You are on page 1of 14

Journal of the Air & Waste Management Association

ISSN: 1096-2247 (Print) 2162-2906 (Online) Journal homepage: http://www.tandfonline.com/loi/uawm20

Emissions of Sulfur Trioxide from Coal-Fired Power


Plants

R.K. Srivastava , C.A. Miller , C. Erickson & R. Jambhekar

To cite this article: R.K. Srivastava , C.A. Miller , C. Erickson & R. Jambhekar (2004) Emissions of
Sulfur Trioxide from Coal-Fired Power Plants, Journal of the Air & Waste Management Association,
54:6, 750-762, DOI: 10.1080/10473289.2004.10470943

To link to this article: https://doi.org/10.1080/10473289.2004.10470943

Published online: 22 Feb 2012.

Submit your article to this journal

Article views: 4817

View related articles

Citing articles: 96 View citing articles

Full Terms & Conditions of access and use can be found at


http://www.tandfonline.com/action/journalInformation?journalCode=uawm20
TECHNICAL PAPER ISSN 1047-3289 J. Air & Waste Manage. Assoc. 54:750 –762
Copyright 2004 Air & Waste Management Association

Emissions of Sulfur Trioxide from Coal-Fired Power Plants


R.K. Srivastava and C.A. Miller
Office of Research and Development, National Risk Management Research Laboratory, Air
Pollution Prevention and Control Division, U.S. Environmental Protection Agency, Research
Triangle Park, North Carolina

C. Erickson and R. Jambhekar


Babcock Borsig Power, Inc., Worcester, Massachusetts

ABSTRACT boilers and discusses approaches for mitigating emission


Emissions of sulfur trioxide (SO3) are a key component of of this mist.
plume opacity and acid deposition. Consequently, these
emissions need to be low enough to not cause opacity INTRODUCTION
violations and acid deposition. Generally, a small fraction As understanding of the adverse effects of air pollution
of sulfur (S) in coal is converted to SO3 in coal-fired has grown, so also has the complexity of coal-fired power
combustion devices such as electric utility boilers. The plant design and operation, especially with regard to air
emissions of SO3 from such a boiler depend on coal S pollution control systems. Control of air pollutant emis-
content, combustion conditions, flue gas characteristics, sions is not only a legal requirement but also is becoming
and air pollution devices being used. It is well known that a financial necessity, as salability of effluents and trading
the catalyst used in the selective catalytic reduction (SCR) of emissions increase the direct monetary value of emis-
technology for nitrogen oxides control oxidizes a small sions control. The days when one must only consider the
fraction of sulfur dioxide in the flue gas to SO3. The extent nuisance value of fly ash are long past.1
of this oxidation depends on the catalyst formulation and As plant complexity has increased, so have unex-
SCR operating conditions. Gas-phase SO3 and sulfuric pected consequences of changing segments of the total
acid, on being quenched in plant equipment (e.g., air chemical process that occurs between fuel preparation
preheater and wet scrubber), result in fine acidic mist, and ultimate emissions. One of the more discernible ad-
which can cause increased plume opacity and undesirable verse consequences is the formation and emission of sul-
emissions. Recently, such effects have been observed at fur trioxide (SO3)/sulfuric acid (H2SO4), as highlighted by
plants firing high-S coal and equipped with SCR systems the recent and well-publicized experiences of a power
and wet scrubbers. This paper investigates the factors that plant in Ohio.2 Although not directly subject to emission
affect acidic mist production in coal-fired electric utility limits, SO3 is important to consider during the design and
operation of coal-fired utility boilers for a number of
environmental and plant performance reasons.
IMPLICATIONS The formation of SO3 will occur during the combus-
Formation and emissions of SO3/H2SO4 can potentially tion of sulfur (S)-bearing fuels such as coal and heavy fuel
lead to plant operation-related problems and environmental oils. Virtually all of the SO3 converts to H2SO4 as flue gas
concerns. The emissions of SO3 from a boiler depend on a
is cooled in the air preheater (APH). Relatively high con-
number of interacting complex factors including coal sulfur
content, combustion conditions, flue gas characteristics, centrations of SO3/H2SO4 in the boiler, stack, or plume
and air pollution devices being used. Understanding the can cause adverse impacts to plant equipment and to the
paramenters leading to excessive generation of SO3/H2SO4 environment. Impacts on plant equipment can include
emissions can assist in selection of practical control ap- corrosion, fouling, and plugging and may require addi-
proaches. This paper investigates the factors that affect
tional hardware or changes in operation to minimize SO3/
SO3/H2SO4 emissions from coal-fired electric utility boilers
H2SO4 concentrations and the resulting adverse impacts.
and discusses approaches for mitigating these emissions.
The paper is expected to be a state-of-the-art reference on
the subject for regulators, utility industry personnel, and Health and Environmental Effects
others stakeholders. Formation of visible H2SO4 droplets depends upon
the concentration and dew point of H2SO4 and the

750 Journal of the Air & Waste Management Association Volume 54 June 2004
Srivastava, Miller, Erickson, and Jambhekar

concentration of sub-micron particles upon which the


acid can condense. As seen in Figure 1, H2SO4 dew point
is a function of water vapor and H2SO4 concentration and
increases as both these variables increase.3 The curves in
Figure 1 determine the fraction of H2SO4 in the vapor and
condensed phases for temperatures below the dew point.
For example, a 250 °F flue gas with 10% water vapor and
8 ppm H2SO4 will have a vapor-phase acid gas concentra-
tion of ⬃2 ppm, with the remaining 6 ppm in the con-
densed phase.
If temperatures at the stack are low enough and water
vapor and H2SO4 concentrations are high enough, the
condensed-phase concentration of H2SO4 can be at a level
that results in the formation of a visible plume attached to
the stack. Once the gases leave the stack, the rate of
cooling and subsequent H2SO4 condensation competes Figure 2. Iso-opacity plot calculated for a bimodal distribution of
coarse and sub-micron particles. The sub-micron particle mode is at
with the plume dilution by entrainment of ambient air
0.15 ␮m diameter.5 The gray area represents typical sub-micron mass
into the plume. Even in cases where stack conditions are concentrations in the stack of a coal-fired utility boiler.6
such that the H2SO4 is completely in the vapor phase at
the stack exit, a detached plume may still form shortly of sub-micron particle concentrations downstream of an
downwind as temperatures drop below the dew point. electrostatic precipitator (ESP).6
In most cases, and particularly in the case where The curves in Figure 2 illustrate that H2SO4 concen-
H2SO4, water, and sub-micron solid particles are present, tration can have a strong impact on opacity when fine
condensation is the dominant formation mechanism. particle concentrations are present at realistic levels. Be-
The opacity of these plumes is most strongly influenced cause the condensing acid particles can nucleate to form
by the concentrations of condensing species and sub- particles even at very low levels of pre-existing sub-
micron fly ash particles present in the stack gases. The size micron particles, further reduction of sub-micron solid
distribution of the sub-micron fly ash also can have a particles may not significantly reduce plume opacity. As
noticeable effect on plume opacity at low to moderate an example, a unit emitting 16 ppm of vapor-phase
H2SO4 concentrations, because these particles act as con- H2SO4 and 5 mg/m3 of sub-micron particles is predicted
densation sites for the condensing vapor-phase H2SO4. to result in a plume opacity of ⬃35%. An opacity of 20%
The relationship between sub-micron particle concentra- can be achieved by reducing sub-micron PM concentra-
tion, H2SO4 concentration, and plume opacity is shown tion to ⬃1.6 mg/m3 (i.e., ⬃70% reduction) or by reducing
in Figure 2 for flue gas with a sub-micron particle mode at H2SO4 to ⬃8 ppm (a 50% reduction). For units with
0.15 ␮m diameter.4,5 The gray area denotes a typical range limited ability to further reduce PM emissions, control of
SO3 (and, subsequently, H2SO4) may be the only viable
option for meeting plume-opacity requirements. Because
essentially all the SO3 is converted to H2SO4 at or below
stack temperatures, Figure 2 also illustrates the impact of
increases in SO3 formation on plume opacity. A unit with
1 mg/m3 of sub-micron particles in the stack gases is
predicted to experience doubling of plume opacity from
⬃5 to 10% when H2SO4 concentrations increase from 5 to
10 ppm.
Data on direct adverse effects on human health are
inconsistent, but studies indicate that aqueous acidic
aerosols at typical and even elevated ambient concentra-
tions generally have minimal effects on symptoms and
mechanical lung function in young healthy adults. How-
ever, there are studies that have shown changes in muco-
ciliary clearance and modest bronchoconstriction in asth-
Figure 1. Dew point of H2SO4 as a function of H2SO4 concentration matics exposed to elevated concentrations of 400 ␮g/m3
for different water vapor concentrations.3 or more.7,8 Of more concern than the effects of exposure

