Sie sind auf Seite 1von 8

This article was downloaded by: [Indian School of Mines]

On: 08 January 2014, At: 18:15


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: Mortimer
House, 37-41 Mortimer Street, London W1T 3JH, UK

Journal of the Air Pollution Control Association


Publication details, including instructions for authors and subscription information:
http://www.tandfonline.com/loi/uawm16

Electrostatic Precipitation Of Fly Ash


Harry J. White

Consultant , Carmel , California , USA


Published online: 13 Mar 2012.

To cite this article: Harry J. White (1977) Electrostatic Precipitation Of Fly Ash, Journal of the Air Pollution Control
Association, 27:1, 15-22, DOI: 10.1080/00022470.1977.10470386
To link to this article: http://dx.doi.org/10.1080/00022470.1977.10470386

PLEASE SCROLL DOWN FOR ARTICLE


Taylor & Francis makes every effort to ensure the accuracy of all the information (the Content) contained
in the publications on our platform. However, Taylor & Francis, our agents, and our licensors make no
representations or warranties whatsoever as to the accuracy, completeness, or suitability for any purpose
of the Content. Any opinions and views expressed in this publication are the opinions and views of the
authors, and are not the views of or endorsed by Taylor & Francis. The accuracy of the Content should not
be relied upon and should be independently verified with primary sources of information. Taylor and Francis
shall not be liable for any losses, actions, claims, proceedings, demands, costs, expenses, damages, and
other liabilities whatsoever or howsoever caused arising directly or indirectly in connection with, in relation
to or arising out of the use of the Content.
This article may be used for research, teaching, and private study purposes. Any substantial or systematic
reproduction, redistribution, reselling, loan, sub-licensing, systematic supply, or distribution in any
form to anyone is expressly forbidden. Terms & Conditions of access and use can be found at http://
www.tandfonline.com/page/terms-and-conditions

Harry J. White

Downloaded by [Indian School of Mines] at 18:15 08 January 2014

Consultant, Carmel, California

Rapid growth of the power industry and the need for ever
increasing power generation efficiency led in about 1920 to the
successful development of pulverized coal as a basic fuel for
large steam generating units in the United States. The pulverized coal process for power generation was pioneered in
Milwaukee, Wisconsin, and soon gained recognition as a major
technical advance over the old stoker-fired power boilers.1 The
new firing method brought in its wake, however, severe and
complex gas cleaning problems. In this form of firing, the residual ash from the coal is carried in the furnace gas stream
in the form of suspended finely divided particles which are
emitted from the stack unless effective collection means are
provided. These minute fly ash particles may comprise 10%
or more of the mass of the coal burned, and therefore would
constitute an intolerable nuisance unless removed from the
gas stream before emission from the stack.
In the power generation industry, application of electrostatic precipitation to the fly ash removal problem is almost
as old as the use of pulverized coal, with the first precipitator
going into service in 1923 at the Trenton Channel plant of
Detroit Edison.2'3 Various inertial and mechanical collectors,
such as wet scrubbers and cyclone collectors, had been tried
but proved ineffectual in solving the severe air pollution
problem inherent in the pulverized coal firing process. The
first precipitator installation at the Trenton Channel plant
consisted of three units specified for 90% collection efficiency
and treating a total gas flow of 800,000 acfm. A period of several years was required to overcome the many problems encountered before achieving successful operation, but with this
practical achievement wide acceptance of the precipitator for
high efficiency removal of fly ash was assured in the power
industry. Today a total of more than 1300 fly ash precipitator
installations having a rated gas flow of over 500 million acfm
have been made in the United States.
The extent of the fly ash problem is shown strikingly by the
amount of coal being burned by the utilities, which in 1975
totaled over 400 million tons. The corresponding quantity of
fly ash which could be emitted would be about 30 million tons
annually, assuming an average ash content of 10% and an
average retention of ash in the furnace of 30%. The magnitude
of the problem of fly ash emission control for a large modern
power plant may be illustrated by considering a coal-fired
generating unit producing, for example, 600 Mw electric
output. Such a unit typically will exhaust flue gas at the rate

