flow field was insignificant. Decreasing the
flow velocity (i.e. decreasing the ratio mentioned to below 0.2) the ionic wind became more obvious and had a negative effect on the collection efficiency. The results were theoretical only, and the
reduction in collection efficiency seemed to be due to re-entrainment caused by turbulence. Again
it can be said that this does not apply to wet surfaces.
Shu & Lai (1995) in their paper, studied improved heat transfer due to an electric field. The electric
field results in a secondary air-flow (i.e. ionic wind) directed towards a grounded surface, enhancing the heat-transfer in the system. The net effect of this secondary flow is an additional mixing of
the flow and destabilisation of the thermal boundary layer leading to raised heat-transfer coefficients. The enhancement is most significant regarding small air-flows, giving a heat-transfer several times higher than one without electric field. The results of Shu & Lai are theoretical. They
further suggest that the enhancement in heat-transfer is possible due to the oscillatory flow-field.
Whenever two flows, being of the same order of magnitude are opposed or perpendicular to each
other, some degree of oscillation is to be expected.
Of the papers cited here discussing ionic wind and particle collection, the only one implying an
improvement in particle collection efficiency due to ionic wind is Kercher (1969), and that regardless of whether it is a wet or dry collecting surface. A wet surface might influence the conclusions
of the other studies.
The paper of Shu & Lai gives an opportunity to consider the results in chapter 4, and also to try to
measure the ionic wind in the air cleaner studied in this project. The heat- and mass-transfer coefficients are measured with and without an electric field, and any difference in levels would then be
explained by the ionic wind. In the following sections some experimental results are presented.

30

7.2. Experimental detection of ionic wind


The experiments were made using the equipment described in chapters 1 and 4. A more detailed
discussion about regression models and also about ionic wind can be found in (Berg and Altner
1999). The regression model presented in this chapter is an improvement to that presented in chapter 4. See also chapter 13.
The experimental results were entered into the overall energy balance for the system, giving as a
result, the heat- and mass-transfer coefficient . Thereafter, a regression model of the same type as
before is generated. The energy and mass balances give the following equations from which t2 and
are calculated. Solving the equations, assuming a steady temperature t1 = 6 C and calculating
the Reynolds and Nusselt numbers in accordance with equations (4.10) and (4.11), we obtained a
set of points which are shown in figure 15.

t
q& el + (m& a x 2 m& a x1 ) c t c Atot + c 1 Atot m& w t1 c w
2
t 2 =
A
c tot m& w c w
2

(7.2)

and

x1 + x2

cv + q& el
2

t1 + t 2 t1 + t 2

Atot
2
2

(t1 t 2 ) m& a ca +

(7.3)

where t represents the temperature in the vicinity of the evaporating surface, the lower indexes 1
and 2 denote the properties at the inlet and outlet, and q& el is a small amount of electrical losses
added as heat to the system, measured in form of a corona current. In this case, the corona current
comes from the fact that the air is lead through a passage between two high voltage electrodes
(Ogawa 1984). The overall effect of the corona current was in this case measured as q& el 400 W.
Here q& el is added completely to the air since the resistance of air, compared to other resistances is
of a different order of magnitude. Anyhow, in this case the influence of q& el 400 W is so small that
it could just as well be omitted. Furthermore using q& el in equation (7.3) assures a maximum influence of the input electrical effect, in the form of ionic wind, on the heat- and mass-transfer coefficient, and yet no measurable differences can be detected.

31

Nusselt number as a function of Reynolds number


5000
4500
4000
3500

Nu

3000
2500
2000
1500
1000
500
0
0

50000

100000

150000

200000

250000

Re
Experiments (without electricity)

Experiments (with electricity)

Regression modell

Figure 15: The experimentally obtained Nusselt numbers as a function of the


Reynolds numbers and a regression model (equation 7.4a, shown as a drawn line)
fitted to the Nusselt numbers.