Volume 54 June 2004 Journal of the Air & Waste Management Association 751
Srivastava, Miller, Erickson, and Jambhekar

to typical ambient levels of SO3/H2SO4 aerosol alone is limit the combined sulfur dioxide (SO2) and SO3 emission
exposure to this aerosol mixed with other ambient PM concentrations to a daily average of 50 mg/Nm3 and a
constituents, including elemental carbon and metals.9 30-min average of 200 mg/Nm3. There is no separate
In situations where SO3/H2SO4 aerosol combines emission limit for SO3 alone.14
with a sinking plume, the ambient concentrations near In the United States, SO3 emissions are included as
the stack can reach significantly higher levels than those part of limits on opacity and particulate matter (PM)
normally experienced. In such cases where meteorologi- emissions, and H2SO4 is listed as a hazardous air pollutant
cal and operating conditions combine to form a near- (HAP) under Title III of the Clean Air Act Amendments of
stack acidic fog, damage to property and vegetation can 1990.15 Under Title III, electric utility generating stations
occur if the conditions are sustained for an extended are exempt from the mass emission limits of 10 t (9.09
period. Anecdotal evidence of adverse health impacts, tonnes) per year of any single HAP or 25 t (22.7 tonnes)
such as burning eyes, sore throats, and headaches, has per year of combined HAPs set for other industrial
also been reported in such cases.2 sources. Thus, although H2SO4 is a listed HAP and is
emitted in significant quantities by coal-fired utility gen-
Impacts on Plant Hardware erating stations,16 there is no regulatory requirement to
The most common problem associated with elevated con- control these emissions as HAPs. They must, however, be
centrations of SO3 is low-temperature (⬍300 °F [150 °C]) reported to the U.S. Environmental Protection Agency
corrosion.10 Once formed, SO3 reacts easily with the mois- (EPA) under the Toxics Release Inventory (TRI) require-
ture in combustion flue gases to produce H2SO4. Below its ments. The H2SO4 is typically measured and reported as
dew point, H2SO4 condenses and will collect in relatively equivalent SO3.
low-temperature areas of the flue gas path, such as the air EPA has defined primary PM as particles that are
heater, and corrode the contacted metal components. The either emitted directly as a solid or liquid or are emitted as
dew point varies with H2SO4 and water concentrations in a vapor but condense or react upon cooling and dilution
the flue gas (see Figure 1), but is typically between 200 and in the ambient air to form solid or liquid PM immediately
300 °F (95 and 150 °C, respectively), corresponding to low after discharge from the stack.17 Secondary PM is com-
(8%) moisture, low (0.1 ppm) H2SO4 conditions and high posed of particles that form through chemical reactions in
(15%) moisture, high (20 ppm) H2SO4 conditions, respec- the ambient air well after dilution and condensation have
tively. At higher temperatures (above 1000 °F [540 °C]), occurred.18 Under this definition, SO3 or H2SO4 that is in
SO3 can cause corrosion in superheaters and reheaters.11 the vapor phase in the furnace or stack and either reacts to
These high-temperature reactions are less common than form SO42⫺ particles or condenses to form liquid H2SO4
the low-temperature condensation and corrosion reac- immediately after leaving the stack is considered part of
tions. In addition to forming corrosive H2SO4, SO3 reacts primary PM emissions and is therefore subject to PM
to form sulfate (SO42⫺) particles, especially ammonium emission limits.
sulfate [(NH4)2SO4] and ammonium bisulfate (NH4HSO4), In some states or localities, plants may be required to
in boilers using ammonia (NH3), which can lead to foul- control SO3 to maintain opacity standards. Even though
ing or plugging of APH passages.12 Regular maintenance is opacity measurements in the stack may be within regula-
conducted at plants to address the impacts described. This tory limits, the formation of a visible plume caused by the
maintenance can include monitoring and replacement of presence of SO3 may result in plume opacities higher than
materials, washing of APH passages, and other activities. allowable levels and a need to control SO3 emissions.
In some cases, modification or addition of hardware Most states have a 20% opacity limit for stationary
may be required to minimize corrosion or plugging. At sources, including coal-fired utility boilers, and many also
plants using wet flue gas desulfurization (FGD) systems, have provisions for higher opacity levels during unit
the reduction in flue gas temperature can result in in- startup. There are some exceptions to the 20% limit, with
creased H2SO4 condensation and subsequent corrosion. a few states having opacity limits of 40% and at least one
In such cases, the flue gas may require reheating to ensure (West Virginia) with a 10% opacity limit.
it remains above the dew point, thereby resulting in in- The formation of a detached plume is indicated when
creased operation costs.13 in-stack opacity measurement values are lower than those
in the downwind plume. While in-stack opacity measure-
REGULATORY OVERVIEW ments are made using continuous emission monitors,
There are currently no U.S. regulations that directly limit measurements of the downwind plume opacity are cov-
emissions of SO3/H2SO4 aerosols from utility boilers. ered under EPA Method 9.9 The increase in opacity from
Other regulatory programs may, however, require SO3 in-stack measurements to downwind is often caused by
control, as discussed later. German emission standards the condensation of vapors such as H2SO4 and can result

752 Journal of the Air & Waste Management Association Volume 54 June 2004
Srivastava, Miller, Erickson, and Jambhekar

in regulatory violations and good neighbor policy con- Table 1. The reactions involving SO3 in the combustion zone. The rate constants
cerns. are listed in the form k ⫽ AT ␤ exp(⫺EA /RT ) with units in mol-cm-sec.