of 2.5 million acfm at a temperature of 300 F. Using a typical


ash loading of 5 grains/scf at the air preheater outlet, the dust
emitted from the furnace will amount to about 1200 lb/min,
or 800 ton/day. To deal effectively with this problem, collecting equipment must be provided to clean continuously 2.5
million acfm of hot flue gas with an efficiency of over 99%, and
with ash disposal means for some 800 ton/day.
The basic features which have made electrostatic separators
particularly attractive for this gasvcleaning problem are: (1)
they can be designed to provide high collection efficiency for
all sizes of particles from submicroscopic to the largest present;
(2) they are economical in operation because of low internal
power requirements and inherently low draft loss; (3) they can
treat very large gas flows; (4) they are very flexible in gas
temperatures used, ranging in the power field from as low as
200F to as high as 800F; and (5) they have long useful
life.
The basic purpose of this paper is to present a broad fundamental and practical discussion of the important phases of
fly ash precipitator application and technology. Subjects
covered are:
1.
2.
3.
4.
5.
6.
7.
8.

Outlook for future growth.


Fundamental operating principles.
Fly ash and furnace gas characteristics.
Precipitator design.
Hot precipitators.
Precipitator equipment.
Precipitator problems.
Case histories of precipitator installations.
These are examined in the context of the pace setting requirements of the federal Clean Air Act of 1970 which established new and much more severe standards both for performance and reliability of air pollution control equipment for
power plants and other heavy industry. Fundamental advances in the science and technology as well as procurement
methods for precipitators for high efficiency control of fly ash
emissions have emerged under the force of the legislation. It
is hoped that the paper will clarify these advances and provide
guidance for air pollution control engineers generally, as well
as specifics for those directly concerned with the ongoing
application and performance problems of electrostatic precipitators for collection of fly ash under today's complex
conditions.

Dr. White, recipient of the Frank A. Chambers Award for "Outstanding achievement in the science
and art of pollution control," is now a consultant in Carmel, CA. He is the author of several definitive
books on air pollution control.
This paper will be published in four parts in JAPCA in the coming months. Sections 1 and 2, as
identified above, are in this issue. Section 3 will appear in February, Sections 4 and 5 in March, and
Sections 6, 7, and 8'will conclude the presentation in April.

January 1977

Volume 27, No. 1

15

Outlook for Future Growth

Continued rapid growth of electrostatic precipitation in the power

2000

industry is anticipated because of increasing dependence on coal for


power generation and the stringent environmental requirements

1750 -

mandated by clean air legislation. New levels of scientific understanding and technical sophistication in the precipitation field will be
necessary to successfully meet the performance and reliability

=1500 -

challenges inherent in this unprecedented growth.

2 1250 -

Downloaded by [Indian School of Mines] at 18:15 08 January 2014

1000 -

Future expansion of the power industry and continued


dependence on coal as a major energy source for electric power
production are obviously important factors in assessing the
future growth of fly ash precipitation. Past growth of the
electric power industry in this country has been remarkable,
with power production doubling about every ten years since
1900, and the use of electricity underlying many of our material advances. In the past few years there has been some
slowing in the rate of increase of electric power production
because of economic conditions and the uncertainties and
disruptions of the energy crises, but the long-term outlook
appears to be for continued steady growth, although possibly
at a somewhat slower rate than in the past.
The major energy source for electric power production in
the U. S. has been and continues to be coal. Other energy
sources for power production include petroleum, natural gas,
hydro, and more recently nuclear reactors and geothermal.
Recurring crises and shortages of petroleum and natural gas,
as well as the swirling controversies associated with these
fuels, point to decreasing use for power generation in the
future.
Most hydropower sites in this country either have been
developed already or are unlikely to be available because
of environmental restrictions.
Predictions of the growth of nuclear power have proved to
be over optimistic in the past because of the serious technical, economic, environmental, and political problems
encountered. There seems to be little reason at present to
expect that nuclear power will account for more than a small
percentage of the total power needed for years to come.
Geothermal energy seems limited to relatively small
amounts in selected geographical areas.
Other approaches to the energy problem, such as solar energy, are still in the early experimental stage and offer little
promise for substantial commercial use for perhaps decades.
Therefore, the huge coal resources of the U. S., coupled with
the manifold problems associated with alternate energy
sources, make it almost certain that coal will remain the major
fuel for electric power production in this country for many
years to come. The idea of converting coal to a clean fuel by
removing ash and sulfur before burning, thereby rendering
cleaning of the combustion gases unnecessary, is attractive,
but proposed methods for accomplishing this appear to be
only in the experimental stage at present.
Displacement of electrical precipitation by other gas
cleaning methods for fly ash collection is, of course, possible,
but entails many difficult problems including high con16

