Sei sulla pagina 1di 104

Continuous Measurement Technologies

for SO3 and H2SO4 in Coal-Fired Power


Plants

Technical Report

Effective December 6, 2006, this report has been made publicly available in
accordance with Section 734.3(b)(3) and published in accordance with Section
734.7 of the U.S. Export Administration Regulations. As a result of this publication,
this report is subject to only copyright protection and does not require any license
agreement from EPRI. This notice supersedes the export control restrictions and
any proprietary licensed material notices embedded in the document prior to
publication.
Continuous Measurement
Technologies for SO3 and H2SO4
in Coal-Fired Power Plants

1009812

Final Report, September 2004

EPRI Project Manager


C. Dene
R. Himes

EPRI • 3412 Hillview Avenue, Palo Alto, California 94304 • PO Box 10412, Palo Alto, California 94303 • USA
800.313.3774 • 650.855.2121 • askepri@epri.com • www.epri.com
DISCLAIMER OF WARRANTIES AND LIMITATION OF LIABILITIES
THIS DOCUMENT WAS PREPARED BY THE ORGANIZATION(S) NAMED BELOW AS AN
ACCOUNT OF WORK SPONSORED OR COSPONSORED BY THE ELECTRIC POWER RESEARCH
INSTITUTE, INC. (EPRI). NEITHER EPRI, ANY MEMBER OF EPRI, ANY COSPONSOR, THE
ORGANIZATION(S) BELOW, NOR ANY PERSON ACTING ON BEHALF OF ANY OF THEM:

(A) MAKES ANY WARRANTY OR REPRESENTATION WHATSOEVER, EXPRESS OR IMPLIED, (I)


WITH RESPECT TO THE USE OF ANY INFORMATION, APPARATUS, METHOD, PROCESS, OR
SIMILAR ITEM DISCLOSED IN THIS DOCUMENT, INCLUDING MERCHANTABILITY AND FITNESS
FOR A PARTICULAR PURPOSE, OR (II) THAT SUCH USE DOES NOT INFRINGE ON OR
INTERFERE WITH PRIVATELY OWNED RIGHTS, INCLUDING ANY PARTY'S INTELLECTUAL
PROPERTY, OR (III) THAT THIS DOCUMENT IS SUITABLE TO ANY PARTICULAR USER'S
CIRCUMSTANCE; OR

(B) ASSUMES RESPONSIBILITY FOR ANY DAMAGES OR OTHER LIABILITY WHATSOEVER


(INCLUDING ANY CONSEQUENTIAL DAMAGES, EVEN IF EPRI OR ANY EPRI REPRESENTATIVE
HAS BEEN ADVISED OF THE POSSIBILITY OF SUCH DAMAGES) RESULTING FROM YOUR
SELECTION OR USE OF THIS DOCUMENT OR ANY INFORMATION, APPARATUS, METHOD,
PROCESS, OR SIMILAR ITEM DISCLOSED IN THIS DOCUMENT.

ORGANIZATION(S) THAT PREPARED THIS DOCUMENT

University of California at Riverside

ORDERING INFORMATION
Requests for copies of this report should be directed to EPRI Orders and Conferences, 1355 Willow
Way, Suite 278, Concord, CA 94520, (800) 313-3774, press 2 or internally x5379, (925) 609-9169,
(925) 609-1310 (fax).

Electric Power Research Institute and EPRI are registered service marks of the Electric Power
Research Institute, Inc. EPRI. ELECTRIFY THE WORLD is a service mark of the Electric Power
Research Institute, Inc.

Copyright © 2004 Electric Power Research Institute, Inc. All rights reserved.
CITATIONS

This report was prepared by

College of Engineering-Center for Environmental Research and Technology


University of California
Riverside, CA 92521

Principal Investigators
J. Pisano
D. Fitz

This report describes research sponsored by EPRI.

The report is a corporate document that should be cited in the literature in the following manner:

Continuous Measurement Technologies for SO3 and H2SO4 in Coal-Fired Power Plants, EPRI,
Palo Alto, CA: 2004. 1009812.

iii
REPORT SUMMARY

The use of Selective Catalytic Reduction (SCR) technology to reduce NOx emissions from coal-
fired power plants can result in the oxidation of SO2 in the flue gas to SO3 with a number of
undesirable consequences, including the emission of acid aerosols from the stack leading to a
visible plume. Injected additives can control SO3, but no method is currently available to
continuously monitor SO3 and make it possible to optimize additive use. This report investigates
the possibility of developing new continuous SO3 measurement techniques, including an in-situ
method using Fourier Transform Infrared (FTIR) spectroscopy that is ready for field testing.

Background
Control of SO3 often entails the use of an additive injected into the flue gas. Control of the
additive feed rate, however, is based on manually collected SO3 measurements during initial
testing of additive use. No method is currently available to optimize the additives use over the
normal range and variability of unit operation. The availability of a continuous SO3 monitoring
method would provide significant cost savings by enabling the minimization of additive feed
rates used to control SO3 emissions. This minimum would be achieved when the resulting SO3
concentrations are reduced enough to avoid a visible plume, while remaining high enough to
condition the fly ash for high efficiency collection by the Electrostatic Precipitator. Continuous
accurate measurement of SO3 would also detect increases in SO3 levels that result from changes
in fuel specifications, combustion conditions, and other factors so that appropriate corrective
actions could be taken.

Objectives
• To discuss issues concerning SO3 in coal-fired power plants and describe present
measurement technologies for SO3 and H2SO4
• To investigate the possible use of Fourier Transform Infrared (FTIR) spectroscopy and
Differential Optical Absorption Spectroscopy (DOAS) for continuously measuring SO3 and
H2SO4
• To provide an annotated bibliography of literature available concerning SO3 issues for coal-
fired power plants.

Approach
The project team conducted an exhaustive literature search on spectroscopic information relating
to identification of SO3 or H2SO4 spectra, measurement techniques for SO3 and H2SO4
(specifically gas phase), or other SO3 related issues for coal fired power plants. The team
analyzed the limitations of current SO3 and H2SO4 measurement methods. They described two
possible alternatives, FTIR spectroscopy and DOAS, and discussed what needs to be done to
implement these technologies.

v
Results
SO3 has adverse effects on plant operations, primarily due to the formation of corrosive H2SO4 on
air heater and/or ductwork surfaces. SO3 can react with residual ammonia slip from SCR systems
and form ammonia bisulfate, which can lead to potential fouling of the air heater. SO3 also has
environmental consequences since it augments the formation of visible plumes and is a precursor
to the formation of particulate level PM2.5. The EPA requires US utilities to estimate sulfuric acid
emissions from their facilities to the Toxic Release Inventory (TRI). A reliable, in-situ SO3
monitor could help coal-fired boiler operators address these concerns by improving their overall
SO3 emission control while avoiding the problems associated with current methods.
For in-situ measurements, optical spectroscopic techniques have been shown to be a reliable
option. Both SO3 and H2SO4 can be monitored via Fourier Transform Infrared Spectroscopy
(FTIR) to the detection limits required for coal-fired applications. The dust loading in power
plant ducts requires that the FTIR technology be interfaced with a dust attenuating probe for in-
situ measurements. Such a probe has been devised by EPRI and is ready for field testing.
Differential Optical Absorption Spectroscopy (DOAS) also has promise for future development,
but currently available systems measure SO2 only and have not yet been configured for in-situ
measurements of SO3.

EPRI Perspective
EPRI is planning field tests of a phase discrimination probe that uses an FTIR analyzer to
conduct continuous in-situ measurement of SO3 over a line-of-sight within the flue gas duct
anywhere downstream of the economizer outlet. The performance of the monitor will be
compared against wet chemical reference methods to establish the accuracy of the method. For
details on this project, see EPRI Project Opportunity 1009582.

Keywords
Sulfur Trioxide (SO3)
Selective catalytic reduction (SCR)
Visible plumes
Continuous monitoring
Fourier transform infrared (FTIR)
Differential Optical Absorption Spectroscopy (DOAS)

vi
ABSTRACT

A continuous SO3 monitoring method would provide significant cost savings to the electric
utility industry by enabling power plant operators to minimize the use of additives to control SO3
emissions resulting from the operation of Selective Catalytic Reduction (SCR) systems. This
study examines issues concerning the measurement of SO3 in coal-fired power plants, and
describes the problems associated with presently available technologies for SO3 and H2SO4
measurement. Two optical methods, Fourier Transform Infrared (FTIR) spectroscopy and
Differential Optical Absorption Spectroscopy (DOAS), have the potential to allow for
continuous in-situ monitoring. An FTIR system interfaced with a dust-attenuating probe has
already been developed for in-situ measurements and will be field-tested in the near future.

vii
ACKNOWLEDGMENTS

The authors would like to thank Dr. Bob Spellicy of IMACC (Industrial Monitor and Control
Corporation of Austin, Texas) for his valuable assistance in discussing the possibility of using
FTIR (Fourier Transform Infrared Spectroscopy) technology for the measurement of SO3 and
H2SO4 and for providing a report showing their experience in conducting an SO3 measurement
experiment using FTIR. The authors would also like to thank Dr. Claudia Sauer, Mr. Matt Smith
and Mr. Adrian Afan of UCR for their assistance in locating many of the references used in the
formation of this report.

ix
CONTENTS

1 INTRODUCTION ....................................................................................................................1-1
1.1 Report Objectives............................................................................................................1-2
1.2 Formation of SO3 .............................................................................................................1-2

2 PRESENT MEASUREMENT TECHNIQUES FOR SO3 AND H2SO4 ......................................2-1


2.1 Corrosion Probes ............................................................................................................2-1
2.2 EPA Method 8, the Determination of Sulfuric Acid Mist and Sulfur Emissions From
Stationary Sources. ...............................................................................................................2-2
2.3 Modified EPA Method 8, the Determination of Sulfuric Acid Mist and Sulfur
Emissions From Stationary Sources. ....................................................................................2-3
2.4 Controlled Condensation Method....................................................................................2-3
2.5 Severn Science Analyser ................................................................................................2-5

3 PROBLEMS INVOLVED WITH CURRENT SO3 MEASUREMENT TECHNIQUES...............3-1

4 POSSIBLE CONTINUOUS MEASUREMENT TECHNOLOGIES FOR MEASURING


SO3 AND H2SO4 .........................................................................................................................4-1
4.1 Spectroscopy ..................................................................................................................4-1
4.2 FTIR (Fourier Transform InfraRed Spectroscopy)...........................................................4-1
4.3 DOAS (Differential Optical Absorption Spectroscopy) ..................................................4-13

5 CONCLUSION........................................................................................................................5-1