Reaction A ␤ EA /R
FORMATION OF SO3
Formation in the Boiler
SO3 ⫹ H ⫽ HOSO ⫹ O 2.5e05 2.92 25,300
The S in coal has inorganic and organic components. The
SO3 ⫹ O ⫽ SO2 ⫹ O2 2.0e12 0 10,000
inorganic component is predominantly pyrite, which can SO3 ⫹ SO ⫽ SO2 ⫹ SO2 1.0e12 0 5000
exist both as distinct particles (excluded) and bound in SO2 ⫹ O ⫹ M ⫽ SO3 ⫹ M 9.2e10 0 1200
the coal matrix (included). The organic component is part SO2 ⫹ OH ⫽ SO3 ⫹ H 4.9e02 2.69 12,000
of the various organic structures present in coal. The HOSO2 ⫽ SO3 ⫹ H 1.4e18 ⫺2.91 27,600
fractions of included and excluded pyrite and organic HOSO2 ⫹ O ⫽ SO3 ⫹ OH 5.0e12 0 0
component delivered to a furnace are known to be depen- HOSO2 ⫹ OH ⫽ SO3 ⫹ H2O 1.0e12 0 0
dent on the fineness to which the coal is ground. During HOSO2 ⫹ O2 ⫽ SO3 ⫹ HO2 7.8e11 0 330
the combustion of coal, virtually all the S gets oxidized to
gaseous species such as SO2 and SO3, with SO2 being the
principal oxide. Although detailed chemical mechanisms 0.87% nitrogen (N), 0.78% S, 9.34% ash, 21.23% water,
for oxidation of fuel S are not understood well at present, and 11.16% O. The bituminous coal contains 69.79% C,
it is believed that this oxidation proceeds through rapid 4.79% H, 1.34% N, 2.95% S, 7.47% ash, 5% water, and
formation of SO2, occurring on a timescale comparable to 8.65% O. The initial compositions resulting from com-
that of fuel oxidation reactions. Because SO2 formation is plete combustion of the coals are then equilibrated at
so rapid, its concentration can be estimated using equi- 1647 °C (1900 K) to provide the input composition for the
librium calculations. chemical kinetic calculations based on the S chemistry.
The primary reaction that results in SO3 formation in Note that the H2-O2-carbon monoxide (CO) subset of the
flames is hydrocarbon mechanism given in ref 21 is used in these
calculations. This subset includes 25 reactions involving
11 species.
SO 2 ⫹ O ⫹ M 3 SO 3 ⫹ M (1)
Figures 4 and 5 plot SO3 mole fraction and the ratio
SO3 mole fraction/SO2 mole fraction versus time, respec-
This recombination reaction proceeds rapidly near the
tively. Because reaction 1 is exothermic, little SO3 is
combustion zone in the presence of super-equilibrium
formed initially at high temperatures in the furnace. As
concentrations of oxygen (O) atoms. The consumption of
the flue gas cools, SO3 is produced, and the rate of pro-
SO3 near flames occurs primarily via
duction is high in the convective region of the boiler (see
Figures 3 and 4). SO3 production via gas-phase S kinetics
SO 3 ⫹ HO 2 3 HOSO 2 ⫹ O 2 (2) is complete before the flue gas enters the economizer. The
results also reflect that SO3 production increases with coal
and

SO 3 ⫹ O 3 SO 2 ⫹ O 2 (3)

A few other reactions are considered to result in formation


and destruction of SO3 near the flame zone, but the effect
of these on SO3 concentrations is relatively minor. Table
1 summarizes the reactions involving SO3 in the combus-
tion zone.19
To understand SO3 formation, calculations were
made using the S chemistry described previously with gas
compositions and temperature-time history characteristic
of a coal-fired boiler.20 This history is shown in Figure 3.
Initial gas-phase compositions are assumed to result from
complete combustion (with 10 and 30% excess air) of a
Montana sub-bituminous and a Western Kentucky bitu-
minous coal. The sub-bituminous coal is composed of
53.26% (by weight) carbon (C), 3.35% hydrogen (H), Figure 3. Temperature-time history for a coal-fired power plant.20

Volume 54 June 2004 Journal of the Air & Waste Management Association 753
Srivastava, Miller, Erickson, and Jambhekar

Figure 4. SO3 produced during coal combustion.

S content and furnace excess air level. This is because extent of SO3 formation caused by catalytic oxidation of
increases in the S content of coal or furnace excess air SO2. A laboratory study found that conversion of SO2 to
level result in a corresponding increase in the concentra- SO3 in the presence of fly ash ranged between ⬃10 and
tion of SO2 and/or O, which, in turn, results in greater 27% for temperatures between 500 and 700 °C, and es-
production of SO3 via reaction 1. Finally, Figure 5 reflects sentially increased linearly with the iron oxide content of
that SO2 conversion rate (i.e., SO3/SO2 molar ratio) for the the ash.23 These conversion rates are substantially higher
modeled coals ranges between 0.1 and 0.65%, approxi- than those found in coal-fired boilers, for which data
mately, and is relatively independent of coal type at a suggest furnace/economizer conversion to be ⬃0.8 to
specific excess air level. Because boilers are operated with 1.6% for bituminous and 0.05– 0.1% for sub-bituminous
between 10 and 30% excess air, a 0.65% conversion of coals.24 The above laboratory study results indicate that
coal S to SO3 should provide a conservative estimate of temperatures and residence times greater than, and car-
SO3 production in the furnace. bon ash contents lower than, those typically found in
In addition to the SO3 formation in the furnace dis- coal-fired utility boilers are required to achieve high oxi-
cussed above, additional formation takes place in the dation rates.
temperature range of 1100 – 800 °F (593– 427 °C) found in
the economizer region of the boiler. This formation re-
sults from oxidation of SO2 via molecular oxygen (O2) Formation of SO3 in SCR Reactors
catalyzed by iron oxides present in both ash and tube The SCR technology is increasingly being used at power
surfaces.22 This oxidation mechanism depends on several plants to control emissions of oxides of nitrogen (NOx). In
operating and design parameters, including SO2 concen- the SCR process, NH3 is injected into the flue gas within a
tration, ash content and composition, convective pass temperature range of ⬃600 –750 °F (315– 400 °C), up-
surface area, gas and tube surface temperature distribu- stream of a catalyst. Subsequently, as the flue gas contacts
tions, and excess air level. Because the impact of these the SCR catalyst, NOx, which predominantly is nitric ox-
parameters depends on site-specific factors (e.g., cleanli- ide (NO) in combustion devices, is chemically reduced to
ness of tube surfaces), it is difficult to characterize the molecular nitrogen (N2). In the most commonly used SCR

754 Journal of the Air & Waste Management Association Volume 54 June 2004
Srivastava, Miller, Erickson, and Jambhekar

Figure 5. SO3 conversion during coal combustion.