H 750 500 -

250 -

1920

1930

1940

1950
Year

1960

1970

1980

Figure 1. Electric energy production by public utilities in the


United States, 1920-1975.

sumption of energy, achieving the reliability needed for power


plant use, and attaining high-efficiency removal of submicron
particles which increasingly are being indicted as injurious to
health.
Turning now to quantitative data, Figure 1 shows the total
electric energy production per annum in the U. S. for the period 1920 to the present, and also the amount produced from

1920

1930

1940

1950
Year

1960

1970

1980

Figure 2. Coal consumption by public utilities in the United


States, 1920-1975.

Journal of the Air Pollution Control Association

99.9

99.8
i

_ 99.7

99.5
S-99.3 -

Maximurn,/

8 98. "97 95
93
90
1920

100 -

1920

1930 1940 1950 1960 1970


Year

1980

Downloaded by [Indian School of Mines] at 18:15 08 January 2014

Figure 3. Growth in installed fly ash precipitator capacity in


the United States.

coal for the same period. The amount produced from coal has
consistently comprised about one-half of the total power
generated. Coal consumption per annum by the public utilities
is shown in Figure 2. Reasonable projections of these historic
growth patterns support the expectation for continued high
coal consumption for power production in the United
States.
Data on the installed fly ash precipitation capacity in this
country and the trend in collection efficiencies, beginning with
the first application in the early 1920's, are plotted in Figures
3 and 4. Particularly notable are the very rapid increase in cfm
capacity and the trend toward ultra-high efficiencies, which
have occurred in the past few years. Collection efficiencies for
the earliest precipitators were 90%, reaching about 99% average value in 1970, and about 99.8% maximum value currently. The rapid increase in collection efficiency in recent
years is directly attributable to the air pollution control requirements stemming from the federal clean air legislation
of 1967 and 1970.4
The overall growth of fly ash precipitation in the U. S. is
even more remarkable when both the increases in the cfm
capacity and in collection efficiency are taken into account.
The higher efficiency requirements are especially important
since the increase in precipitator size is nonlinear with efficiency. For example, a fly ash precipitator for 99% efficiency
typically will be three or four times larger than one of similar
capacity for 90% efficiency. Taking both size and efficiency
factors into account, the per annum growth rate for new fly
ash precipitators turns out to be some 20 times greater for the
1965-70 era than for the 1945-50 era. This rapid growth is
evident in Figure 5 which shows the annual sales of precipitators in the United States for the period 1966-1975, of which
about three-fourths is accounted for by fly ash precipitators.
Precipitator development has been influenced in several
important ways beyond increased use and higher collection
efficiencies by the federal clean air legislation and emission
standards.5* First, the requirements for meeting emission
standards are backed up by enforcement provisions which can
curtail plant output or even force complete shutdown of
generating units if such is necessary for compliance with
emission standards. The practical result of these enforcement

January 1977

Volume 27, No. 1

/j

^r

/ Average

1930 1940 1950 1960 1970 1980


Year

Figure 4. Trends in the average and maximum collection efficiencies of fly ash precipitators in the United States.

provisions has been to require much more conservatively designed precipitators and to compel higher design and construction standards than were previously acceptable in the
precipitator industry.
Second, the emission limits on sulfur oxides5 have caused
the power industry to turn toward much greater use of low
sulfur coals as a means of meeting these stringent limitations.
This in turn complicates the precipitator problem because of
the greater probability of high resistivity ash with these coals.
Although effective methods are known for dealing with high
resistivity, these methods require a higher level of technical
sophistication and result in increased costs as compared with
fly ash collection from higher sulfur coals.
A third complication can arise from the need to provide
effective removal of both the fly ash and the SO2 from the flue
gas. This introduces several new factors in the overall air
pollution control system for a power plant, and also several
different control strategies may need to be considered. The
technical and economic feasibility of the various possible
strategies have yet to be firmly established. However, it now
appears that dry collection of the fly ash in a precipitator, either before or after the air preheaters, followed by removal of
the SO2 in a wet scrubber may turn out to be the preferred
method.