6 REFERENCE LIST FROM LITERATURE SEARCH..............................................................6-1


6.1 Spectroscopic References ..............................................................................................6-2
6.2 Measurement Technique References ...........................................................................6-10
6.3 SO3 Coal-Fired Power Plant Related References .........................................................6-11

xi
xii
LIST OF FIGURES

Figure 2-1 Schematic of Controlled Condensation SO3 Sampling System (CCS) .....................2-4
Figure 4-1 Typical Michelson Interferometer .............................................................................4-2
Figure 4-2 Example Interferogram of Ambient Air Containing High Level of CH4
(Courtesy of IMACC) ..........................................................................................................4-3
Figure 4-3 Interferogram of the Source Background (Left), and Resulting Transformed
Single Beam Spectrum (Right)...........................................................................................4-5
Figure 4-4 Interferogram of the Gas Sample Containing Styrene (Left), and Resulting
Transformed Single Beam Spectrum (Right) .....................................................................4-5
Figure 4-5 Resulting IR Spectrum of Styrene ............................................................................4-6
Figure 4-6 Plot of Experimental Data Taken by IMACC at Their Laboratory Facility
Investigating SO3 IR Spectra Compared to the SO3 Reference (middle) and the SO2
Reference (bottom). (By Permission of IMACC) ................................................................4-8
Figure 4-7 Plot Of Experimental Data Taken by IMACC at Their Laboratory Facility
Investigating H2SO4 IR Spectra Compared to the H2SO4 Reference (Bottom) in the
-1
1100 to 1260 Cm Region. (By Permission of IMACC)......................................................4-9
Figure 4-8 Plot of Experimental Data Taken by IMACC at Their Laboratory Facility
Investigating H2SO4 IR Spectra Compared to the H2SO4 Reference (Bottom) in the
-1
800 to 940 Cm Region. (By Permission of IMACC).......................................................4-11
Figure 4-9 Plot of Experimental Data Taken by IMACC at a Coal Fired Power Plant in
the Unit Duct. The Error Bars are the 3 Sigma Errors Returned by the FTIR (With
Permission of IMACC)......................................................................................................4-12
Figure 4-10 DOAS Technique , in Which I 0 is Replaced by I 0' , the Light Intensity in the
Absence of a Structured Absorption Band. ......................................................................4-14
Figure 4-11 Components of the UCR Developed DOAS System Employed for the
Measurement of Low Level (30-100 ppbV) SO2 From Vehicle Exhaust...........................4-15
Figure 4-12 The Absorbance Spectra Over the Wavelengths From 220 nm Up to 300 nm
for SO3, for 100 ppmV of SO3, a Five-Meter Sample Cell and a Diode Array
Spectrometer....................................................................................................................4-17
Figure A-1 Optical Configuration Employed at Gas Fired Power Plant for Continuous
TDL Measurements of NH3 ....................................................................................................... A-6
Figure A-2 NH3 Slip Measured Compared to Reagent Flow at the Gas Fired Utility Over
a Two-week period............................................................................................................ A-7
Figure A-3 NH3 Slip Measured Compared to Reagent Flow, Load and Reagent Dilution
(H2O) at a 700 MW Coal Fired Power Plant ..................................................................... A-9

xiii
Figure A-4 Linear Regression Plot Showing the Inter-Comparison Between the Wet
Chemical NH3 Measurements to That as Measured by the Probe Tunable Diode
Laser NH3 Measurements. .............................................................................................. A-10
Figure A-5 Shows the NH3 Slip Results and Correlation Coefficient of the TDL Data
During the Adjustment of Urea Flow With Newly Installed Injectors. .............................. A-12
Figure A-6 Shows the Which Shows the Laser Signal Power as Measured With the
Probe and Without the Probe Over Pathlengths Up to 10 Meters................................... A-14

xiv
LIST OF TABLES

Table A-1 Lists Results From the Traverse of All 8 Ports. ..................................................... A-13

xv
1
INTRODUCTION

NOx emissions from coal-fired boilers in ozone non-compliant regions are being restricted to
levels that often require the application of Selective Catalytic Reduction (SCR) technology. One
consequence of the catalyst within SCR systems is the oxidation of a fraction of the SO2 in the
flue gas to SO3. A number of balance-of-plant impacts can result from the increased SO3
emissions, which include (a) increased propensity to form ammonium bisulfate within the air
heater, (b) corrosion of downstream surfaces at temperatures below the acid dew point, and (c)
increased opacity due to the emission of acid aerosols from the stack.

Control of SO3 often entails the use of an additive injected into the flue gas. Control of the
additive feed rate, however, is based on manually collected SO3 measurements during initial
testing of additive use. No method is currently available to optimize the additive use over the
normal range and variability of unit operation. The availability of a continuous SO3 monitoring
method would provide significant cost savings by enabling the minimization of additive feed
rates used to control SO3 emissions. This minimum would be achieved when the resulting SO3
concentrations are reduced enough to avoid a visible plume, while still being high enough to
condition the fly ash for high efficiency collection by the ESP.

Even when additives are not utilized, accurate measurement of SO3 emissions can serve to
monitor and detect an increase in SO3 levels resulting from a change in fuel specifications,
combustion considerations, SCR operation etc., so that appropriate corrective actions can be
taken. A continuous SO3 monitor would also help plants that burn a low-sulfur coal to minimize
the amount of SO3 injected to condition fly ash for high ESP removal efficiencies.

The formation of sulfur compounds is an integral part of the combustion of coal in power plants.
Most of the fuel sulfur is oxidized to sulfur dioxide (SO2), while a small percentage can be
further oxidized to sulfur trioxide (SO3) depending on the temperature and nature of the reaction.
Though levels of SO3 are nominally 1-25 ppmv, it is very reactive and forms H2SO4 which can
promote corrosion.

SO3 levels can also be substantially increased by the use of NOX control technologies. Since
Selective Catalytic Reduction units (SCR’s) are becoming more prevalent in the industry this is
becoming an area of concern. This increase in SO3 formation can create further problems by
reacting with unreacted ammonia, leading to the formation of visible plumes as well as fouling
of air heaters and ductwork.

Finally, the US Environmental Protection Agency has listed SO3 in its Toxic Release Inventory
(TRI) requiring the determination of sulfuric acid emissions from power plants. Since 1998, US
utilities report their annual releases of toxic chemicals to the TRI.

1-1
Introduction

1.1 Report Objectives

This report was commissioned with five goals in mind;

1. To briefly discuss issues concerning sulfur trioxide (SO3) with respect to coal-fired power
plants.

2. To list and describe present measurement technologies for SO3 and H2SO4

3. To describe some of the problems involved with current SO3 measurement techniques

4. To investigate the possibility of utilizing other measurement technologies for continuously


measuring SO3 and H2SO4 and describe the requirements for implementation.

5. To list and reference current literature available concerning SO3 issues for coal fired power
plants

1.2 Formation of SO3

The formation of sulfur compounds is an integral part of the combustion of coal in power plants.
Most of the fuel sulfur is oxidized to SO2 while a small percentage can be further oxidized to
SO3 depending on the temperature and nature of the reaction. In the case of homogeneous gas
phase reactions below temperatures of 1650ºF (900ºC), the formation of SO3 is characterized by
Reaction 1 (Squires, 1982).

SO2 + O2 → SO3 + O Equation 1-1

At higher temperature conditions greater than 3100ºF (1700ºC), SO2 is favored; however, the
rapid cooling of the combustion products can shift the equilibrium to increase the formation of
SO3. This reaction rate decreases with temperature until equilibrium can no longer be
maintained. The residual SO3 can also be adsorbed by the ash depending on the alkalinity of the
ash.

The formation of sulfuric acid can result from the combination of water vapor and SO3 at
temperatures below 750ºF (400ºC). As the additional water combines with the H2SO4 vapor, an
aerosol can form in the air heater. The design of many utility boilers incorporate regenerative air
heaters to transfer heat from the gases leaving the economizer to the inlet air flowing to the
burners. When the flue gas is cooled from levels greater than 660ºF (350ºC) to 300ºF (150ºC),
SO3 and water vapor combine to form H2SO4. If the temperature is sufficiently low, the H2SO4
can reach a saturation point and condense, or remain in equilibrium with the vapor phase H2SO4.
This is particularly difficult to model or characterize when the load is changing and the
equilibrium is similarly variable, making dewpoint predictions difficult.

The effect of dew point variability is directly related to the formation of an adherent film on the
air heater and other cold duct surfaces. The resultant corrosion contributes to the formation of
iron sulfates and increases the deposition rates of unburnt fuel particles which leads to further
fouling of the air heater surfaces (Johnston, 1982). As might be expected the corrosion rates are

1-2
Introduction

proportional to the rates of deposition. These accumulations can absorb up to 30% of their total
weight in acid.

The rate of absorbtion can range from 0.1 to 0.5 grams acid/kg ash for eastern bituminous coal.
The use of lower sulfur content sub bituminous coal, containing 0.5% sulfur, with higher alkali
content will effectively adsorb most of the sulfuric acid while the bituminous coal with higher
sulfur content may not adsorb the high concentration amounts (Levy, 1998).

Deposition rates of ash and the acid dew point temperature combined can have a significant
impact on the surface corrosion rates. The corrosion mechanism in utility applications is based
on thin film deposition as opposed to more typical acid immersion. This mechanism also is
directly affected by the acid dew point temperature in the surrounding gas. The acid deposition
and corrosion rate will reach a maximum at approximately 54ºF (12ºC) below the acid dewpoint.

The formation of SO3 can have the following adverse effects on both the plant operation and the
surrounding environment.
1. Formation of particulate level PM2.5, a known health hazard (Sanyal and Parker, 1999).

2. Degradation of the SCR vanadium catalyst and ultimately deactivation of the SCR unit
(Duellman et al, 2002).

3. Corrosion of air heater and/or ductwork surfaces (Halstead and Talbot, 1980).

4. Formation of ammonium bisulphate due to ammonia slip from SCR or SNCR processes
(Bionda, 2002).

5. Formation of visible plumes (Mueller and Imhoff, 1994)

6. Flyash contamination (Bayless et al, 1996)

Although of significance to the operation of coal-fired boilers with medium to high sulfur
contents, these issues are not the focus of the current report. A recent article in the Air & Waste
Management Association summarizes the factors affecting SO3/H2SO4 emissions from coal-fired
boilers, as well as approaches for mitigating these emissions (Srivastava et. al., 2004). The
following focuses on present measurement techniques for SO3, as well as their inherent
problems. Two potential continuous SO3 measurement approaches are then assessed for their
potential application to coal-fired boiler flue gas streams.

1-3
2
PRESENT MEASUREMENT TECHNIQUES FOR SO3
AND H2SO4

Historically, there are several analytical techniques for determining SO3 concentrations in flue
gas. This section will briefly describe these conventional methods and indicate which of the
techniques is most commonly used today. The subsequent section will describe two novel
measurement methods for SO3 and H2SO4, as well as outline the requirements and potential
advantages of each. The two new methods are both optically based and will be discussed
objectively as to their potential for continuous SO3 monitoring. The two spectroscopic
techniques are FTIR (Fourier Transform Infrared Spectroscopy), the other being DOAS
(Differential Optical Absorption Spectroscopy).

On a whole, it is has been found that the measurement of SO3 is quite difficult. This is primarily
due to:
1. the low concentrations of SO3 present in coal-fired boiler flue gas streams,
2. SO3 tends to adsorb onto surfaces,
3. most measurement techniques of SO3 are based upon the result of SO2 oxidization and can be
biased due to the presence of a much higher concentration of SO2.