process layout, known as hot-side SCR, the catalyst is example illustrates that the level of oxidation across the
located between the economizer and the APH. catalyst can have a significant impact on SO3 concentra-
It is well known that the catalyst used in the SCR tion.
technology oxidizes a small fraction of SO2 in the flue gas In general, for a given catalyst material experiencing
to SO3:25–29 the same flue gas conditions, the oxidation rate of SO2 to
SO3 (or conversion rate) is inversely proportional to area
SO 2 ⫹ 1 ⁄ 2 O 2 3 SO 3 (4) velocity (AV), which is simply the ratio of flue gas volu-
metric flow rate to geometric catalyst surface area. This
implies that the conversion rate is proportional to catalyst
The extent of this oxidation depends on the catalyst for-
volume (and, hence, geometric surface area) and gas res-
mulation and SCR operating conditions. Generally, this
idence time in the catalyst. Thus, the conversion rate, ␬,
oxidation can range from 0.25 to 0.5% of SO2 in bitumi-
can be expressed as
nous and from 0.75 to 1.25% in low-S sub-bituminous
coal applications. To examine what this oxidation means,
consider an SCR application with an SO2 oxidation guar- ␬ ⫽ K/AV (5)
antee of 0.5%. Also assume that the concentrations of SO2
and SO3 in the flue gas at the inlet of the SCR reactor are where the constant of proportionality, K, is a function of
2000 and 20 ppm, respectively. Then, based on oxidation catalyst material and design as well as flue gas proper-
across the catalyst, the concentrations of SO2 and SO3 at ties;30 that is,
the exit of the SCR reactor will be 0.995 ⫻ 2000 ⫽ 1990
ppm and 2000 ⫺ 1990 ⫹ 20 ⫽ 30 ppm, respectively. Thus,
K ⫽ f(catalyst material, catalyst design,
the SO3 loading in the flue gas at the exit of the SCR
reactor will be 50% more than that at the inlet. This flue gas properties) (6)

Volume 54 June 2004 Journal of the Air & Waste Management Association 755
Srivastava, Miller, Erickson, and Jambhekar

A systematic study of oxidation of SO2 to SO3 over hon- APH design does not have surfaces that are periodically
eycomb SCR catalysts has been reported in ref 31. The exposed to combustion air; except for small leakages in
findings of this study are as follows: welds and seal, no mixing of the flue gas and combustion
(1) The conversion rate depends primarily on the air occurs. Therefore, the resultant H2SO4 formed is not
vanadium content of the catalyst and, therefore, removed by evaporation into the combustion air and
can be controlled by adjusting this content; passes directly out of the APH. Modeling of the acid con-
(2) The oxidation reaction is considerably slower densation phenomenon in tubular APHs is described in
than diffusion of SO2 within the pores of the ref 35.
catalyst. Therefore, the entire volume of the cat- In addition to the condensation-evaporation of SO3
alyst is active in oxidation of SO2 to SO3 in con- discussed above, an additional conversion process takes
trast to reduction of NOx to N2, which, being place in boilers using SCR. In such applications, NH3 is
diffusion-limited, occurs mainly at the catalyst injected as a reagent in the SCR process. A minor fraction,
surface. The rate of oxidation is linearly propor- 2–5 ppmv, of injected NH3 slips past the SCR catalyst and
tional to catalyst wall thickness. Accordingly, re- does not react with NOx. This fraction of NH3, known as
ducing the wall thickness should not affect NOx NH3 slip, reacts with SO3 downstream of the SCR reactor
reduction but should reduce SO2 to SO3 conver- and forms (NH4)2SO4 and NH4HSO4 salts and, thereby,
sion; results in removal of SO3 from the flue gas.12 This salt
(3) The reaction rate is of variable order in SO2 con- formation can be detrimental to the APH performance if
centration, increases with temperature, is inde- APH passages become plugged and pressure loss across the
pendent of concentrations of O2 and water in APH results in forcing offline washing. The amount and
practical applications, is strongly inhibited by type of salt formed will depend on the amount of NH3
NH3 and is slightly enhanced by NOx. slip; based on typical concentrations of SO3 and H2O in
flue gas, the amount of NH3 slip is the limiting factor in
SO3 DEPLETION OR CONVERSION TO H2SO4 salt-formation reactions.
Processes in APHs
Utility boilers use APHs to transfer heat from hot flue gas Localized Condensation of H2SO4 in the Duct
exiting the economizer to combustion air flowing into the between APH and PM Control
boiler. These APHs are available in rotary regenerative and Condensation of H2SO4 also can take place in the duct
tubular designs, with the former used more widely. Typ- between the APH and the PM control device.35,36 For
ically, the flue gas temperature at the APH inlet is between example, because of the rotating heat transfer element
600 and 700 °F (316 and 371 °C) and ⬃300 °F (149 °C) at employed in the regenerative APH, gas flow stratification
the exit. SO3 is hygroscopic and, therefore, absorbs vapor- across the flow cross-section at the APH inlet is enhanced
phase moisture at temperatures above its dew point to in the APH. As a result, strong transverse variations in gas
form H2SO4 vapor.22 This process occurs in the APH. The temperature and H2SO4 vapor concentration can exist in
extent of conversion of SO3 to H2SO4 depends on the the gas leaving the APH. These variations can lead to
temperature distribution in the APH and flue gas moisture localized condensation of H2SO4. Based on operating con-
content. However, virtually all SO3 converts to H2SO4 at ditions, localized condensation also can occur in units
temperatures of 400 °F or less. If local metal temperatures using tubular APHs. For a given combination of flue gas
in the APH flow passages drop below the acid dew point, H2SO4 concentration and moisture content, Figure 1 can
some H2SO4 condenses on these surfaces as liquid drop- be used to determine if condensation is occurring.
lets (aerosol). This rate of condensation is dependent on
the wall temperature and H2SO4 concentration in the flue Fly Ash Adsorption and Removal of H2SO4 in PM
gas. Control Equipment
In a regenerative APH, where flow passages are peri- In addition to the processes described, some H2SO4 gets
odically exposed to hot flue gas and relatively cold incom- adsorbed on fly ash in the APH and downstream equip-
ing combustion air, evaporation of condensed H2SO4 oc- ment. The rate of this adsorption depends on the temper-
curs on exposure to air. The rate of this evaporation is ature of the flue gas, concentration of H2SO4, and fly ash
dependent on the moisture content of air and metal sur- properties, in particular alkalinity.23,37 The adsorption
face temperature of the APH flow passage.32,33 may increase rapidly as the flue gas reaches the cold end
Results of an extensive field test program indicate of the APH and may continue in the duct between the
that ⬃40% of the flue gas SO3 present at the regenerative APH and the PM control equipment. The adsorbed H2SO4
APH inlet is removed in the APH by the condensation- gets removed with the fly ash in the PM control device
evaporation mechanism discussed above.34 The tubular (ESP or baghouse).