From the data points in figure 15 no significant difference can be seen between those cases with or
without an electric field, i.e. Nusselt numbers and therefore can be held as equal in both cases.
Shu & Lai (1995) amongst others, wrote about an additional mixing of the gas due to the applied
high voltage (i.e. ionic wind), resulting in an enhancement in the heat-transfer, but they also stated
that the influence would be greatest at small flow velocities (i.e. low Reynolds numbers). In the
case under study (the system was described earlier in chapters 1 and 4) the level of turbulence is
extremely high and even a voltage as high as 130 kV (i.e. q& el 400 W), should not cause measurable differences in the heat- and mass-transfer coefficients.
From these data points the regression model becomes:

Nu = 0.032 Re0.95

(7.4a)

Nu = 0.037 Re 0.95 Pr 0.45

(7.4b)

or, to use a more familiar form

Considering the ratio d/l and using the more general equation (4.13) gives for (7.4b)

d 2 / 3
Re 0.95 Pr 0.45
Nu = 0.024 1 +
l

32

(7.5)

When comparing these equations with the equations (4.12a) and (4.12b), one can see that the exponent of Re does not equal one, as suggested in the previous section, but is a little bit lower, which
was already predicted as the most probable case. Some further experiments giving regression models of the same kind as equations (7.4a) and (7.4b) are described in chapter 13.

33

8. Particle diffusion
In order to evaluate the need to consider particle diffusion when modeling an air cleaner of the kind
studied during this project, some equations are presented. The diffusion coefficient for a particle is
given by (Hinds 1982)

D=

k T Cc
3 d p

(8.1)

where k is the Boltzmanns constant, T is the absolute temperature and Cc is the Cunninghams
correction factor (see chapter 2). The equation is called the Stokes-Einstein equation for aerosol
particle diffusion coefficients.
For deposition on a collecting surface by diffusion from a turbulent flow, the situation is complicated, and no explicit solution exists. The customary assumption made, is that the turbulent flow
provides a constant concentration, n0, which is uniform beyond a thin diffusive boundary layer
(the same assumption concerning the concentration was made earlier for the Deutsch formula). In
this thin layer, the concentration decreases from the concentration of the main flow to zero at the
surface. The deposition velocity is given by

Vdep =

(8.2)

where is the thickness of the diffusion boundary layer. The value of depends on flow mechanisms, the nature of the velocity boundary layer and the size of the particles. One suggestion for
found in (Hinds 1982, page 148) is

28.5 d t D 1 / 4

g

7 / 8

Re

1/ 4

(8.3)

where dt is the tube diameter. Larger particles get thrown partway into the diffusion boundary layer
and have a shorter distance to diffuse, whereas smaller particles, having smaller inertia, have to
diffuse the whole distance. To express the diffusive deposition in terms of collection efficiency, the
following is given

Vdep A

D = 1 exp

&
V

34

(8.4)

In the equation (8.4) A is the collection surface area. For decreasing particle sizes the diffusional
effects become more pronounced, whilst for larger particles, inertial effects increasing with particle
size, govern collection efficiency. In between, a gap appears where neither of the processes work
properly. This gap in the efficiency curve can be seen for particles in the 0.1 to 1.0 m range.
When calculating the collection efficiency for diffusional deposition in a turbulent flow for the air
cleaner studied in this project, a circular tube with a flow, giving an equal level of turbulence (i.e.
-value) is chosen in order to achieve collection efficiencies of the right magnitude. Therefore a
pipe of d= 0.5 m with the flow velocity wg=15 m/s is used. To get approximately the same residence
time as in the air cleaner, a tube length L52 m is used, resulting in a collection surface area A 82
m.

Collection efficiency (%) as a funtion of particle size

Efficiency [%]

100
80
60
40
20
0
0.001

0.01

0.1

10

100

Particle diameter [m]

Figure 16: Collection efficiency for deposition by diffusion in a turbulent pipe flow.