2.1 Corrosion Probes

Corrosion probes have been designed that allow for estimation of SO3 levels by direct analysis of
the probe for corrosion after a period of time (Littler et al., 1992). For these types of probes the
temperature needs to be recorded and the corrosive effects are measured based on the time period
that the probe was in the corrosive environment. The material of the probe was investigated for
corrosive effects based on time and temperature, and a scale was set to yield approximate acid
gas levels. The probe developed was designed for removal after 2-3 hours based on typical SO3
levels.

A similar type of corrosion probe was designed that employed weight loss measurement of a test
piece that was inserted into the probe itself. The test piece under examination was insulated
from the probe via a refractory collar. The test piece is also air cooled so as to be thermally
isolated from the probe. The corrosion rate is normally determined by the weight loss of the test
piece. These type of test pieces have also been designed as corrosion test discs, whereby direct
analysis of the corrosion can be done by examining the corrosive properties of the discs
themselves.

Another type of corrosion probe was developed by the British Coal Utilisation Research
Association. This type of probe contains a stainless steel lance, and also requires temperature

2-1
Present Measurement Techniques for SO3 and H2SO4

monitoring. The probe is inserted directly into the flue gas, and after about 30 minutes is
removed with condensed acid washed from each of the sections. The washings are titrated for
total acidity and a graph of acid deposition can be plotted against temperature. Using a graph,
the maximum rate of build up of acid along with the dewpoint temperature can be estimated.
Though this technique does not give a direct readout of dewpoint or SO3 concentration, it does
give a good indication of the corrosive potential of the flue gas.

2.2 EPA Method 8, the Determination of Sulfuric Acid Mist and Sulfur
Emissions From Stationary Sources.

EPA Method 8, “The Determination of Sulfuric Acid Mist and Sulfur Emissions from Stationary
Sources”, was designed to estimate acid mist emissions from sulfuric acid plant gas streams that
typically contain very little moisture and insignificant amounts of other contaminants. Below is
a brief description of this technique.

Principle And Applicability

Principle. A gas sample is extracted isokinetically from the stack. The sulfuric acid mist
(including sulfur trioxide) and the sulfur dioxide are separated, and both fractions are measured
separately by the barium-thorin titration method.

Applicability. This method is applicable for the determination of sulfuric acid mist (including
sulfurtrioxide (SO3) in the absence of other particulate matter) and sulfur dioxide (SO2)
emissions from stationary sources.

Minimum detection limits: Collaborative tests have shown that the minimum detectable limits of
the method are 0.05 mg/m3 (0.03 x 10-7lb/ft3) for SO3 and 1.2 mg/m3 (0.74 x 10-7lb/ft3) for SO2. No
upper limits have been established. Based on theoretical calculations for 200 ml of 3 percent
hydrogen peroxide solution, the upper concentration limit for SO2 in a 1.0 m3(35.3 ft3) gas sample
is about 12,500 mg/m3 (7.7 x 10-4lb/ft3). The upper limit can be extended by increasing the
quantity of peroxide solution in the impingers.

Interferences: Possible interfering agents of this method are fluorides, free ammonia, and
dimethylaniline. Particulate sulphate salts such as NaSO4 and CaSO4 will also cause a positive
bias to the SO3 measurement. If any of these interfering agents are present (this can be
determined by knowledge of the process), alternative methods, are required.

For Method 8, an isokinetic sample of flue gas is extracted through a heated glass-lined probe
into a series of impingers. In the first of the impingers resides a solution of 80% isopropyl
alcohol (IPA), the solution is cooled in an ice bath. By keeping the isopropyl alcohol solution
cool, condensation of the H2SO4 is nearly complete and collected by the isopropyl alcohol
solution. A filter is inserted downstream of the alcohol impinger and any uncollected H2SO4
aerosol is trapped by this filter. Isopropyl alcohol is used because it inhibits the interferences
from SO2, this is accomplished because of the limited solubility of SO2 in organic alcohols and
the additional fact that alcohol inhibits the oxidation of the dissolved SO2. This method also
measures SO2 by placing two impingers containing a 3% solution of hydrogen peroxide
downstream of the first impinger.

2-2
Present Measurement Techniques for SO3 and H2SO4

SO3 and H2SO4 are captured as SO42- in the front half of the sampling train which includes the
isopropyl alcohol containing first impinger and the filter. The SO2 is oxidized to sulfate ions in
the two impingers containing the hydrogen peroxide solution.

The method is quite limited and not suitable for measurements in coal-fired utility boilers,
because a small amount of SO2 is absorbed by the isopropyl alcohol in the first impinger and is
oxidized to SO3. Since it has been found that for coal-fired boilers the concentration of SO2 can
be up to a thousand times that of SO3, this positive bias error introduced by the sampling train
can actually exceed the concentration of SO3 in the flue gas. As of this time, efforts have been
made to further inhibit the oxidation of SO3 using antioxidants, the results have been mixed and
overall have not proved successful.

Apparatus: Commercial models of this train are available. For those who desire to build their
own, however, complete construction details are described in APTD-0581 (Construction Details
of Isokinetic Source-Sampling Equipment, EPA, 1971, document#PB-203-060). The operating
and maintenance procedures for the sampling train are described in APTD-0576 (Maintenance,
Calibration and Operation of Isokinetic Source Sampling Equipment, EPA, 1972, document
#PB-209-022).

2.3 Modified EPA Method 8, the Determination of Sulfuric Acid Mist and
Sulfur Emissions From Stationary Sources.

As described above in the description of the EPA Method 8, this method was designed to
measure sulfuric acid emissions from sulfuric acid plants. These type of plants do not contain
water vapor, as it is not a combustion process and not applicable for use in utility flue gas which
does contain moisture. The major problem with measuring SO3 emissions from utility flue gas
with Method 8 stem from the fact that SO2, is absorbed by water as it condenses in the
isopropanol impinger and therefore can result in a positive bias to the measured SO3 values.

A modified EPA Method 8 was developed that employs a cyclone separator on the inlet end of
the probe to remove water droplets. This cyclone separator in theory allows the sulfuric acid mist
to pass through it and be collected on the walls of the probe and on a high purity filter located at
the outlet end of the probe. The filter collects any particulate matter and sulfuric acid mist
which is subsequently analyzed. The probe and the filter are maintained at a temperature slightly
above that of the stack gas.

The analysis involves a barium perchlorate titration using thorin as an indicator or alternatively
may be performed using ion chromatography. The cyclone catch and filter are analyzed for
calcium by ion chromatography to determine scrubber mist droplet carryover. The modified
method has not proven to be highly accurate and has led to the development of other methods as
described in the next section (Jackson et al., 1981).

2.4 Controlled Condensation Method

The Controlled Condensation Method was designed to measure the vapor phase concentration of
SO3 and H2SO4 flue gas streams. The sampling system and methodology was originally

2-3
Present Measurement Techniques for SO3 and H2SO4

developed by Goksoyr and Ross. It consists of a heated glass-lined probe which is followed by a
quartz filter nominally heated to at least 500°F (260 C). The sample then passes through a glass
coil condenser which acts as a liquid-cooled condenser and allows the controlled condensation of
H2SO4. The liquid is maintained in the temperature range of 150°F (66 C). A schematic of a
system is shown in Figure 2-1. The sample is cooled in the condenser to the acid dewpoint
temperature at which the H2SO4 condenses. The gas temperature is kept above the water
dewpoint to prevent any interference from SO2 while the heated quartz filter system removes
particulate matter. The condensed acid is titrated with barium perchlorate using thorin as the
indicator.

For particulate laden flue gas, SO3 testing is performed using a controlled condensate sampling
system. The system includes a probe which is heated to 550°F (288 C). A quartz wool filter is
used in the probe tip to separate particulate matter from the sample stream. The entire probe
apparatus is heated to the target temperature prior to initiating the test. The sample gas is
extracted with a nonreactive quartz probe that is maintained above the acid dew point
temperature. Quartz wool plugs are used at the tip of the sample probe to remove particulate
matter. The sample gas is filtered and passed through cooling coils and a frit, which is
maintained at a temperature between the acid and water dew points (approximately 145-165°F ;
(63 – 74 C).This approach allows the condensation of SO3 while preventing the condensation of
water and associated SO2 capture. The sample is withdrawn at a rate of nominally 0.25 scfm. A
total sample volume of 7 - 8 scf is typical, requiring a 30-minute sample time.

Figure 2-1
Schematic of Controlled Condensation SO3 Sampling System (CCS)

2-4
Present Measurement Techniques for SO3 and H2SO4

At the completion of each test, the coil is rinsed with distilled water. An aliquot of each sample
is sent to be analyzed for sulfate.

One issue with the current tests is the potential reaction of SO3 with the sorbent collected on the
filter. If the unreacted sorbent were to combine with SO3, the resulting sulfate would be trapped
in the filter. The SO3 value for the test would not include this portion and would therefore be
biased low. To account for this one typically collects the filters from each test run and stores
them separately in distilled water. One filter solution sample from each set of triplicate tests
would also be analyzed for sulfates.

2.5 Severn Science Analyser

The Severn Science analyser was developed by the Central Electricity Generating Board of
England over 20 years ago for continuously measuring SO3 in flue gases (Jackson et al., 1981) .

The instrument was originally designed to be portable and have a range of 0 - 50 ppm SO3.

The principle of operation for this analyzer is based on the following:

• SO3 or H2SO4 in the flue gas is absorbed as sulphate ions in an aqueous solution of propanol
in water.
• This solution is subsequently passed through barium chloranilate resulting in the following
reaction taking place :

SO4 2- + BaC6O4C12 + H+ → BaSO4 + HC6O4Cl2¯ Equation 2-1


¯
The resulting acid chloranilate ions (HC6O4Cl2 ) formed from the above reaction absorb light
at 535 nm and can be continuously measured spectroscopically in a flow photometer
• If the ratio of flow rates for the flue gas and the IPA absorbing solution are kept constant, the
concentration of the chloranilate ions can be directly related to the sulphate ion concentration
• Since the sulfate ion is formed from either SO3 or H2SO4 in the flue gas, then the chlorinate
ions can subsequently be related to the concentration of SO3 or H2SO4
• The method of sampling is extractive and the flue gas is sampled through a silica tube which
is heated to 390ºF (200ºC) to prevent acid condensation.
• The silica tube is connected to a heated filter to remove particulate matter.
• The gas makes contact with the IPA absorbing solution in a heated block
• The solution flows through the barium chloranilate reservoir bed at a constant rate where the
sulphate ions are retained and acid chloranilate ions are released as described by Equation
2-1 above.
• Prepared calibration solutions of dilute sulfuric acid are used as standards.

2-5
Present Measurement Techniques for SO3 and H2SO4

A major potential advantage of this technique is that interference from other flue gas constituents
is decreased. It should be noted, however, that although the instrument has been used in a
number of power plants in the USA, plant operators report that since the instrument requires the
presence of a very skilled operator there has been relatively insignificant demand for the
instrument.