756 Journal of the Air & Waste Management Association Volume 54 June 2004
Srivastava, Miller, Erickson, and Jambhekar

The firing of sub-bituminous coals, which generally For sub-micron mist droplets with diameters less than
have S contents on the order of 0.5%, results in fly ash ⬃0.05 ␮m, Brownian diffusion is known to be the primary
with a relatively high amount of alkali (20 –30% by component of mass transfer.40 In practical situations, de-
weight). Such alkaline ash adsorbs virtually all H2SO4 in pending on the relative velocity between the sub-micron
the flue gas.38 In such cases, SO3 injection is required for mist droplets and the collecting liquid form (droplets,
improving ESP performance. In contrast, based on fly ash wetted walls, liquid sheets, and bubbles), the Brownian
properties and temperature, the majority of SO3 in flue motion-induced mass transfer is often not adequate to
gas of a bituminous coal-fired boiler may or may not be result in high-efficiency capture of sub-micron droplets.
adsorbed in fly ash.30 Finally, hot-side ESPs operate at Therefore, generally, larger droplets in the mist can be
high-enough temperatures where little adsorption of removed in the scrubber, but a significant portion of the
H2SO4 occurs. sub-micron droplets are not removed and are emitted
As mentioned above, adsorption of H2SO4 on fly ash from the stack. This explains why wet FGD systems are
depends on the alkalinity of the fly ash and the concen- relatively inefficient in removing SO3/H2SO4. In general,
tration of H2SO4. Figure 6 illustrates the general adsorp- ⬃50% of the H2SO4 entering the scrubber may be re-
tion characteristics of fly ash as a function of the molar moved in the scrubber.
ratio of alkali content of fly ash and SO3 concentration at
the inlet of the APH. The alkali content of fly ash is MEASUREMENT OF SO3
defined as the molar sum of magnesium oxide (MgO) and The controlled condensation system (CCS) has been used
calcium oxide (CaO) in the ash. This figure represents a since the 1960s for the measurement of SO3 and H2SO4 in
correlation of data from Babcock Borsig Power field-test- flue gas streams.41,42 The general arrangements of the CCS
ing programs concerning SO3 production and capture in and thimble holder are shown in Figures 7 and 8, respec-
flue gas systems. tively. The measurement system consists of a quartz-lined
heated probe that draws gas through a quartz thimble for
Aerosol Formation in Wet FGD Systems the removal of PM. The probe and thimble are heated to
As the H2SO4 vapor-containing flue gas passes through a avoid condensation of SO3 vapor in the gas sample. The
wet FGD system, it is rapidly cooled below the acid dew gases enter a temperature-controlled condenser where the
point. Because the rate of this cooling is greater than the SO3 is condensed on the wall to be collected and mea-
rate of absorption of H2SO4 vapor in the scrubber solu- sured after the sample run via a deionized water rinse.
tion, the dew-point crossover results in H2SO4 mist with The measurement techniques for SO3 and H2SO4 re-
sub-micron droplets.30,39 In a wet scrubber, generally, the cently have been under review for accuracy and improve-
mass transfer between this mist and the scrubbing liquid ment for TRI reporting. Recently, a review and verification
occurs through inertial impaction, gravitational settling, project has been undertaken by EPRI to qualify the
Brownian diffusion, diffusiophoresis, and thermophoresis. CCS method and compare field data with available

Figure 6. Fly ash SO3 capture rate downstream of furnace.

Volume 54 June 2004 Journal of the Air & Waste Management Association 757
Srivastava, Miller, Erickson, and Jambhekar

Figure 7. CCS sampling train.

prediction methods used to estimate SO3 emissions for MITIGATION OF SO3 EMISSIONS
TRI reporting.37 The EPRI project compared laboratory The mitigation of SO3 has been an active area of research
data with field test data for a variety of coals with a range for many years.13,23,44,45 This research has resulted in the
of S and fly ash chemical compositions. The study further development, refinement, and implementation of differ-
investigated the SO3 removal rates and efficiencies of the ent techniques and methods for the successful mitigation
plant equipment used for field verification. of SO3 in the flue gases of fossil-fuel–fired boilers. Re-
The EPRI project has concluded that alkaline ash will cently, a guide that summarizes SO3 mitigation options
produce a bias in the SO3 measurement. The bias is a and their respective success in either full-scale or pilot
result of SO3 removal in the thimble holder by the alka- testing has been written.10 The mitigation options are
line ash collected on the quartz thimble. This bias is seen described below.
as significant for low-S, high-alkaline ash fuels (Powder
River Basin coals). The CCS was found to have a bias of Alkali Addition into the Furnace
20 –25% low readings for high S, acidic ash fuels and The injection of alkali in the flue gas stream has been a
greater than 40% for Powder River Basin coals. A possible method of SO3 mitigation for almost 30 years.44 The lo-
solution to the bias is utilized in Europe.43 The test cation in the flue gas stream and the delivery method of
method uses a system similar to the CCS; however, the the alkali have been studied and tested depending on
quartz thimble is replaced with a small tubular ESP. The operating and site conditions.
tubular ESP removes the ash to the sidewalls away from The addition of alkali into the furnace recently has
the gas stream; in contrast, the quartz thimble filters the been proven to be effective at the full scale.30,39 However,
flue gas through the collected ash. Furthermore, the Eu- the method of delivery and the effectiveness of the alkali
ropean system can distinguish between gaseous and aero- used varies from site to site depending on specific condi-
sol SO3 using deionized water procedures on various col- tions. The addition of limestone to coal before pulveriza-
lection plates. The European SO3 method system has been tion for the control of SCR catalyst arsenic poisoning has
used extensively in the United States on a variety of coals been shown to be an effective method of furnace SO3
with repeatable and reliable results. control. Recent commissioning data46 have shown at least

758 Journal of the Air & Waste Management Association Volume 54 June 2004
Srivastava, Miller, Erickson, and Jambhekar

Figure 8. CCS thimble holder.

a 50% reduction in furnace SO3 from limestone addition. of the SO3 in the flue gas at various plants depending on
The injection of alkaline sorbents (Ca- and Mg-based slur- the injected material and injection rate.10 The sorbents
ries) has been shown to be effective at controlling SO3 are introduced into the flue gas as either a dry powder or
emissions of the furnace.39 The slurry injection method mixed with water to form slurry before injection. The
has successfully obtained SO3 furnace conversion reduc- location of injection in the flue gas stream varies. Plants
tions of 40 – 80% but has been found to be sensitive to with APH cold-end corrosion problems may elect to inject
injection location and elevation. The effects of the MgO the sorbent before the APH; however, adequate APH
sorbent slurries on SCR catalyst activity are currently un- cleaning equipment is required with this configuration.
der investigation. To date, no studies have measured the In general, the common injection location for SO3 con-
potential benefit of SCR catalyst arsenic poisoning control trol is between the APH and the ESP. The injection of
by MgO sorbent injection. The addition or injection of sorbent before the ESP must, however, consider the effect
alkali in the furnace does not influence the conversion on particulate control. The ESP will have higher inlet
rate of the SCR catalyst. mass loading, and the fly ash will have different resistivity
characteristics. It has been reported that dry injection of
Alkali Injection after the Furnace hydrated lime has resulted in strong sparking and lower
The injection of other alkaline materials after the furnace/ operating currents in the ESP during pilot-scale testing.10
economizer exit has been used for the control of SO3 for
both APH corrosion and stack emissions.10 The primary Ammonia Injection before the ESP
sorbents used are compounds such as hydrated lime, lime- The injection of NH3 after the APH and before the ESP has
stone, MgO, and sodium carbonate. The selection of a been shown to be ⬍90% effective in the removal of SO3 in
sorbent for a given site will depend on economic factors, full-scale application.10 This method of mitigation re-
such as availability and required SO3 removal rates. Alkali sults in the formation of (NH4)2SO4 and NH4HSO4 salts in
injection has successfully removed between 40 and 90% the ESP, depending on the NH3 and SO3 concentration