Analogous to diffusive and field charging, chapter 2, where equations (2.9) and (2.12) were added,
equations (5.16) and (8.4) could be combined to one - the error being very small because of the
above-mentioned gap in the collection efficiency. The resulting equation would become

k p wn + Vdep A

T + D = 1 exp

&
V

(8.5)

The influence of the diffusional deposition, however, is hereafter neglected since, as seen in figure
16, it is significant in a particle size range with very little practical interest to the paper industry.

35

9. Particle agglomeration
The phenomena studied so far in this text are all collecting particles from the gas flow regardless of
particle concentration. For particle agglomeration (or coagulation) this is different as agglomeration occures due to collisions between particles. No mass is actually separated from the gas flow
due to agglomeration, but the number of particles is reduced.
The theories for particle agglomeration assume that when two particles collide with one another
due to a relative motion between them, they attach to each other and form larger particles. The size
of the particle growing with the collision, is not affected that much, but for the other particle the
change is much more dramatic it vanishes completely. The result is a decrease in number concentration, and as the agglomeration is more significant for small particles, the number size distribution is also displaced towards larger particles.
The mechanisms for particle collisions are various, and some idealistic models can be found in the
pertinant literature. When the relative motion between particles is due to Brownian motion the
process is called thermal agglomeration (or coagulation), whereas when external forces are involved, the process is called kinetic agglomeration. Such forces are gravity, electrical forces or
turbulent eddies.
The simplest case is thermal agglomeration of monodisperse spherical aerosols. The theory is based
on how particles diffuse to the surface of an observed particle. The rate of collisions, which is equal
to the rate of change in number concentration, can be given by (Hinds 1982):

dN
= 4 d p D N 2
dt

(9.1)

where D is the diffusion coefficient for particles (equation 8.1). If the change in number concentration with time due to agglomeration is to be described, the following equation is obtained by integrating equation (9.1):

N( t ) =

N0
1 + N0 4 d p D t

(9.2)

where N0 is the number concentration at t=0. The diffusion coefficient D decreases rapidly with
increasing particle size. Therefore N(t) will differ from N0 only for small particles. The same can be
said when the time period is short. As an example, the net result of agglomeration can be neglected
over a 10-minute period if the number concentration is less than 10 6/cm. Since the residence time
in the studied air cleaner is given in terms of seconds rather than minutes, and since the number
concentration reaches 106/cm only for the smallest particles, the influence of (thermal) agglomeration (of monodisperse spherical aerosols) can be neglected in the paper industry. The curve Number distr out in figure 17 represents equation (9.2) with t=5 seconds.
For polydisperse aerosols the rate of agglomeration is significantly higher. Small particles having a
36

large diffusion coefficient D diffuse well to larger particles, having a higher surface area. Thus
agglomeration between particles of different size will affect the number size distribution faster. The
bigger difference in size, the faster the agglomeration will proceed. For the mass size distribution
shown in figure 17, which should represent a typical distribution curve for paper converting dust,
having a total mass concentration of 21 mg/m, the increased agglomeration for polydisperse aerosols will not significantly change the number size distribution curve coming out of the system.

2.5

1.E+08

2.0

1.E+04
1.E+02

1.5

1.E+00
1.E-02

1.0

1.E-04
1.E-06

0.5

Mass distribution [mg/m]

Number distribution [1/cm]

1.E+06

1.E-08
1.E-10
0.001

0.01

0.1

10

100

0.0
1000

Particle diameter [ m]
Number distr in

Number distr out

Mass distr

Figure 17: Mass and number distributions for an aerosol (rp=1000 kg/m, spherical
particles). The total mass concentration in this example is 21 mg/m.