Published data on the accuracy and repeatability of the Severn Science SO3 analyser were
difficult to find and are not readily available. Although the manufacturer maintains that the
system provides accurate results over the operating range for typical coal fired flue gas SO3
measurements, the measurement is extractive and will suffer from similar problems as the other
methods described. Further information may be obtained from their website address:
www.severnscience.co.uk.

2-6
3
PROBLEMS INVOLVED WITH CURRENT SO3
MEASUREMENT TECHNIQUES

As has been discussed, the measurement of SO3 is quite difficult primarily due to the low
concentrations of SO3 and its reactive nature. Most measurement techniques can yield biased
results due to their extractive sampling procedures, as well as by the presence of much higher
concentrations of SO2. There are also two other considerations which can make present
measurement technologies questionable in providing accurate representative values for SO3 or
H2SO4.

One of these considerations include the presence of SO3 in the form of an acid aerosol (H2SO4)
which can adsorb onto surfaces. The other is that in addition to the above measurement issues,
the techniques described represent point measurements, which by their nature assume a
homogeneous distribution of the SO3 in the flue gas duct. This assumption is likely not
representative of most boiler flue gas streams.

Most of the problems involved with accurate measurements of SO3 stem from the extractive
nature of the sampling techniques employed. SO3 is highly reactive and tends to adsorb onto
surfaces. In fact, a considerable amount of the SO3 in the flue gas will react rapidly with water
vapor to form H2SO4 vapor. This H2SO4 vapor will also combine with additional water molecules, to
form liquid aerosol droplets should the flue gas temperatures be below the acid dewpoint.

A considerable amount of variability has been experienced in trying to estimate how much of the
SO3 is formed as H2SO4 aerosol in power plant flue gases. Approximations have been made by
estimating that the proportion of H2SO4 aerosol formed depends upon the moisture content of the
gas, as well as the temperature and concentration of H2SO4 in the vapor phase (Bionda, 2002).
These aerosol droplets will not immediately vaporize, even in the flue gas stream where the
average temperature is far above the acid dewpoint. With this in mind, extractive techniques,
specifically the controlled condensation method that incorporates a pre-filter to prevent
particulates affecting the SO3 determination, could also remove acid aerosol droplets, thereby
reducing the SO3 measurement.

Stratification has long been a concern in making single point or path measurements of a target
gas in the effluent gas at coal fired power plants (Bionda, 2002). The fabrication and handling of
long glass probes is not practical, and multiple sampling ports are frequently limited by
accessibility for this type of measurement. Since current SO3 measuring techniques based on wet
chemistry are not appropriate for multiple point sampling due to lengthy turnaround times, what
would be best is to have an in-situ type measurement of SO3 in conjunction with characterization
of the duct work so as to choose a path that would be representative of the SO3 concentration
throughout the duct.

3-1
4
POSSIBLE CONTINUOUS MEASUREMENT
TECHNOLOGIES FOR MEASURING SO3 AND H2SO4

Based on the literature search conducted for this project, as well as meetings with a company
involved with spectroscopic measurements (i.e. IMACC), two potential approaches were
identified with the potential for making continuous SO3 measurements.

4.1 Spectroscopy

Spectroscopy can be defined as the investigation of the interaction that occurs between a defined
source of electromagnetic radiation with a target sample. The way the emissive properties of the
radiative source, usually some sort of light source, are perturbed depends upon the sample
investigated. For example, molecules of gas can absorb electromagnetic radiation leading to
molecular vibrations or rotations, with the frequencies that are absorbed by quantum theory
being unique to each different type of molecule in the target sample. Therefore each molecule
has its own characteristic absorption pattern over the electromagnetic spectrum. For the
investigation of SO3 two possible spectroscopic techniques were examined, which when applied
to SO3 entail absorption features over two distinctively different regions in the electromagnetic
spectrum. The techniques will be discussed along with their potential use in coal fired
applications.

Spectroscopic methods basically evaluate the concentration of the molecule investigated through
Beer-Lambert’s Law, which states that the fraction of light intensity transmitted through a gas is
given by:

I/Io = exp (-σ(ψ)NL) Equation 4-1

Where I and Io are the transmitted and incident powers respectively, L is the absorption path
length (cms), and N would then be the concentration of the absorbing molecules in number of
molecules per cubic centimeter. In this example σ(ψ) is the wavenumber dependent absorption
cross section in square centimeters per molecule.

4.2 FTIR (Fourier Transform InfraRed Spectroscopy)

Fourier transform infrared spectroscopy began to come into its own in the early 1950s when
experimental groups first built and tested high resolution spectrometers. Today, commercial
Fourier transform spectrometers are widely available. Aided by fast computers which perform
Fourier transforms quickly, FTIR spectrometers are used to make spectroscopic measurements in
many diverse disciplines. This has led to FTIR technology being used in the development of
many commercial instruments by companies such as IMACC, Nicolet, Midac, Bruker, Bomem,

4-1
Possible Continuous Measurement Technologies for Measuring SO3 and H2SO4

Unisearch Associates and ABB to name a few. Most of the commercial instruments employed in
this type of infrared spectroscopy use an interferometer originally designed by Michelson to
measure the speed of light as shown in Figure 4-1 below.

Figure 4-1
Typical Michelson Interferometer

The interferometer works by taking the entire incident beam of radiation from the source and
dividing it into two paths with a beam splitter. The beam splitter is usually some non-absorbing
film whose transmittance and reflectance are both approximately 50%. The beam splitter must
also permit the transmission of the wavelengths (nominally 2-10 microns) employed for this type
of spectroscopic measurement. Usually they are made of germanium. Therefore the source
radiation incident on the beamsplitter is divided into two parts. One of the paths goes to a fixed
mirror while the other path goes to a moving (translating) mirror. When the position of the
translating mirror is continuously varied along an axis collinear to the source, an interference
pattern I(x) is generated as the two phase shifted beams interfere with each other. By smoothly
translating one mirror, the optical path difference (OPD) is x=2L (where x is twice the distance L
traveled by the translating mirror). This optical path difference (also called retardation,) between
the beams traveling to the fixed and the moving mirrors will be the same for all wavelengths of
light. This leads to two boundary conditions. First, when the two beams have traveled the same
distance, they will be in phase and thus when they recombine at the beam splitter, will interfere
constructively. Second, when the movable mirror is at a distance twice that of the fixed mirror,
the two beams will be out of phase with each other and interfere destructively. By moving the
translating mirror at a constant velocity, the signal at the detector will vary sinusoidally. The
time varying component is the only component that is important in spectroscopic measurements
and is called the interferogram (Figure 4-2). The actual signal measured at the detector will
depend on the beamsplitter, detector response at different wavelengths and the emission light
source.

4-2
Possible Continuous Measurement Technologies for Measuring SO3 and H2SO4

Figure 4-2
Example Interferogram of Ambient Air Containing High Level of CH4 (Courtesy of IMACC)

The specific intensity Ik(x) can be derived for the source energy at a single wave number k by

Equation 4-2

where J(k) is the incident intensity and <T(k)> is the averaged beam splitter transmission
function. Since cos(kx) is an even function, the interferogram will be symmetrical about the
white light fringe. Since the resolution of a fourier transform spectrometer increases with
increasing optical path difference, the maximum spectral resolution is achieved by using the
entire available translation distance to measure only one side of the interferogram.

However, in order to maximize the signal-to-noise ratio (by avoiding the slight overhead
incurred in switching the direction of stage motion), both sides of the interferogram can be
measured, yielding a so-called "two-sided" interferogram. Two-sided interferograms contain
two measurements of each interferogram point per scan, but can only achieve half the optical
path difference (and therefore half the spectral resolution) of one-sided scans. Because one-sided
interferograms transform to real spectra, no explicit information on the interferogram phase is

4-3
Possible Continuous Measurement Technologies for Measuring SO3 and H2SO4

available, although phase problems do show up as anomalous spectral baselines. Two-sided


interferograms transform to complex spectra (they have two pieces of information per
frequency), allowing phase errors to be directly measured as a function of frequency. Two-sided
scans are therefore extremely useful for examining alignment of the optics and other potential
instrumental problems. As already mentioned, a perfectly aligned instrument with no phase
errors will produce completely symmetric interferograms whose transform will have zero
imaginary part over all frequencies in the passband. The total intensity measured for a given
OPD x from radiation at all wave numbers is found by integrating which is equivalent to
-1
applying the inverse Fourier cosine transform Fc

Equation 4-3

The FTIR spectrometer generates the infrared spectrum of a given sample by calculating the ratio
of the signal obtained by scanning air (empty beam) to the signal obtained by scanning the
sample gas.

This process is schematically illustrated in Figures 4-3 through 4-5, for the FTIR analysis of
styrene, (taken from the Columbia University web site (www.columbia.edu/ccnmtl/draft/dbeeb/chem-
udl/spectrometer.html)). First an interferogram of the source (background) is scanned (Figure 4-3
left), and then transformed into a single beam spectrum (Figure 4-3 right) and stored in computer
memory.

4-4
Possible Continuous Measurement Technologies for Measuring SO3 and H2SO4

Figure 4-3
Interferogram of the Source Background (Left), and Resulting Transformed Single Beam
Spectrum (Right)

The sample, (in this case containing styrene), is then sampled by many possible ways, either
extractively to a sample cell, in-situ or open-air and the interferogram of the sample gas is
scanned (Figure 4-4 left), and then transformed into a single beam spectrum (Figure 4-4 right)
and stored in computer memory.

Figure 4-4
Interferogram of the Gas Sample Containing Styrene (Left), and Resulting Transformed
Single Beam Spectrum (Right)

The ratio between the two single-beam spectra, in computer memory, is calculated and the
"double beam" presentation with a flattened baseline is produced. The features present in the
background spectrum correspond to the emission profile of the source, the optical efficiency, or
detectivity of the detector, the absorption of atmospheric water, and gaseous CO2. The ratio
process compensates for these effects and they don't appear in the spectrum of the sample as
shown in Figure 4-5.

4-5
Possible Continuous Measurement Technologies for Measuring SO3 and H2SO4

Figure 4-5
Resulting IR Spectrum of Styrene

By comparing the resulting IR spectrum of styrene to stored calibration spectra of styrene the
concentration of styrene in this sample can be ascertained. It is important to note that calibration
spectra are temperature and pressure dependent and accurate IR measurements require stored
calibration spectra at the temperature and pressure of the sample gas.

To investigate whether this technique could provide accurate measurements of SO3 and H2SO4, a
literature search specific to infrared measurements of these two gases was made. The results are
included in the listed references section of this report. What was clear was that there had been
many IR measurements of SO3, using both FTIR and mid infrared tunable diode laser
spectroscopy. Tunable diode laser spectroscopy has been used extensively in coal fired NH3
measurements, (refer to Appendix A, which describes this technique and its use with the
EPRI/UCR phase discrimination probe), but employ Near-IR lasers as the radiation source. The
mid infrared systems needed for SO3 measurements require cryogenic laser cooling and reduced
pressure sampling, which are very expensive, delicate, and far too complicated to be practical for
measurements in a coal-fired boiler environment.