Volume 54 June 2004 Journal of the Air & Waste Management Association 759
Srivastava, Miller, Erickson, and Jambhekar

ratios. The formation of NH4HSO4 is expected when NH3 maintenance parameters.47 While dry ESPs are typically
to SO3 molar ratios are less than 1; this formation tends limited to power levels of 100 –500 W per 1000 cfm of flue
to decrease the ESP particle loading caused by fly ash gas, WESPs can operate with power levels as high as 2000
agglomeration. NH3 injection before the ESP is used for fly W per 1000 cfm. Because of wet cleansing of the collec-
ash conditioning to increase ESP performance caused by tion system, PM does not accumulate in the collection
the agglomeration effects. With NH3 to SO3 molar ratios electrodes; this mitigates particle re-entrainment. Based
between 1 and 2, increased formation of (NH4)2SO4 is on these factors, WESPs can collect sub-micron particles
expected, with an increase in the particle loading of the and acid mist very efficiently. WESPs can be configured
ESP. The injection of NH3 results in the adsorption of NH3 for vertical or horizontal gas flows in tubular or plate
by the fly ash. Because the fly ash will contain most of the designs. Tubular designs offer smaller footprints and, in
injected NH3, the concentration may exceed acceptance general, are more efficient than the plate type.
limits for ash salability. Additional treatment of fly ash WESPs can be integrated easily with a wet scrubber.
holding ponds and basins may be required if large In fact, integration of the WESP within the wet scrubber is
amounts (⬃30 ppmv) of SO3 are being removed. The use a design option with many attractive features including a
of NH3 for SO3 mitigation is practical on units equipped compact footprint; the ability to integrate the handling of
with SCRs, because NH3 is used as the reagent and is the wash water and solids from the WESP with scrubber
readily available on site. slurry, thereby avoiding the need for separate tank and
blowdown system; and the ability to collect the fine
Fuel Switching and Blending H2SO4 mist, which typically escapes the scrubber because
The firing of sub-bituminous coals typically results in low of its very small droplet size.47
SO3 formation and emission rates. Consequently, switch- In 1986, the first commercial WESP application on a
ing from firing bituminous to sub-bituminous coals can U.S. power plant took place when AES Deepwater, a
be a mitigation option. However, equipment and fuel cost 155-MW cogeneration plant firing petroleum coke as the
factors often make such a change impossible. Many boiler primary fuel, was equipped with a WESP. With the WESP
systems do not have the capacity and the equipment to in operation, the plume opacity at the plant is generally
accommodate the firing of sub-bituminous coal without 10% or less.48
major modifications that make fuel switch economically Recently, an upflow tubular design WESP has been
unacceptable. Coal availability and costs also can con- retrofit at Northern States Power Company’s Sherco Sta-
strain fuel switching. One possible solution is the blend- tion in a wet scrubber/WESP configuration. Two more
ing of bituminous and sub-bituminous coals to create a power plant applications are underway presently: (1) a
blend that has fuel and ash characteristics favorable for 5000-cfm slipstream at Bruce Mansfield Station; and (2) a
SO3 emissions. This strategy is currently used for control- plate-type WESP for integration with Powerspan’s ECO
ling SO2 emissions. When sub-bituminous coals are technology to be demonstrated at First Energy’s R.E.
blended with bituminous coals, the overall S content of Burger plant.
the fuel is reduced, resulting in a reduction of the SO2 Tests at the Sherco Station (WESP retrofit to the outlet
concentration in the flue gas as well as the conversion of section of the wet scrubber) allowed the scrubber to main-
SO2 to SO3 in the furnace and SCR and, thereby, resulting tain a 70% SO2 reduction while keeping particulate emis-
in an overall reduction in SO3. Also, the sub-bituminous sions at 0.01 lb/106 Btu and opacity under 10%. Full
coal ash contains large percentages of alkaline materials conversion of all scrubber modules at the plant with
that further assist in the capture of SO3 in the APH and WESPs is now underway.49 The WESP at the Mansfield
ESP. Station is achieving greater than 95% removal of SO3 and
PM2.5 and stack flow with near 0 opacity.50
Wet ESPs
Similar to dry ESPs, wet ESPs (WESPs) operate in a three- Changing the Operation of APH
step process involving (1) charging of the entering parti- Increasing the heat transfer (or cooling of the flue gas) in
cles; (2) collection of the particles on the surface of an the APH would appear to be a potentially viable strategy
oppositely charged surface; and (3) cleaning the collec- for removing some of the SO3/H2SO4 in the APH. This, in
tion surface. Both technologies employ separate charging turn, would lead to increased condensation of H2SO4 in
and collection systems. However, the collecting surface in the APH and also to improved plant efficiency. However,
WESPs is cleaned with a liquid, in contrast to mechanical the dew point at the APH outlet is ⬃230 °F (110 °C),
cleaning in dry ESPs. Consequently, the two technologies thereby limiting the SO3 removal to ⬃90%.30 On the
differ in the nature of particles that can be removed, other hand, the potential for corrosion in the APH and
the overall efficiency of removal, and the design and downstream duct would increase with increased H2SO4