500

< 100 -

1966
* Particulate emissions for new generating units are limited under the standards to a maximum of 0.1 lb/million Btu heat input, equivalent to about 0.08 gr/scf for a typical fly ash,
and to visible emissions not exceeding Ringelmann 1. Collection efficiencies of 99+% are
usually necessary to meet these standards, with some allowance for safety margins.

/K

1968

1970
1972
Year

1974

1976

Figure 5. Annual sales of precipitators in the United States for


1966-1975.

17

Fundamental Operating Principles

Downloaded by [Indian School of Mines] at 18:15 08 January 2014

Purpose of this section is to set forth the basic principles and theory
underlying the precipitation process, which are necessary for understanding the design and operation of fly ash precipitators. Included
are the migration velocities of particles, the idealized efficiency
equation, and the effects on performance of nonuniform particle size
and other nonuniformities. Semiempirical and statistical methods
sometimes used for precipitator design are also examined.

Basically, an electrical precipitator must provide three essential functions: (1) the suspended particles must be given
an electrical charge; (2) they must be subjected to an electric
field to remove them from the gas stream to a suitable collecting electrode; and (3) means must be provided for removing the particle layers from the electrode surfaces to an
outside receptacle with as little loss as possible.
In practice, electric charging of the particles is accomplished
by means of ions produced in the high voltage d-c corona.
Under typical precipitator operating conditions the charge
attained on a 1 micron particle, for example, is about 250
elementary charges.
In a Cottrell or single-stage precipitator the collecting field
is also provided by the high voltage d-c corona. Typical
values of the collecting field are 4-5 kv/cm, and the electric
force acting o n a l / i particle, for example, is about 3000
times gravity. It should be noted that the existence of these
very high collection forces, especially for the finer particles
which are the most difficult to collect, is one of the major
factors which set electrostatic precipitators apart from other
methods of particle collection. Furthermore, the forces are
exerted directly on the particles themselves, whereas in
mechanical collecting methods it is necessary to treat the
whole mass of gas in order to exert the necessary separating
forces on the particles. As a result, electrostatic precipitators
are inherently economical in power consumption and are
free of the high draft losses which characterize most mechanical or inertial collection methods.
In a Cottrell precipitator the actual power required to clean
one million cfm of flue gas at 99%, for example, turns out
to be only about 200 kw, as compared to over 2000 kw consumed by a typical scrubber for the same application.
Removal of the collected particle layers is accomplished by
rapping (by impact, or by vibration, of the electrodes). The
rapping process is critical to overall performance, as a balance must be struck between keeping the electrodes sufficiently clean, and overrapping which causes excessive loss
of already collected particles by reentrainment into the gas
stream.

conditions. The simplest case to consider is that of uniform


conditions for particle size, gas flow, and electric field distribution. Extensions of the theory can be made to include
non-uniformities for these factors. However, the disturbances
and particle losses caused by reentrainment, gas sneakage, and
rapping cannot be calculated from theory, but must be accounted for by modifications of the parameters based on
empirical data from performance tests on existing precipitator
installations. In practice, the losses due to rapping and the like
should be minimized insofar as possible in order to achieve
best performance and so that the design can be based essentially on the theoretical equations.
Gas flow in industrial precipitators is always turbulent and
the migration velocities for particles below about 5 /im in diameter are small compared to the gas velocities normally used
in precipitators. Under these conditions, the motions of the
particles are dominated by the tortuous gas flow patterns
characteristic of turbulence. Particle collection occurs only
when particles move close enough to be trapped by the electric
field near a collecting electrode surface. The theoretical
problem of particle capture under these conditions can be
treated by well-known classical probability theory,6 and leads
to the equation
= l -.