IR spectra for SO3 and isomers of SO3 have been worked on by many researchers as Jones et al.,
1969, Kaldor et al., 1973, Ivanov et al., 1987, Ortigoso et al., 1989, Khanna et al., 1995, Jou et
al., 1996 and Maki et al., 2001. Earlier work (Jones et al., and Kaldor et al.) found from their
investigations that for SO3, peak absorbance occurs around the 1350-1420 cm-1 and 2400-2470
cm-1. This region has been used as the defined characteristic absorption region for IR SO3
measurements. Later work has been looking more at the asymmetric SO3- stretch at 1160-1190
-1 - -1
cm and the symmetric SO3 stretch between 1015-1080 cm . Research has been to identify the
vibrational or rotational absorption feature and estimate molecular constants for that specific
absorption.

IR spectra for H2SO4 have also been worked on by researchers as Milne et al., 1967, Jou et al.,
1996, Givan et al., 1998, They found from their investigations that for H2SO4, peak absorbance
-1 -1
occurs around the 800-950 cm region and the 1100-1280 cm region. This region has been used
as the defined characteristic absorption region for IR H2SO4 measurements. Again research has
been primarily to identify the characteristics of the vibrational or rotational absorption feature
and estimate molecular constants for that specific absorption.

4-6
Possible Continuous Measurement Technologies for Measuring SO3 and H2SO4

Three references showed measurements of SO3 using FTIR technology in harsh conditions,
similar to those expected at a coal-fired power plant. The first one was entitled “Monitoring of
Mining by Infrared Methods” by Onishchenko, A. M et al, and utilized FTIR for the
measurement of SO3 in ambient air in mines. The second was entitled, “Sulfur Trioxide
Generator and Sulfur Trioxide Quantitative Analysis Method Using Fourier-Transform IR
Spectrometer” by T. Satake et. al., measured SO3 extractively out of a utility boiler using FTIR.
Finally, a third reference came from personal correspondence with Dr. Bob Spellicy of IMACC
(Industrial Monitor and Controls Corporation of Austin, Texas). IMACC had performed a
measurement program with a coal-fired power plant in the United States and provided us with a
copy of an internal report of the measurements taken at this facility (Appendix B).

IMACC’s report presented results of two tests to demonstrate H2SO4 and SO3 monitoring with
FTIR. The first set of tests were laboratory tests conducted at IMACC’s facilities in Austin,
Texas. These were controlled tests to demonstrate detection, evaluate detection limits and to test
compatibility of material. The highlights of the lab experiments are shown in Figures 4-6, 4-7
and 4-8, which depict the H2SO4 and SO3 IR spectrums, for regions that experience little SO2 or
H2O interferences. The determination of the best spectral region for FTIR analysis is one of the
critical requirements in utilizing this technology.

For their laboratory setup, IMACC used a propylene glycol temperature-controlled bath to hold a
flask filled with H2SO4 at constant temperature. They added glass beads to the bath to aid in
diffusion of the air flow through the acid in order to achieve a better equilibrium. A pump was
placed at the output of the FTIR cell and was used to pull ambient air through the H2SO4 flask,
through a 50 ft. heat traced line and through the FTIR cell.

All lines were heated to 300ºF (150ºC) while the cell was held at 365ºF (185ºC) to avoid
condensation of the H2SO4. The flask was filled with 96% H2SO4 and it was heated to 175ºF
(80°C). At this temperature, the equilibrium vapor pressure of pure H2SO4 in water should be
41.4 ppm and that of SO3 should be 0.13 ppm.

To confirm that the FTIR was capable of seeing H2SO4/SO3 and not some contaminant, collected
spectra were viewed directly to identify the spectral signatures. In Figure 4-9 the portion of the
-1
observed spectrum from 2400 to 2530 cm is shown on top and this can be compared to the
library reference spectra for both SO3 in the middle and SO2 on the bottom.

4-7
Possible Continuous Measurement Technologies for Measuring SO3 and H2SO4

**Fri Sep 27 10:44:43 2002


0.052

0.050
Abs

0.048

0.046

*GAS5.040 SO3 7897.3 ppm 0.2m/760T/25C 8 mbar (not broadened p=6.08 Torr ) 20 cm 25 C 2 cm-1 => 0.5 cm-1

0.03
Abs

0.02

0.01

-0.00
SO2 800 ppm 5.2m/650t/185C Synthetic
0.03

0.02
Abs

0.01

-0.00
2520 2500 2480 2460 2440 2420
Wavenumbers (cm-1)

Figure 4-6
Plot of Experimental Data Taken by IMACC at Their Laboratory Facility Investigating SO3 IR
Spectra Compared to the SO3 Reference (middle) and the SO2 Reference (bottom). (By
Permission of IMACC)

It is clear that both SO3 and SO2 are present. Figures 4-7 and 4-8 show two regions where H2SO4
absorbs. Figure 4-7 is the 1100 cm-1 to 1260 cm-1 region and Figure 4-8 the 800 to 940 cm-1
region. In both of these, the sample spectrum is on top and the reference spectrum for H2SO4 is
on the bottom. Except for a little water vapor in the 1200 cm-1 region, both measurements are
identical to the H2SO4 reference. Clearly H2SO4 was also observed.

Once they had conducted the lab tests, field tests were undertaken at a power plant in the United
States. These tests attempted to reproduce the laboratory results and to perform real-time duct
measurements at the plant. Since the methodology was extractive, when IMACC first conducted
the field experiments, their first tests were to investigate whether there would be losses of either
SO3 or H2SO4, through the sampling lines. As discussed earlier, because of the reactive nature of
both of these target gases, extractive sampling leads to concern whether the measured sample
will generate results indicative of what is in the gas stream in the duct or will be biased high
from the high background SO2 concentrations, or biased low through sample line losses. With
FTIR, since the regions investigated for SO3 or H2SO4 would not be perturbed by the high
concentrations of SO2, the major concern was sample line losses.

4-8
Possible Continuous Measurement Technologies for Measuring SO3 and H2SO4

0.40 **Fri Sep 27 10:44:43 2002

0.35

0.30

0.25
Abs

0.20

0.15

0.10

0.8 817ppm H2SO4 22.1m/650/185C ,10 ppm SO3 (SO3 = 5.0ppm cal against 7897 ppm SO3) Checks w/ Applied Optics vol17 No.7 p976.

0.7

0.6

0.5
Abs

0.4

0.3

0.2

0.1

1260 1240 1220 1200 1180 1160 1140 1120 1100 1080
Wavenumbers (cm-1)

Figure 4-7
Plot Of Experimental Data Taken by IMACC at Their Laboratory Facility Investigating H2SO4
IR Spectra Compared to the H2SO4 Reference (Bottom) in the 1100 to 1260 Cm-1 Region. (By
Permission of IMACC)

IMACC’s tests showed that there were inherent problems in measuring the sample extractively.
First they experienced losses in the stainless steel couplings first used, and they were replaced
with all teflon material couplings. IMACC also concluded stainless steel seems to show
significant losses for H2SO4 and SO3 particularly below 250ºF – 300ºF (120ºC - 150ºC), and
should be avoided throughout the sampling system. IMACC also recommended that a probe
utilizing only teflon is likely necessary for low loss transfer, although this would limit available
sampling locations in a coal-fired boiler flue gas stream to temperatures compatible for teflon.
They also found that by sampling extractively and with the high moisture levels that most of the
SO3 will disappear, and be converted to H2SO4. They also observed an interesting result in that
the stopping of the sample flow resulted in a slow decrease in H2SO4, and to a lesser degree in
SO3. This is indicative of adsorption of the acid on the walls of the cell reducing its
concentration with time.

This type of effect is seen with many polar compounds and is not unexpected. It is for this
reason that continuous flow is required to properly sample these gases and that there will be a
period required for conditioning the walls of the sample lines and sample cell.

To minimize line losses, IMACC preconditioned the sample lines through prolonged exposure to
H2SO4. The problem with this approach is two-fold. The results may still be biased due to the
uncertainty of further absorption of the H2SO4 on the sample line walls, or possibly even some
desorptive effects from the preconditioning. Also since there is a time element with any dynamic

4-9
Possible Continuous Measurement Technologies for Measuring SO3 and H2SO4

sample as found in flue gas, it is unclear as how to be certain that the sample gas at the time
sampled is indicative of the actual sample of the flue gas.

However, problems that arose from IMACC’s initial measurements at a coal-fired power plant
were mostly from the extractive sample technique and not from the performance of the FTIR.
Their measured results showed levels of H2SO4 from 30 to 60 ppmV with minimum detection
limits of approximately 2 ppmV with a 10 meter path length.

The data from a single day of measurements is presented in Figure 4-9. Note that for this test,
sampling started at 18:00, and that the sample appears to reach equilibrium at approximately
18:45. There appears to be a drop in concentration at 10:00 pm, which was the time that IMACC
switched the probe material from Nylon to Teflon. Also during this time the flow was stopped
and ambient air was leaked into the sample cell. After switching to Teflon, the concentrations
did not change significantly. Initially, the sulfuric acid concentration rose significantly but then
fell back down appearing to equilibrate at about the same level as before the switch. Clearly this
shows that the extractive technique, creates its own set of problems and that if the FTIR could be
configured to make in-situ measurements that it could provide credible results with good sensitivity.

This leads to the question of the possibility of interfacing the FTIR with an optical system that
would allow in-situ monitoring, and thereby remove the uncertainties that arise with extractive
sampling complications.

It is highly possible to develop a sending and receiving optical assembly in conjunction with the
EPRI/UCR developed probe, which would interface with an FTIR. This would require transfer
optics from the FTIR spectrometer to the probe and collection optics to direct the return beam to
the detector. Since the probe attenuates the dust loadings by 70% or more depending on flue gas
flow rates, optical pathlengths of 20 meters or more are possible. This would yield minimum
detection limits of better than 0.3 ppmV for H2SO4 and 1 ppmV for SO3.

The advantage would not only be limited to the measurement of these two gases, as IMACC
demonstrated the capability of their FTIR system to not only detect and quantify concentrations
of H2SO4 and SO3 but also monitor other stack gases of interest in power plant applications, such
as H2O, CO, CO2, NH3, NO, NO2, SO2, and HCl.

The success of their initial tests indicate that further work should be undertaken specifically if
the problems with extractive sampling could be eliminated.

4-10
Possible Continuous Measurement Technologies for Measuring SO3 and H2SO4

**Fri Sep 27 10:44:43 2002

0.30

0.25
Abs

0.20

0.15

0.10

0.05
817ppm H2SO4 22.1m/650/185C ,10 ppm SO3 (SO3 = 5.0ppm cal against 7897 ppm SO3) Checks w/ Applied Optics vol17 No.7 p97
0.7

0.6

0.5
Abs

0.4

0.3

0.2

0.1

940 920 900 880 860 840 820


Wavenumbers (cm-1)

Figure 4-8
Plot of Experimental Data Taken by IMACC at Their Laboratory Facility Investigating H2SO4
IR Spectra Compared to the H2SO4 Reference (Bottom) in the 800 to 940 Cm-1 Region.
(By Permission of IMACC)

In their existing system collimated light is sent across the duct/stack where it hits a retro-
reflecting mirror on the far side. The light returning from the retro mirror re-enters the
spectrometer and eventually bounces off the beam splitter, and is subsequently focused onto the
detector. In this way, the light double passes the duct providing a path length twice the
duct/stack dimension. This is highly viable if the light passes through the EPRI/UCR probe in
coal-fired applications as the dust attenuation provided by the probe permits this type of
monostatic configuration .