760 Journal of the Air & Waste Management Association Volume 54 June 2004
Srivastava, Miller, Erickson, and Jambhekar

condensation. Consequently, more frequent soot blowing 4. Damle, A.S.; Ensor, D.S.; Sparks, L.E. Prediction of the Opacity of
Detached Plumes Formed by Condensation of Vapors; Atmos. Environ.
may be required to control this corrosion. These factors 1984, 18, 435-444.
would need to be considered when deciding to change the 5. Damle, A.S.; Ensor, D.S.; Sparks, L.E. Options for Controlling Conden-
sation Aerosols to Meet Opacity Standards; J. Air Pollut. Control Assoc.
APH operation to mitigate SO3 emissions. Data reflect that 1987, 37, 925-933.
6. Markowski, G.R.; Ensor, D.S.; Hooper, R.G.; Carr, R.C. A Submicron
⬃25% increase in H2SO4 condensation may be possible by Aerosol Mode in Flue Gas from a Pulverized Coal Utility Boiler; Envi-
lowering the flue gas temperature at the APH exit by ⬃40 ron. Sci. Technol. 1980, 14 (11), 1400-1402.
7. Leikauf, G.D.; Spektor, D.M.; Albert, R.E.; Lippmann, M. Dose-
°F (22 °C). Dependent Effects of Submicrometer Sulfuric Acid Aerosol on Particle
Clearance from Ciliated Human Lung Airways; Am. Ind. Hyg. Assoc. J.
1984, 45, 285-292.
SUMMARY 8. Utell, M.J.; Frampton, M.W.; Morrow, P.E. Air Pollution and Asthma:
Clinical Studies with Sulfuric Acid Aerosols; Allergy Proc. 1991, 12,
Formation and emissions of SO3/H2SO4 potentially can 385-388.
lead to plant operation-related problems and environ- 9. Method 9 —Visual Determination of the Opacity of Emissions from Sta-
tionary Sources; 40 CFR 60, Appendix A; Government Printing
mental concerns. The formation of SO3 is complex, de- Office: Washington, DC, 1996. Available at: www.epa.gov/ttn/emc/
promgate/m-09.pdf (accessed April 2004).
pending upon fuel, operating parameters, and plant con- 10. Peterson, J.; Jones, A.F. SO3 Mitigation Guide; EPRI Report Number
figuration, and understanding the parameters leading to TR-104424; EPRI: Palo Alto, CA, October 1994.
11. Singer, J.G. Combustion—Fossil Power Systems, 3rd ed.; Combustion
excessive generation of SO3 and subsequent formation of Engineering Inc.: New York, 1981.
H2SO4 can assist in selections of practical control ap- 12. Burke, J.M.; Johnson, K.L. Ammonium Sulfate and Bisulfate Formation in
Air Preheaters; EPA-600/7– 82-025a; Industrial Environmental Research
proaches. Elevated SO3 concentrations can lead to corro- Laboratory, U.S. Environmental Protection Agency: Research Triangle
Park, NC, 1982.
sion, the formation of sulfite scale, and fouling and plug- 13. Zetlmeisl, M.J.; McCarthy, K.J.; Laurence, D.F. Investigation of Some
ging of low-temperature plant components, and can add Factors Affecting Nonflame Generation and Suppression of SO3; In-
dust. Eng. Chem. Prod. Res. Dev. 1983, 22, 710-716.
to the particle loading to control equipment. In some 14. Miller, C.A., U.S. Environmental Protection Agency. Personal commu-
nication with J. Bock of Siemens AG, Redwitz, Germany, June 18,
cases, elevated SO3 concentrations can lead to the forma- 2002.
tion of visible plumes at the stack exit or shortly down- 15. Clean Air Act Amendments of 1990; Public Law 101–549; Government
Printing Office: Washington, DC, November 15, 1990. Available at:
stream of the stack, resulting in noncompliance with local www.epa.gov/epahome/laws.htm (accessed April 2004).
regulations. 16. U.S. Environmental Protection Agency. Study of Hazardous Air Pollut-
ant Emissions from Electric Utility Steam Generating Units—Final Report,
SO3 can be formed in the boiler during the combus- Volume 1; EPA-453/R-98 – 049; Office of Air Quality Planning and
Standards: Research Triangle Park, NC, 1998.
tion of S-bearing fuels or downstream, particularly in SCR 17. U.S. Environmental Protection Agency. Air Quality Criteria for Particu-
reactors. SO3/H2SO4 can condense on low-temperature late Matter; EPA-600/P-95/001 (NTIS PB96 –168224); National Center
for Environmental Assessment: Research Triangle Park, NC, April
components and gets adsorbed by fly ash. Adsorption is 1996.
much greater in the high alkaline ashes resulting from 18. U.S. Environmental Protection Agency. Consolidated Emissions Re-
porting; Fed. Reg. 2002, 67, 39602-39616.
sub-bituminous coal firing. SO3/H2SO4 traditionally has 19. Alzueta, M.U.; Bilbao, R.; Glarborg, P. Inhibition and Sensitization of
Fuel Oxidation by SO2; Combust. Flame 2001, 127, 2234-2251.
been measured using extractive controlled condensation 20. Senior, C.L.; Sarofim, A.F.; Zeng, T.; Helble, J.J.; Mamani-Paco, R. Fuel
methods. Process. Technol. 2000, 63, 197-213.
21. Warnatz, J.; Maas, U.; Dibble, R.W. Combustion; Springer-Verlag: New
If needed, SO3/H2SO4 emission can be mitigated us- York, 1996; pp 67-71.
ing a variety of methods. Injection of alkali materials into 22. Hardman, R.; Stacy, R.; Dismukes, E. Estimating Sulfuric Acid Aerosol
Emissions from Coal-Fired Power Plants. Presented at the DOE-FETC
the furnace, either with the fuel or in slurry form, has Conference on Formation, Distribution, Impact, and Fate of Sulfur
Trioxide in Utility Flue Gas Streams, Pittsburgh, PA, March 1998.
resulted in reductions of up to 80%. Post-furnace injec- 23. Marier, P.; Dibbs, H.P. The Catalytic Conversion of SO2 to SO3 by Fly
tion of alkali materials can achieve up to 90% reductions Ash and the Capture of SO2 and SO3 by CaO and MgO; Thermochim.
Acta 1974, 8, 155-165.
but can increase particle loadings and ash resistivity char- 24. Monroe, L. An Updated Method for Estimating Total Sulfuric Acid Emis-
sions from Stationary Power Plants; Research and Environmental Affairs
acteristics. NH3 injection can also reduce SO3/H2SO4 by Department, Southern Company Generation and Electricity Market-
⬍90% and may result in increased particle loading to the ing: Birmingham, AL, 2001.
25. Nova, I.; dall’Acqua, L.; Lietti, L.; Giamello, E.; Forzatti, P. Study of
downstream collection systems. In plants with adequate Thermal Deactivation of a De-NOx Commercial Catalyst; Appl. Catal.
operational and equipment flexibility, fuel switching and B: Environ. 2001, 35, 31-42.
26. Forzatti, P.; Nova, I.; Beretta, A. Catalytic Properties in DeNOx and
blending can be used to reduce formation and emissions SO2-SO3 Reactions; Catal. Today 2000, 56, 431-441.
27. Dunn, J.; Koppula, P.R.; Stenger, H.G.; Wachs, I. Oxidation of Sulfur
of SO3. WESPs are also an option for control of SO3/ Dioxide to Sulfur Trioxide over Supported Vanadia Catalysts; Appl.
H2SO4, and a variety of designs have been successfully Catal. B: Environ. 1998, 19, 103-117.
28. Dunn, J.; Stenger, H.G.; Wachs, I. Oxidation of Sulfur Dioxide over
demonstrated for collection of acid mists and opacity Supported Vanadia Catalysts: Molecular Structure-Reactivity Relation-
control. ships and Reaction Kinetics; Catal. Today 1999, 51, 301-318.
29. Dunn, J.; Stenger, H.G.; Wachs, I. Molecular Structure—Reactivity
Relationships for the Oxidation of Sulfur Dioxide over Supported
Metal Oxide Catalysts; Catal.Today 1999, 53, 543-556.
REFERENCES 30. Gutberlet, H.; Hartenstein, H.; Licata, A. SO2 Conversion Rate of
1. Combustion Engineering; Fryling, G.R., Ed.; Combustion Engineering DeNOx Catalysts—Effects on Downstream Plant Components and
Inc.: New York, 1966. Remedial Measures. Presented at PowerGen 1999, New Orleans, LA,
2. Hawthorne, M. Plant Spewing Acid Clouds; Columbus Dispatch, August 1999.
3, 2001, p1C. 31. Svachula, J.; Alemany, L.J.; Ferlazzo, N.; Forzatti, P.; Tronconi, E.;
3. Verhoff, F.H.; Banchero, J.T. Predicting Dew Points of Flue Gases; Bregani, F. Oxidation of SO2 to SO3 over Honeycomb DeNOxing
Chem. Eng. Prog. 1974, 70 (8), 71-72. Catalysts; Ind. Eng. Chem. Res. 1993, 32, 826-834.