e-(A/V)w

(1)

where t] is the probability of particle capture, V the gas flow


rate, A the collecting electrode surface area, and w the migration velocity of the particle.* This equation was first derived by Deutsch in 1922 using a differential equation method,7 and was discovered even earlier by Anderson in 1919 in
an empirical form based on field experiments.8
Derivation of theoretical expressions for the collection of
large particles that move in the electric field with velocities
comparable to or exceeding the gas velocity is much more
difficult and no such solution exists at present. However, this
is of little practical importance since chief interest in electrostatic precipitation centers on the fine particles and it is
easy to show that the larger particles will be captured at a rate
greater than that predicted by the exponential formula.
Migration Velocity.

This may be calculated from theory by equating the electrostatic force on the particle to the drag force, giving the
relation *
6irda

(2)

Here, q is the particle charge, Ep the precipitating field, 6 the


gas viscosity, a the particle radius, A the mean free-path of the
gas molecules, and a a dimensionless parameter found to be
about 0.86 for atmospheric air.
Eq. (2) is accurate to within a few percent for smooth
spherical particles of radii 0.05 txm < a < 30 /xm in air. The

Efficiency Equations

The fundamental theory of electrostatic precipitation is


sufficiently advanced to permit derivation of theoretical
equations for the collection efficiency under certain idealized
18

* The quantity eo appears in the expression for q and in Eq.(3) because of the use of S.I. units,
and has the value 8.854 X 10~12 farad/meter.
t The units in Eq.(2) may be either cgs electrostatic (esu) or Systeme Internationale (SI).
The latter system is used herein in conformity with the present vogue. Thus w is expressed
in meters/sec.

Journal of the Air Pollution Control Association

Downloaded by [Indian School of Mines] at 18:15 08 January 2014

term (1 + a X/a) is the Cunningham correction to Stokes' law.


The mean free-path X depends on the composition, pressure,
and temperature of the gaseous medium. For air at normal
pressure of 760 mm Hg, typical values of X are found to be
0.069 /t at 70F, 0.106 at 300F, and 0.128 at 600F. The mean
free-path varies inversely with the gas pressure, so that at
one-half normal pressure, for example, the values of X would
be doubled. It is evident that the Cunningham correction
factor is significant for particles up to several microns diameter for gas temperatures of 300 F and higher, the correction
increasing with the temperature. Although the composition
of flue gas differs somewhat from that of air, it is sufficiently
accurate for most purposes to use values of w calculated from
the mean free-paths for air.
The resistance to motion of nonspherical particles is not in
general susceptible to theoretical analysis. Particles departing
only slightly from spheres will not be much affected, but
others such as needle or star shaped particles obviously will
be greatly affected. However, since most of the particles
comprising fly ash tend to be spherical in shape, Eq. (2) is a
reasonable approximation for calculation of w.
Calculation of the particle charge q in the high voltage corona is far from elementary. Complications arise because of
the presence of two different particle charging mechanisms
usually identified as field charging9 and diffusion charging.10
In field charging, ions from the corona are driven onto the
particles by the electric field; this is the major charging process
for particles larger than 1 or 2 ;um diameter. In diffusion
charging, the corona ions are carried to the particles by virtue
of their kinetic energies; this process is most important for
submicron particles. In the intermediate size range from a
few-tenths to 1 or 2 nm diameter, both processes are important. Theories of the two basic charging processes are well
understood, but the combination of the two leads to complex
mathematical equations which cannot be solved by analytical
methods. Therefore, recourse must be had to approximate or
to computer methods.11 Results of particle charge calculations
by refined computer methods compare quite well with experimental values.12
For the purposes of this paper it is sufficient, in calculating
the particle charge, to treat the fly ash particles as essentially
conductive. Consider, first, particles larger than 1 or 2 nm
diameter. The saturation particle charge is given by q =
12TreoEca2, and the Cunningham correction factor can be
neglected, so that Eq. (2) simplifies to
(3)