In summary looking at the spectra taken for H2SO4 and SO3 utilizing the FTIR technique and
previous work done using this technique specifically for coal-fired applications the following
steps would be required to provide continuous measurements of SO3 in a coal-fired flue gas
environment:
1. The extractive sampling system was shown to create sampling issues, and ultimately is not
practical for the measurement of H2SO4 and SO3 from power plant flue gas streams.
IMACC has developed a cross-stack FTIR system to monitor industrial stacks and ducts in-
situ. As has been discussed, cross-stack monitoring has the advantage that the sample is
never perturbed so there is no question of a sample transport system modifying the sample
composition.

4-11
Possible Continuous Measurement Technologies for Measuring SO3 and H2SO4

70

60
3 Sigma Error
50

40
H2SO4 (ppm)

30

20

10

0
18:00 18:15 18:30 18:45 19:00 19:15 19:30 19:45 20:00 20:15 20:30
-10
Time

Figure 4-9
Plot of Experimental Data Taken by IMACC at a Coal Fired Power Plant in the Unit Duct.
The Error Bars are the 3 Sigma Errors Returned by the FTIR (With Permission of IMACC)

2. IMACC’s commercial instrument is designed to operate unattended for months at a time,


thereby meeting the criteria of a continuous type monitor for SO3 and H2SO4.
3. The key issue is that the dust loadings in power plants would make FTIR cross duct
measurements difficult as the signal is attenuated significantly. Any optical signal will see a
decrease in the signal to noise if the optical power of the light source is attenuated. When
the light source is attenuated to levels in the measuring environment where it is approaching
being optically thick, the signal to noise is far to low for measurements. This was the
problem for the approach taken by EPRI/UCR in the design of the phase discrimination
probe. The probe reduces the dust loadings in the optical path by 70% or more depending on
duct flow.
4. The cross-stack FTIR system would be enhanced if designed for in-situ measurements
utilizing the EPRI/UCR phase discrimination probe in a monostatic configuration. This is
highly possible as the probe has already been interfaced with an IR source in a similar
configuration.
5. Characteristic IR spectra for H2SO4 and SO3 and the region where they can be monitored with
the least amount of interference, have already been developed, but these are for sample gas
temperatures of 300ºF (150ºC). Since the IR spectra is temperature dependent, the IR spectra
at typical flue gas temperature for H2SO4 and SO3, as well as any other gas of interest, would
need to be generated and incorporated in the FTIR’s analytical library.

4-12
Possible Continuous Measurement Technologies for Measuring SO3 and H2SO4

4.3 DOAS (Differential Optical Absorption Spectroscopy)

Differential Optical Absorption Spectrometry (DOAS) is a UV/VIS optical measurement


technique based on Beer-Lambert’s Law. Since absorption coefficients in the UV are generally
orders of magnitude higher than in the IR region, sensitivity is usually enhanced, which makes
DOAS a method for measuring relatively low concentrations. On the other hand, UV absorption
features are much broader than those typically associated with infrared spectra, which leads to
two potential problems.

First, only species with absorption features that are considerably narrower than the typical
measuring window of 15–40 nm can be detected and quantified. This constraint significantly
reduces the number of compounds that can be measured.

Second, the light intensity without absorbing species in the optical path, I 0 , has to be known to
apply Beer-Lambert’s Law.

Broad absorption features, however, are affected by numerous species in the light path, and with
light scattering are often indistinguishable from the background and thus make the measurement
of I 0 difficult.

One approach to address these issues incorporates differential optical absorption, in which I 0 is
replaced by I 0' , where I 0' is the light intensity in the absence of a structured absorption band
(refer to Figure 4-10). The DOAS technique solves for the concentration using a modified form
of the Beer-Lambert’s Law in which a differential absorption coefficient ε ' replaces the absolute
absorption coefficient σ(ψ). C is the species concentration, L the optical path length, and I the
intensity at an absorption peak.

1 I'
C= ln 0 Equation 4-4
ε '⋅L I

The University of California at Riverside (UCR) has developed a DOAS system for the very low
level measurement (ppbV) of SO2 (Pisano et al, 2003). SO3 formed from the further oxidization
of SO2 likely has strong absorption lines in a similar region as SO2. Therefore, what was
required to assess the applicability of a DOAS measurement approach for SO3, was a UV
spectrum for SO3 from which estimations of the suitability of the present UCR DOAS system for
making SO3 measurements at a coal-fired plant could be made.

4-13
Possible Continuous Measurement Technologies for Measuring SO3 and H2SO4

"true" I0

broad absorption band

I'0

"differential"

Wavelength

Figure 4-10
'
DOAS Technique , in Which I 0 is Replaced by I 0 , the Light Intensity in the Absence of a
Structured Absorption Band.

DOAS System Description

The UCR differential optical absorption spectrometer (DOAS), was specially designed and built
for rapid, extractive detection of sulfur dioxide at low ppb levels. The integrated system consists
of the extractive sampling train, the White cell (White 1942), and the spectrometer. An overall
schematic layout of the system is shown in Figure 4-11 and explained in the following paragraphs.

For reactive molecules, such as SO2 or SO3, as well as to allow fast response times, the residence
time in the sample cell and thus the volume of the cell was made as small as possible. To
provide both a small cell volume and a long path length, the design uses a folded optical path in
the form of a White cell. Since exhaust gases are very corrosive, as well as hot and humid, and
contain a high numbers of particles, direct exposure of the mirrors to the sample gas was
avoided. Therefore, the White cell was designed as an open cell with the sample cell residing
between the mirrors, rather than having the mirrors inside the sample cell, as is the usual
configuration. Different from the aluminum coated spherical mirrors, the quartz windows are
easy and fast to remove and clean, as well as relatively inexpensive to replace. An extractive
system similar to this would be practical for coal-fired flue gas applications.

4-14
Possible Continuous Measurement Technologies for Measuring SO3 and H2SO4

Figure 4-11
Components of the UCR Developed DOAS System Employed for the Measurement of Low
Level (30-100 ppbV) SO2 From Vehicle Exhaust

The complete sample train extracts the engine exhaust gas from a sample port added to the
vehicle’s exhaust system between the engine and the catalyst. A pump sucks the exhaust gas
from the insulated sample port through a heated particle filter (300ºF ; 150ºC) and a heated 5 m
long PFA-line (300ºF ; 150ºC) into the heated sample cell (250ºF ; 120ºC) for the UV
smeasurements. Care was taken to avoid any cold spots in this sampling path.

The gas pressure inside the sample cell was slightly below atmospheric due to the extractive
sampling technique. All interior surfaces of the sample train are either stainless steel or PFA, the
only exception being the particle filter, which consists of bonded quartz fibers. The 1.4 m long
rectangular sample cell is equipped with fused silica windows at each end to allow the UV light
to enter and exit, and rests between the concave mirrors of the open White cell. The total volume
of the DOAS sample cell is approximately 3.19 liters (l), which gives a residence time of
15 seconds for a sample flow of 10 l/min (normalized to room temperature) and a sample
temperature of 230ºF (110ºC). The heating system is a set of flat Kapton-insulated etched-foil
heaters bonded to the stainless-steel tube with a high temperature silicone adhesive.

The placement and the power loading of the individual heating elements have been balanced to
allow an even temperature distribution along the sample cell. A tunable PID controller maintains
the cell temperature at a user-defined level.

4-15
Possible Continuous Measurement Technologies for Measuring SO3 and H2SO4

To avoid possible interferences from other exhaust constituents, the spectral range was restricted
to the UVB region (290–330 nanometer (nm)). As was subsequently determined, this region
would be unsuitable for SO3 measurements and a new grating would be required.

For the 290–330 nanometer (nm) region the use of inexpensive and readily available low power
(20 Watt) halogen bulbs as a light source was feasible. The spectral output of this light source
was tested and shown to drop off significantly below 270 nm. This characteristic would make
the present light source unusable for SO3 measurements.

Looking at the present UCR system, the light is focused onto the first mirror of the open White
cell. Light exiting the White cell passes through an optical filter to remove the visible
wavelengths, and is then focused onto the round tip of an optical fiber bundle. At the
spectrometer end, the individual strands in the fiber bundle are oriented in a line, thus acting as
the entrance slit for the spectrometer.

The spectrometer consists of a single working element, a concave, 1200 groove/millimeter (mm)
holographic diffraction grating optimized for 300–400 nm. The concave diffraction grating both
wavelength-disperses the light and focuses it onto the single-stage thermoelectric cooled charge-
coupled device (CCD) detector array. The effective resolution of the spectrometer is 0.2 –0.3 nm.
Spectra are integrated over 3 seconds and analyzed immediately.

The White cell is aligned for 10 or 14 passes, resulting in 14 m or 19.6 m optical path length,
respectively.

To investigate whether the UCR DOAS could be used to measure SO3, a UV absorption spectra
was measured. A sample gas containing approximately 100 ppmV of SO3 was generated and a
diode array spectrometer with a 5 meter path absorption cell was used for the spectral
measurements over the wavelengths from 220 nm up to 300 nm. The absorbance spectra is
shown in Figure 4-12. The absorbance is shown in absorbance units, for the relative quantity of
SO3 used.

4-16
Possible Continuous Measurement Technologies for Measuring SO3 and H2SO4

0.0003

0.0002

0.0001

0
Absorbance

220 230 240 250 260 270 280 290

-0.0001

-0.0002

-0.0003

-0.0004
Wavelength (nm)

Figure 4-12
The Absorbance Spectra Over the Wavelengths From 220 nm Up to 300 nm for SO3, for 100
ppmV of SO3, a Five-Meter Sample Cell and a Diode Array Spectrometer.

The major UV absorbance for SO3 lies in the region from 230 to 245 nm, a region where there
would be little overlapping from SO2 absorbance. The absorbance is relatively weak, however,
and would require a much longer path White Cell than that in the UCR system to achieve similar
detection limits as that for SO2. Since the detection limits in the power plant applications are on
the order of 1 ppmV, and not 30 ppbV as achieved for SO2 at a pathlength of 19.6 meters, a
pathlength of 10 meters should be sufficient for 1 ppmV detection limits of SO3. Therefore the
present UCR system could not be used for the measurement of SO3 and would require
modification.

In summary looking at the absorbance of SO3 compared to that of SO2, and utilizing the DOAS
technique with a system comparable to that developed at UCR, the following steps would be
required to confirm the DOAS capability for continuous SO3 measurements.
1. The peak absorbance of SO3 is in a different region than that of SO2. This has a benefit in that
we would likely receive little interference from the high levels of SO2 in the flue gas, usually
100 times that of SO3. This creates the need for both a new grating and a new light source for
the current DOAS system. The new light source would likely be a tungsten mercury lamp,
and the grating would be optimized for the region 230-245 nanometers.