Volume 54 June 2004 Journal of the Air & Waste Management Association 761
Srivastava, Miller, Erickson, and Jambhekar

32. Levy, E.K. Effect of Boiler Operations on Sulfuric Acid Emissions. 45. Radway, J.E.; Exley, L.M. A Practical Review of the Cause and Control
Presented at the DOE-FETC Conference on Formation, Distribution, of Cold End Corrosion and Acidic Stack Emissions in Oil-Fired Boilers.
Impact, and Fate of Sulfur Trioxide in Utility Flue Gas Streams, Pitts- Presented at the Winter Meeting of the ASME, Atlanta, GA, December
burgh, PA, March 1998. 6 –10, 1975.
33. Sarunac, N.; Levy, E. Factors Affecting Sulfuric Acid Emissions from 46. Hutcheson, L.; Martin, M.; Ake, T.; Erickson, C.A.; Nystedt, P.; McFar-
Boilers. Presented at the EPRI/DOE/EPA Air Pollution Control Sympo- land, E.; Favor, C. Selective Catalytic Reduction System Performance at
sium: The Mega Symposium, Atlanta, GA, August 16 –20, 1999. Duke Energy’s Cliffside Unit 5 Power Station. Presented at the EPRI
34. Devito, M.S.; Oda, R.L. Flue Gas SO3 Stratification at ESP Inlets. Pre- 2002 Workshop on Selective Catalytic Reduction, Atlanta, GA, Octo-
sented at the DOE-FETC Conference on Formation, Distribution, Im- ber 22–23, 2002.
pact, and Fate of Sulfur Trioxide in Utility Flue Gas Streams, Pitts- 47. Altman, R.; Offen, G.; Buckley, W.; Ray, I. Wet Electrostatic Precipita-
burgh, PA, March 1998. tion Demonstrating Promise for Fine Particulate Control-Part I; Power
35. Yilmaz, A.; Pechulis, M.; D’Agostini, M.D.; Levy, E.; Sarunac, N.; Eng. 2001 Jan, 37-39.
Maust, P. Cold End Acid Condensation in Tubular Air Preheaters. In 48. Kumar, K.S.; Mansour, A. Wet ESP for Controlling Sulfuric Acid Plume
Proceedings of 1995 Joint Power Generation Conference, PWR-Vol. 28; Following an SCR System. Presented at the ICAC Forum 2002, Hous-
American Society of Mechanical Engineers: New York, NY, 1995; Vol. ton, TX, February 12–13, 2002.
3, pp 473-480. 49. Altman, R. Wet Electrostatic Precipitation Demonstrating Multiple
36. D’Agostini, M.; Levy, E.; Wilson, R. A Method for Calculating Acid Pollutant Control in Industrial Applications Holds Promise for Coal-
Deposition Rates in Regenerative Air Preheaters. In Heat Transfer Equip- Fired Utility Emission Reduction of Acid Mist, PM2.5, and Mercury.
ment Fundamentals, Design Application and Operating Problems; Sath, R., Presented at the U.S. EPA/DOE/EPRI Combined Power Plant Air Pol-
Ed.; ASME-HTD: New York, NY, 1990; Vol. 108, pp 78-87. lution Control Symposium: The Mega Symposium; Chicago, IL, Au-
37. Blythe, G.; Galloway, C.; Rhudy, R. Flue Gas Sulfuric Acid Measure- gust 20 –23, 2001.
ment Method Improvements. Presented at the U.S. EPA/DOE/EPRI 50. Srivastava, R., U.S. Environmental Protection Agency. Personal com-
Combined Power Plant Air Pollution Control Symposium: The Mega munication with J. Reynolds of Croll-Reynolds Co. Inc., Westfield, NJ,
Symposium; Chicago, IL, August 20 –23, 2001. August 8, 2002.
38. Gutberlet, H. Pleasant Prairie Power Plant SO3 Measurements at Different
Test Locations; Riley Power Inc. Internal Report by E. On Engineering,
Report Number 01– 01-100962; Riley Power: Worcester, MA, January
2002.
39. Blythe, G.; McMillan, R.; Davis, J.; Lamison, M.; Beeghly, J.; Benson,
L.; Goetz, E.; Rhudy, R. Furnace Injection of Alkaline Sorbents for
Sulfuric Acid Control. Presented at the U.S. EPA/DOE/EPRI Combined
Power Plant Air Pollution Control Symposium: The Mega Symposium, About the Authors
Chicago, IL, August 20 –23, 2001.
40. Air & Waste Management Association. Air Pollution Engineering Man-
R.K. Srivastava and C.A. Miller are at the Office of Research
ual, Ed. 2; Davis, W., Ed.; Wiley & Sons: New York, NY, 2000. and Development, National Risk Management Research
41. Lisle, E.S.; Sensenbaugh, J.D. The Determination of Sulfur Trioxide Laboratory, Research Triangle Park, NC. C. Erickson and R.
and Acid Dew Point in Flue Gases; Combustion 1965, 36, 12-16.
42. Dismukes, E.B. Conditioning of Fly Ash with Sulfur Trioxide and Ammo-
Jambhekar are at Babcock Borsig Power, Inc., Worcester,
nia; EPA-600/2–75-015 (NTIS PB-247231); U.S. Environmental Protec- MA. Address correspondence to: R.K. Srivastava, Office of
tion Agency, Office of Energy, Minerals, and Industry: Research Tri- Research and Development, National Risk Management
angle Park, NC, 1975.
43. Erickson, C., Babcock Borsig Power Inc. Personal communication with Research Laboratory, Air Pollution Prevention and Control
H. Gutberlet of E. On Engineering, Gelsenkrichen, Germany, July 7, Division, U.S. Environmental Protection Agency, Research
2002. Triangle Park, NC, 27711; fax: (919) 541-0554; e-mail:
44. Reese, J.T.; Jonakin, J.; Caracristi, V.Z. Prevention of Residual Oil
Combustion Problems by Use of Low Excess Air and Magnesium srivastava.ravi@epa.gov.
Additive; J. Eng. Power 1965, Transactions of ASME, Series A, 87, 229-
236.

762 Journal of the Air & Waste Management Association Volume 54 June 2004

You might also like