When observing an airflow, there will always be velocity gradients which result in different velocities for particles. Equations can be found for both laminar and turbulent flows, only the latter being
of practical interest. As the turbulent eddies result in significant velocity gradients, the following
equation is given:

2
3
Turbulent agglomeration b d p g 2 f U
=

Thermal agglomeration 64 D d t

(9.3)

where b is a constant of order 10, U is the average velocity in the duct with the diameter dt and f is
the fanning friction factor (equation 5.2). Equation (9.3) gives the ratio of turbulent agglomeration
to thermal agglomeration. With U =15 m/s and dt = 0.5 m the ratio becomes one (i.e. turbulent
agglomeration exceeds thermal agglomeration) for approximately 0.7 m, and increases rapidly
with increasing particle size. However, as the number concentration decreases rapidly with increasing particle size, the conclusion drawn from figure 17 will not change.
For the agglomeration of charged particles, particles of opposite charges would attract each other
and form agglomerates (clusters) more easily. The net result is a minor change compared to the
agglomeration of uncharged particles, as the effect is balanced by collisions between particles of
37

same sign repelling each other.


In an electrostatic air cleaner the particles all have a negative (or positive) charge (possibly forming
dipoles) and therefore the net result should be, not just equal to agglomeration of uncharged particles, but even lower.
The conclusion is, for the air cleaner studied in this project with paper dust, that the number of
particles is too low for significant agglometation, no matter which theory for agglomeration is
considered. The conclusion is the same even when the particles are charged.

38

10. Properties of dust in the paper industry


The different mechanisms for particle collection described in earlier chapters, all to some extent,
include properties of the particles, and the material of the particles. The particle size, density, dielectric constant and particle shape for individual particles must be known and since the collection
efficiency depends heavily on particle size, the particle size distribution function for the processed
dust must be known in order to properly estimate the final (mass) collection efficiency of the planned
air cleaning system.
Some of the properties are described here in brief, but the shape factor and some aspects concerning
that are presented in a later chapter. The properties presented are based on the Masters thesis
undertaken within the project (Dahlbacka 1999).
Particle size distribution can be split up in three separate fractions; the particles with the smallest
diameter are referred to as the nucleous fraction, the next is called the accumulation fraction and the
largest particles are defined as the coarse fraction. Whether the sizes of particles match any definition for the names used is irrelevant. The names are just used to indicate which fraction is being
discussed.
Since the physical properties of the particles vary for the different fractions, the collection efficiency as a function of particle size will also be different for the different fractions. The three
different fractions have to be treated separately and if a final curve giving the collection efficiency
as a function of particle size for the whole particle distribution is desired it can, for example, be
calculated from the difference between input and output dust. This might lead to some unlinearities
in the collection efficiency curve.
The nucleous and accumulation fractions are treated as ideal spheres, the coarse fraction as fibers
described as cylinders (Dahlbacka 1999). A study was made in order to determine the particle size
distribution function for dust from different paper qualities in different paper production or conversion processes. The qualities were coated paper, newsprint and tissue paper. The relative size distributions are shown in figure 18.
For the different distributions the following values are used (table 3), based on the study mentioned
and literature:

39

N
uc
le
ou
s
A
cc
um
ul
at
io
n
Co
ar
se

Table 3 : Particle properties.

r
4
Tissue paper
p
640
k
0.0085

0.4289

1.7
Newsprint
p
640
k
0.0278

0.4456

1.79
Coated paper
p
640
k
0.0818

0.485

1.81

2.5

950
0.261
10.9
2.51

800
0.731
174.8
76.9

kg/m
m
m

850
0.255
6.55
2.34

800
0.717
174.8
76.9

kg/m
m
m

790
0.567
6.27
2.21

800
0.351
174.8
76.9

kg/m
m
m

Relative particle size distribution for different paper


qualities
18.0 %
16.0 %

Fraction [%]

14.0 %
12.0 %
10.0 %
8.0 %
6.0 %
4.0 %
2.0 %
0.0 %
0.1

10

100

Particle diameter [m]


Tissue paper

Coated paper

Newsprint

Figure 18: Relative particle size distribution for different paper qualities.

40

1000

You might also like