4-17
Possible Continuous Measurement Technologies for Measuring SO3 and H2SO4

2. The extractive sampling system was for automotive exhaust and heated to 300ºF (150ºC).
This would not be practical for the measurement of SO3 extractively in power plant flue
gases. The sampling system would have to be designed to be able to maintain higher
temperatures so as to maintain the equilibrium between SO3 and H2SO4.
3. The sample cell placed in between the White cell was heated to temperatures up to 250ºF
(120ºC), this would also need to be modified to reach temperatures that would maintain the
equilibrium between SO3 and H2SO4.
4. It may be possible to develop an optical system that would allow in-situ monitoring,
obviating the need for extractive sampling complications. A sending and receiving optical
assembly in conjunction with the EPRI/UCR phase discrimination probe would be required
(Appendix A). This approach for a DOAS system would parallel commercially developed
systems by both TECO of the USA and OPSIS of Sweden which utilize a similar type of
optical configuration, but for the primary pollutants (NOX, and ozone (O3)).
5. The DOAS technique would only measure SO3 and not H2SO4, as the feasibility of doing
H2SO4 by DOAS would require taking UV spectra of H2SO4 and determining if there would
be a region where both species could both be measured concurrently.

4-18
5
CONCLUSION

SO3 will always be present to a limited degree in a coal-fired boiler, as it is unavoidable


whenever fossil fuels containing sulfur are used. SO3 emissions are quite variable and depend on
many parameters. Specifically, SO3 emissions are primarily influenced by the concentration of
sulfur in the fuel, the flue gas temperature, and the level of excess air, as well as the fly ash
composition and size. For boilers firing sulfur bearing fuels, SO3 is typically present at levels
between 1-25 ppmV. Most recently, however, the broad implementation of SCR technology for
NOx control has resulted in the further oxidation of SO2 in the flue gas to SO3 through contact
with the SCR catalyst. On units firing a high sulfur coal, this has resulted in SO3 emissions
exceeding 50 ppmV in some cases.

The formation of SO3 has been shown to have adverse effects on plant operations, primarily due
to the formation of H2SO4 which can lead to the corrosion of air heater and/or ductwork surfaces.
With the growing implementation of SCR and SNCR NOX control technologies, SO3 can also
react with residual ammonia slip, forming ammonia bisulfate, which can lead to potential fouling
of the air heater. Environmental considerations also include the formation of visible plumes as
well as being a precursor to the formation of particulate level PM2.5. Finally, power plants need
to estimate sulfuric acid emissions as it is required that US utilities report their annual releases of
toxic chemicals to the Toxic Release Inventory (TRI), a national compilation of chemical
releases designated toxic by the US EPA, of which sulfuric acid is one.

Current methods for estimating SO3 emissions do not take into consideration all pertinent site
specific factors and have limited accuracy. Current methods for measuring SO3 are based on
extractive techniques which can yield biased results due to the reactive nature of SO3. Additional
measurement biases can arise from the further oxidation of SO2 in the sample train. These
extractive techniques require adherence to proper procedures and have been shown to have a
significant degree of variability from tester to tester, leading to testing inaccuracies. The only
continuous SO3 monitor currently available is from Severn Science, which is also extractive.
The instrument has been shown to give credible results, but only with skilled operator use. The
extractive approach has been found to not be practical for long term power plant operation.
These extractive techniques only represent point measurements of the local SO3 in the flue gas
duct, requiring an assumption of homogeneous distribution that is often not representative of
most boilers.
In sum, current measurement methodologies not only have accuracy limitations, but also only
provide a time averaged sample that is not spatially resolved. As such, the physical
measurement may not be representative of the overall SO3 concentration in the flue gas stream.
Clearly what is needed to give the coal-fired boiler operators better overall SO3 emission control,
as well as to help evaluate the SO3 emission models, is a reliable, easy to operate, continuous SO3
monitor. The monitor would provide better accuracy if it could provide in-situ measurements,
obviating the need for sample extraction and the problems associated with that type of sampling.
The availability of such a continuous SO3 monitor would provide significant cost savings by

5-1
Conclusion

enabling the minimization of additive feed rates used to control SO3 emissions. Also by
providing an accurate real time measurement of SO3, operators would be able to detect an
increase in SO3 levels resulting from a change in fuel specifications, combustion considerations,
SCR operation etc., so that appropriate corrective actions could be taken. A continuous SO3
monitor would also help plants that burn a low-sulfur coal to minimize the amount of SO3
injected to condition fly ash for high ESP removal efficiencies.
For in-situ measurements, optical spectroscopic techniques have been shown to be reliable.
Most boilers have opacity meters as part of their continuous emission monitoring equipment.
Tunable diode laser spectroscopy has been incorporated in some coal-fired power plants for the
measurement of NH3 slip. SO3 and H2SO4 have been investigated extensively spectroscopically
in the infrared region where they both can be monitored concurrently with the least amount of
interference. Both gases can be monitored via Fourier Transform Infrared Spectroscopy (FTIR)
to the detection limits required for coal-fired applications. Many commercial companies sell
FTIR analyzers. IMACC’s commercial instrument is designed to operate unattended for months
at a time, thereby meeting the criteria of a continuous type monitor for SO3 and H2SO4 with low
maintenance. With expected dust loadings in power plant ducts of nominally 5 grains per cubic
foot, in situ FTIR measurements would likely not work without being interfaced to some dust
attenuating probe. EPRI/UCR have successfully developed such a probe and have already
interfaced this with a laser IR source. Therefore, this monitoring configuration with the FTIR
and the phase discrimination probe is highly possible, and would likely yield concurrent in-situ
real time measurements of SO3 and H2SO4.
Differential Optical Absorption Spectroscopy (DOAS) was also investigated for the continuous
measurement of SO3. The DOAS technique has been used to monitor SO2 under laboratory
conditions at UCR (University of California at Riverside). The goal was to see if the instrument
developed at UCR could be used to make SO3 measurements, and if not, what modifications
would be required. There was very little in the published literature with respect to SO3 UV
spectra, and tests were conducted at UCR to measure the UV spectra of SO3. It was found that
peak SO3 UV absorbance is in the 230-245 nanometer region, a different region than that of SO2.
This indicates that the present UCR DOAS system would need both a new grating and a new
light source to enable the measurement of SO3. The UCR system is also extractive and would
suffer the same problems encountered by all extractive techniques. It may be possible to develop
an optical system that would allow in-situ monitoring, as discussed with the FTIR, but the DOAS
technique presently could only measure SO3 and not H2SO4, as the feasibility of doing H2SO4 by
DOAS would require taking UV spectra of H2SO4 and determining if there would be a region
where both species could both be measured concurrently.
Finally an extensive literature search was undertaken, with regards to three topics. One was
spectroscopic information for SO3 and H2SO4, second was current measurement techniques for
SO3 and H2SO4, and lastly issues associated with SO3 emissions from coal-fired boilers.

5-2
6
REFERENCE LIST FROM LITERATURE SEARCH

One of the components of this report was to undertake an exhaustive literature search on SO3
related publications. Publications were searched with respect to either spectroscopic information
relating to identification of SO3 or H2SO4 spectra, measurement techniques for SO3 and H2SO4
(specifically gas phase), or other SO3 related issues for coal fired power plants.

The literature search results listed were based primarily from two sources. The first set of results
were from searching the database of Melvyl, The catalog of the University of California
Libraries, the second set of results were from using the search engine Google to look at web
based references.

Once the reference was found it was delineated into one of three sections, these being:
1. Spectroscopic Reference
2. Measurement Technique Reference
3. SO3 Coal-fired power plant Related References

Once the reference was deemed appropriate, a copy of the article was secured. The list of
references presented here is in no way complete and only presents the results from these
searches. The references for each section were listed in order alphabetically with respect to the
primary author. For example the reference shown below, lists the title of the paper in bold
typeface, in this example: High-Resolution Infrared Spectra of the ν2, ν3, ν4, and 2ν3 Bands
of 32S16O3, the primary author Maki, Arthur is italicized and in bold typeface, the secondary
authors and reference information are in normal typeface.

High-Resolution Infrared Spectra of the ν2, ν3, ν4, and 2ν3 Bands of 32S16O3.
Maki, Arthur; Blake, Thomas A.; Sams, Robert L.; Vulpanovici, Nicolae; Barber, Jeffrey;
Chrysostom, Engelene T. H.; Masiello, Tony; Nibler, Joseph W.; Weber, Alfons. Mill Creek,
WA, USA. Journal of Molecular Spectroscopy (2001)

The spectroscopic references include the abstract to better describe the research work done here
in this report, since the major goal was to look at alternative technologies, more specifically,
spectroscopic types, for the measurement of SO3 and H2SO4.

Most of the work listed in the 28 spectroscopic references, deal more with identifying regions
that can be used for successful IR measurements than actual measurements or instrumentation.
As can be observed from the list of references very little work has been done on the generating of
the UV spectra, which is why we did the measurement ourselves.

6-1
Reference List From Literature Search

Abstracts for the other articles are not included, only the reference is listed as the topics have
been included here only for background information, but there are far better reports specific to
the many concerns with respect to SO3 briefly discussed in this report.

6.1 Spectroscopic References


1. IR spectrum of H2SO4 species trapped in low temperature solid CO and in CO
containing matrixes. Givan, Aharon; Larsen, Lars A.; Loewenschuss, Aharon; Nielsen,
Claus J. Department of Inorganic and Analytical Chemistry, The Hebrew University of
Jerusalem, Israel. Journal of the Chemical Society, Faraday Transactions (1998), 94(16),
2277-2286.

Abstract
The IR spectrum of the vapors above neat H2SO4 trapped in CO enriched Ar matrixes was
studied and compared to their IR spectra in pure CO matrixes. The authors assigned
vibrational modes of (OC)⋅(H2SO4), (OC)⋅(H2O)⋅(H2SO4), (OC)2⋅(H2SO4) and (OC)⋅(SO3)
species isolated in Ar matrixes in all relevant spectral regions. Bonding between the moieties
of the mixed complexes is indicated. These species are formed by surface diffusion within
the deposited layer and do not represent a vapor phase equil. No spectral evidence was found
for any binary (H2O)m.cntdot.(SO3)n or ternary (CO)l⋅(H2O)m.cntdot.(SO3)n complexes or of
ionic species, to indicate a proton transfer <30 K between the H2SO4 and H2O moieties in
either CO enriched or pure CO matrixes.
2. The ν3 band of sulfur trioxide at high resolution. Henfrey, N. F.; Thrush, B. A. Dep.
Phys. Chem., Univ. Cambridge, Cambridge, UK. Chem. Phys. Lett. (1983), 102
(2-3), 135-8.