where Ec is the charging field strength.* Since the charge for


particles smaller than 1 or 2 /im cannot be expressed analytically, it is necessary to use numerical values of q, determined
by approximate methods, in Eq. (2) in order to calculate the
migration velocities for these very fine particles. Substitution
of the approximate value of w in Eq. (1) then gives the theoretical collection efficiency for particles of uniform size in
precipitators energized by steady d-c voltage.
Voltage Waveform. Derivation of the fundamental precipitation Eq. (1) is based on steady d-c voltages and fields.
In actual practice, however, steady d-c voltages are seldom
used, but rather unfiltered rectified voltages which fluctuate
in synchronism with the alternating current supply. These
fluctuating voltages obviously influence the migration velocity
of the particles, which in turn affects the collection rate.
However, it is readily shown13 that Eq. (1) remains valid
provided that the average value of the migration velocity wav
is substituted for the d-c value of w. In general, the particle
charge when using fluctuating voltages is determined by the
maximum value of the voltage, whereas the collecting field is

* The exponent (A/V)w in Eq. (1) is dimensionless so that A, V, and w must he expressed
in consistent units. In practice, usage varies between English and metric systems.

January 1977

Volume 27, No. 1

determined by the average voltage. Therefore, it is advantageous to keep both the maximum and average values of the
voltage as high as possible.
Variations in Field over the Collection Surface.

For pipe precipitators, the field distribution over the collection surface is uniform, but for duct precipitators it varies
widely because of the effects of the plate baffles and the grid
arrangement of the corona wires. These variations in the
collecting field obviously affect the theoretical efficiency
equation. Since there is no feasible approach for calculating
the effects of the baffles and end effects of the plates, it is usual
to neglect these factors and simply use an average field based
on average current density at the plates. This is an approximation which can be justified only qualitatively in a theoretical sense, but is adequate in a practical sense because of the
presence of other uncertainties of equal or greater magnitude.
Nonuniform Particle Size

In practice, the typical fly ash is composed of a wide range


of particle sizes from under ljitm to over 100 /im in diameter.
The equation for efficiency for nonuniform particle size distributions may be obtained from Eq. (1) by integration
methods. For this purpose, let y(x) represent the size-frequency particle size distribution function, and let y(x) represent the cumulative distribution function. Then, by definition
dy = y(x) dx
and the precipitator loss dQ for particles in the infinitesimal
size interval x to x + dx is given by
dQ = e-{A/V)w(x) dy = e-(A/V)w(x) 7 ( x ) dx
where w{x) is the migration velocity for particles of size x.
Integration gives for the efficiency
n = l-Q

= l-

(4)

Jo

where the integration limits encompass all particle sizes.


Evaluation of the integral in Eq. (4) can be carried out by
analytical methods only in special cases. Otherwise the integral must be evaluated by numerical or computer methods.
Note that if w(x) can be adequately represented by Eq. (3),
then expression {4) simplifies somewhat. Let k be a parameter
defined by
k=

A epEcEp

v~T-

(5)

Then the efficiency equation can be written in the form


~kx y(x)dx

T? = 1 C

(6)

Jo

which can be evaluated analytically in certain cases of practical interest. The most important case is that for the lognormal particle size distribution which is treated in the next
section.
Log-normal Distribution.

Many fly ashes are found to obey the log-normal particle


size distribution law to a first approximation. Substitution of
the appropriate expressions for y(x) in Eqs. (4) and (6) leads
to explicit formulas for the collection efficiencies for cases in
which the log-normal law is valid.
Let 7i(t) be the size-frequency distribution function for the
normal or Gaussian distribution expressed by the relation
,-tV2
yi(t)

(7)
19

Then the log-normal distribution is derived from Eq. (7) by


the transformation14
In (x)/xs
(8)

where xg is the geometric mean size and ag the geometric


standard deviation.
This leads to the expression for 7(x) as follows
dx
Substitution of Eq. (9) in Eq. (4) gives the efficiency
equation for the log-normal distribution
e-tV2

. e-(A/V)w(x)

,
*
dx
(10)
'o
(v 2TT In ag) x
Similarly, substitution of Eq. (9) in Eq. (6) gives the efficiency
equation for the log-normal case where w is proportional to
the particle diameter

Substituting this result in Eq. (13) and integrating gives for


the efficiency
77 = 1
CVa"e-<Au>'A^f(v)dv
(14)
Jo
where vmax is the maximum value of the gas velocity.
Eq. (14) accounts for the effect of uneven gas velocity on the
efficiency of collection, but does not cover losses in efficiency
due to particle reentrainment. The reentrainment losses
cannot, in general, be calculated from theory, but must be
determined empirically from field measurements and observations. There are various approaches for expressing the
reentrainment loss quantitatively. One useful method is to
express the added loss in terms of lowered values of the effective w in regions where reentrainmeht occurs. Calculation
of total loss, taking into account particle reentrainment, is
then similar to the case treated in Eq. (14), except that w now
becomes a decreasing function of gas velocity for velocities
above the critical particle blow-off value.

Downloaded by [Indian School of Mines] at 18:15 08 January 2014

P-t

/2.p-kx

= 1 - 11

Joo ( v

.
2TT

dx

(11)

Nonuniform Particle Size and Gas Flow

In Gg) x

In practice, both the particle size and the gas flow are
nonuniform, and need to be taken into account in formulating
efficiency equations. It is not difficult to develop the theo2
77 = 1 - - i = f " [e-U /2) . e-kxg . et In cA d t ( 1 2 )
retical equation for this case, but the result involves double
integration and requires a great amount of numerical calcuThe lower limit of integration becomes t = <*> which corre- lation to evaluate practical cases. Such calculations can be
performed by computer methods, but otherwise are too time
sponds to x = 0 in Eq. (11). Evaluation of the integral in exconsuming to be justified. An approximate procedure which
pression (12) by analytic means has been studied by Allander
is usually permissible is to calculate the effects of nonuniform
and Matts.15 They were able to relate the integral to the soparticle size and nonuniform gas flow separately. Such a
called aftereffect function which appears in the classical
procedure may be justified in many cases because of other
Jahnke and Emde Tables of Functions, pp. 38-39. Application
approximations and uncertainties which enter into the design
of this method is made in a later section of this paper.
of fly ash precipitators.
It is sometimes convenient to use t instead of x as the
variable of integration in expression (11),

Nonuniform Gas Flow

Unbalanced gas flow through a precipitator lowers performance in two fundamental ways. First, uneven treatment of
the gas lowers collection efficiency in the high gas velocity
zones to a degree not compensated for in the low velocity
zones. Second, particles already captured may be blown off
the plate surfaces in high gas velocity regions and be lost from
the precipitator. Although each of these adverse effects is
important, the second or reentrainment loss predominates
where gas flow is markedly bad. Positive measures are needed
to minimize these losses, including gas flow model studies to
insure high quality gas flow, and limiting design gas velocities
to 5 or 6 ft/sec.
The effect of nonuniform gas flow in lowering efficiency can
be treated in a manner similar to that for nonuniform particle
size distribution. Let f{v) du be the fraction of the precipitator
inlet cross-section area which has a gas velocity between v and
v + dv. Then
f(v) dv =
where Ac is the total cross-sectional area for gas flow through
the precipitator. It follows that the fractional loss dQ through
the area dAc is given by
dQ = e-tdA'dV)w f(v)
fo
(13)
where dA is the element of collection surface associated with
dAc, and dV is the gas flux through dAc. But dA/dV can be
expressed by
dA = A(dAc/Ac) = A
dV
vdAr
Arv
20

Semi-theoretical and Statistical Methods

There are several approaches for describing precipitator


performance and designing precipitators which are based on
semi-empirical and statistical methods rather than on theory.
In one approach, the concept of nonuniform particle size
distribution is generalized to include variations in gas velocity,
corona current density, and electric field. Computations based
on this concept have been carried out for the case where the
generalized distribution of these variables is assumed to be
represented by the log-normal distribution and leads to an
efficiency equation of the form of Eq. (12). To make the resulting expression more useful for design purposes, the originators of this idea have shown that the efficiency can be approximated empirically by the formula16
= l-

e-(wkA/V)

(15)

where the parameter u>k may be considered as the equivalent


of a performance w, and the parameter k is a measure of the
statistical spread of the variable quantities.
Analysis of field performance data for a significant number
of fly ash precipitators shows that k = 0.5 can be taken as
representative. Values of Wk, calculated from Eq. (15) with
77 and V determined from field tests, and A known from the
precipitator design, may then be used for sizing new precipitators. Observe that this approach is a variant of using Eq. (1),
but with the quantity Wk replacing w.
Another approach is to characterize the precipitator analysis and design process as statistical in the manner commonly
used in the social sciences. In essence the underlying physical
principles and theories are abandoned, and instead field
Journal of the Air Pollution Control Association

performance data are treated as statistical quantities. Performance data for a group of precipitators are subjected to
regression analysis for the purpose of isolating and evaluating
the significant parameters such as gas velocity, gas temperature, and sulfur content of the coal burned. An example of one
such statistical technique is given later in this paper.17 A
similar statistical approach has been applied recently to the
design problem for so-called hot precipitators, that is, those
operating at high temperatures ahead of the air preheaters.18
The utility of regression analysis methods rests mainly on the
improvement gained over simple averaging techniques. The
disadvantages are the absence of a fundamental theoretical
foundation, and the possibility of perpetuating design weaknesses and other deficiencies through gradual deterioration
of the statistical design base.

Downloaded by [Indian School of Mines] at 18:15 08 January 2014

9.
10.
11.
12.

13.

References
1. G. A. Gaffert, Steam Power Stations. 2nd ed. McGraw-Hill, New
York, 1940. p. 305.
2. H. M. Pier and A. N. Crowder, "Catching pulverized-coal ash at
the Trenton Channel plant," Power 65: 834 (1927).
3. D. Carlton-Jones, "Electrostatic precipitation offlyash reaches
fiftieth anniversary," J. Air Poll. Control Assoc. 24: 1035
(1974).
4. "The Air Quality Act of 1967" and The Clean Air Act of 1970."
5. U. S. Environmental Protection Agency, "Standards of Performance for New Stationary Sources," Federal Register 36: (159)
Part II, (Dec. 23,1971).
6. A more detailed discussion of the capture process and the derivation of the probability efficiency equation is given in the book:

January 1977

7.
8.

Volume 27, No. 1

14.
15.
16.
17.
18.

H. J. White, "Industrial Electrostatic Precipitation," Addison-Wesley, Reading, MA, 1963. Ch. 6.


W. Deutsch, Ann. derPhysik 68:335 (1922).
E. Anderson and G. H. Home, "Report on Electrical Precipitation
Experiments at the Experimental Installation at Santa Cruz."
Unpublished report from Western Precipitation Company,
1919.
M. M. Pauthenier and M. Moreau-Hanot, "La charge des particles
spheriques dans un champ ionise," J. de Physique et le Radium
3: 590 (1932).
H. J. White, "Particle charging in electrostatic precipitation,"
AIEE Trans. 70:1189 (1951).
W. B. Smith and J. R. McDonald, "Calculation of the charging
rate of fine particles by unipolar ions," J. Air Poll. Control Assoc.
25:168 (1975).
G. B. Hewitt, "The charging of small particles for electrostatic
precipitation," AIEE Trans. 76:300 (1957). This paper although
almost two decades old remains the most reliable and definitive
experimental study of the charging of fine particles in the micron
and submicron range.
H. J. White, Industrial Electrostatic Precipitation, AddisonWesley, Reading, MA, 1963, pp. 170-172.
G. Herdan, Small Particle Statistics. 2nd ed. Elsevier, New York,
1959.
C. Allander and S. Matts, "The effect of particle size distribution
on efficiency in electrical precipitators," Staub 52:738 (1957).
S. Matts and P. Ohnfeld, "Efficient Gas Cleaning with SF Electrostatic Precipitators," bulletin of A. B. Svenska Flaktfabriken,
Stockholm, Sweden.
Barrett, "Plate Type Precipitators, An Analysis of Performance
on Pulverized Fuel Ash, Part II." Unpublished memorandum of
British Experience, Aug. 1967.
A. B. Walker, "Experience with hot electrostatic precipitators
for fly ash collection in electric utilities," Combustion, 14: (Nov.
1974).

21

Das könnte Ihnen auch gefallen