Abstract
The origin of the ν3 band in SO3 was reassigned to 1391.5205(15) cm-1 on the basis of a
Fourier transform spectrum at 0.01 cm-1 resoln. Improved mol. consts. were obtained,
particularly for the dependence of the Coriolis interaction on vibrational excitation.
3. Absorption of IR radiation at weakly forbidden transitions in triatomic molecules.
Ivanov, S. V.; Panchenko, V. Ya. Mosk. Gos. Univ., Moscow, USSR. Kvantovaya
Elektron. (Moscow) (1987), 14(1), 210-13

Abstract
Spectra of the av. probability and d. of weakly forbidden vibration-rotational transitions with
|∆K|>1 were called. for ν3 and ν1 bands in O3 and SO2, resp. Model equidistant spectra were
employed to analyze the contribution of weakly forbidden transitions with ∆K = ±3 into the
excitation of ν1 vibrations in SO2. The ranges of the radiation intensity were detd. in which
the absorption at weakly forbidden transitions reaches the max.

6-2
Reference List From Literature Search

4. Measuring hazardous air pollutants. Jones, Linda E.; Ingram, Michael B. New York
State College of Ceramics, Alfred, NY, USA. Am. Ceram. Soc. Bull. (1995), 74(7), 66-7.

Abstract
The suitability of mass spectrometry with a capillary sampling system was studied for the
accurate detection of a broad range of emissions.
5. Vibrational frequency isotope shifts for sulfur trioxide. Jones, Llewellyn Hosford. Los
Alamos Sci. Lab., Los Alamos, N. Mex., USA. Acta Chem. Scand. (1969), 23(1), 335.

Abstract
Force consts. of 32S18O3, 34S16O3, and 34S18O3 were calcd. and isotopic shifts were estd. The
max. deviation in ν3* and ν4* should be ±0.4 and ±0.2 cm , resp., for S18O3 rather than ±5
.-1

and ±1 cm.-1, resp., as reported by R. Stolevik, et al., ibid. 1967, 21, 1581.
6. Isomers of SO3: infrared absorption of OSOO in solid argon. Jou, Shen-Horng; Shen,
Min-yi; Yu, Chin-hui; Lee, Yuan-Pern. Dep. Chem., Natl. Tsing Hua Univ., Taichung,
Peop. Rep. China. J. Chem. Phys. (1996), 104(15), 5745-53.

Abstract
S trioxide (SO3) isolated in solid Ar at 12 K was irradiated with light at 193 nm from an ArF
excimer laser. Recombination of photofragments O and SO2 produces OSOO that absorbs at
1229.6, 1041.3, and 597.6 cm-1. The assignments are based on obsd. 34S- and 18O-isotopic
shifts. Theor. calcns. using the B-P86 and the B3-LYP d. functional methods were carried
out for 5 isomers of SO3; energies and vibrational wave nos. were predicted for each. Obsd.
line positions, IR intensities, and isotopic shifts fit well with those predicted for cis-OSOO at
the B-P86 level. Further irradn. of the matrix sample with emission at 248 nm from a KrF
laser bleached OSOO and enhanced absorption lines of SO2 and, to a lesser extent, of SO3.
The mechanism of formation of OSOO in a matrix cage is discussed.
7. Assignment of ν2 and ν4 of Sulfur trioxide. Kaldor, A.; Maki, A. G.; Dorney, A. J.;
Mills, I. M. Natl. Bur. Stand., Washington, D. C., USA. J. Mol. Spectrosc. (1973), 45(2),
247-52.

Abstract
Three expts. were performed to resolve an uncertainty in the assignmant of ν2 and ν4 for
SO3: (1) the gas phase Raman spectrum was measured; (2) the ir active combination band ν3
+ ν4 was measured; (3) a band contour calcn. was performed taking account of the l-type
resonance in ν4 and a strong Coriolis resonance between ν2 and ν4. These expts. establish
beyond any doubt that ν2 lies at .apprx.497.5 cm-1 and ν4 lies at about 530.2 cm-1. The
contour calcn. also shows that the Coriolis resonance gives rise to a pos. intensity
perturbation.

6-3
Reference List From Literature Search

8. High-resolution infrared spectrum and the molecular structure of sulfur trioxide.


Kaldor, Andrew; Maki, Arthur G. Opt. Phys. Div., Natl. Bur. Stand., Washington, D. C.,
USA. J. Mol. Struct. (1973), 15(1), 123-30

Abstract
-1
The 2ν32 (E') band of SO3 was measured with a resoln. of 0.03 cm . The anal. of this band
-1
yielded rotational consts. which fit the data to within a std. deviation of 0.003 cm . From the
rotational const. B0 = 0.34857 ± 0.00006 cm-1, the S-O bond distance (r0 = 1.4198 ± 0.0002
.ANG.) agreed with earlier ir measurements, but represented a considerable improvement in
accuracy.
9. Infrared spectra and structure of solid phases of sulfur trioxide: possible identification
of solid SO3 on Io's surface. Khanna, R. K.; Pearl, J. C.; Dahmani, R. Department of
Chemistry and Biochemistry, University of Maryland, College Park, MD, USA. Icarus
(1995), 115(2), 250-7.

Abstract
The IR spectra of solid sulfur trioxide (SO3) have been examd. between 20 and 290 K. The
spectral features of an amorphous low-temp. deposit and four addnl. phases (monomeric, γ,
α, and β) are presented. Two of the emission peaks in the thermal spectrum of Io recorded
by the Voyager IRIS instrument are suggested to be possibly due to solid SO3 mixed with
cryst. SO2.
10. Spectrophotometric determination of sulfur dioxide and sulfur trioxide in a gas
mixtures. Levina, L. M.; Strots, V. O.; Popov, S. A. Inst. Catal., Novosibirsk, USSR.
Zh. Anal. Khim. (1988), 43(4), 655-9.

Abstract
A simple, reliable and nondestructive method has been developed for direct detn. of SO2 and
SO3 in a gas phase. The method is based on measuring the absorbance of the gas mixt. in the
UV region at 230 and 288 nm. The relative error was ≤3.0% for detg. 1-12% SO2.
11. High-Resolution Infrared Spectra of the ν2, ν3, ν4, and 2ν3 Bands of 32S16O3. Maki,
Arthur; Blake, Thomas A.; Sams, Robert L.; Vulpanovici, Nicolae; Barber, Jeffrey;
Chrysostom, Engelene T. H.; Masiello, Tony; Nibler, Joseph W.; Weber, Alfons. Mill
Creek, WA, USA. Journal of Molecular Spectroscopy (2001)

Abstract
New measurements are reported for the IR spectrum of S trioxide, 32S16O3, with resolns.
ranging from 0.0015 cm-1 to 0.0025 cm-1. Rovibrational consts. were measured for the
fundamentals ν2, ν3, and ν4 and the overtone band 2ν3. Comparisons are made with the
earlier high-resoln. measurements on SO3, and the high correlation among some of the
consts. related to the Coriolis coupling of the ν2 and ν4 levels is discussed to understand the
areas of disagreement with the earlier work. Splittings of some of the levels are obsd. and
the splitting const. for K = 3 of the ground state is detd. for the 1st time. Other obsd.
splittings include the K = 1 levels of 2ν3 (l = 2), the K = 2 levels of ν3 and ν4, and the K = 3
levels of ν2. The anal. shows that there are level crossings between the l = 0 and l = 2 states
of 2ν3 that allow one to det. the sepn. of the subband centers for these two states even though

6-4
Reference List From Literature Search

access to the l = 0 state from the ground state is elec.-dipole forbidden. This is a generalized
phenomenon that should be found for many other mols. with the same symmetry. The l-type
resonance const., q3, that causes the splitting of the l3 = ±1, k = ±1 levels of ν3 also couples
the l3 = 0 and 2 states of 2ν3. (c) 2001 Academic Press.
12. Coriolis coupling constants in sulfur trioxide. Milne, J. B.; Ruoff, A. Tech. Hochsch.,
Stuttgart, Ger. J. Mol. Spectrosc. (1967), 23(4), 408-15

Abstract
The ir bands, ν3 and ν4, of gaseous SO3 have been studied. Twenty-one Q-heads could be
identified in ν3 and from their positions and the contour of ν3, ζ3 values have been detd.
The values from both methods show good agreement. ζ4 has been estd. from the intensity
ratio of the Q- and R-branch of ν4.
13. Chemical-resistant near infrared-cutting filter glass. Murakoso, Masatoshi. (Toshiba
Glass Co., Ltd., Japan). Jpn. Kokai Tokkyo Koho (1989), 3 pp.

Abstract
The glass, useful for color correction filters, etc. comprise P2O5 65-85, Al2O3 3-17, SiO2 0.1-5,
SnO2 0.1-3, SiO2 + SnO2 0.3-8, B2O3 0-10, Li2O, Na2O, and/or K2O 0-15, MgO, ZnO, and/or
CaO 1-15, BaO and/or SrO 0-10, ZrO2, TiO2, La2O3, and/or Y2O3 0-10, SO3 ≤1, and CuO 0.5-
8 wt.%. The glass showed high light transmittance at 400-600 nm, water resistance, and
chem. resistance.
14. Reduction of sulfur trioxide contained in boiler flue gases. Nojima, Shigeru; Iida, Kozo;
Naito, Osamu; Uchihashi, Kazumasa. (Mitsubishi Heavy Industries, Ltd., Japan). Jpn.
Kokai Tokkyo Koho (1998), 5 pp.

Abstract
SO3-contg. boiler flue gases are treated by redn. with NH3 in the presence of an Ir catalyst on
porous supports. The porous supports may be made from TiO2, Al2O3, SiO2, ZrO2, silicalites,
and/or metallosilicates.
15. Control of mining atmosphere by IR methods. Onishchenko, A. M.; Krichko, I. B.;
Ivashev, A. V. Inst. Gorn. Dela im. Skochinskog, USSR. Izv. Vyssh. Uchebn. Zaved., Gorn.
Zh. (1991), (8), 1-4.

Abstract
The use of IR methods for remote automatic monitoring of the concns. of gas components
(CH4, CO, CO2, SO3, etc.) in mining atm. and of the condition of the rock massif and for
prediction of abrupt releases of CH4 and remote control of the mining equipment is
discussed.

6-5
Reference List From Literature Search

16. Monitoring of mining by infrared methods. Onishchenko, A. M.; Krichko, I. B.;


Ivashev, A. V. IGD im. Skochinskogo, Moscow, USSR. Fiz.-Tekh. Probl. Razrab. Polezn.
Iskop. (1991), (1), 84-7

Abstract
The uses of IR methods in coal mining are discussed. They include, e.g., control of massif
condition; prediction of gas eruptions; control of mining operations and machines; detn. of
CH4, CO2, CO, SO3 in air; detn. of coal and rock dust in mine air; monitoring of the SiO2
content in air dust.
17. The ν2 and ν4 IR bands of sulfur trioxide. Ortigoso, Juan; Escribano, Rafael; Maki,
Arthur G. Inst. Estruct. Mater., CSIC, Madrid, Spain. J. Mol. Spectrosc. (1989), 138(2),
602-13.

Abstract
The IR spectrum of SO3 was recorded at 0.005 cm-1 resoln. at 460-567 cm-1. The >2000
lines were assigned to the Coriolis-interacting ν2 and ν4 fundamental bands. A rotational
anal. was performed which allowed refining rovibrational parameters of the ground state and
the v2 = 1 and v4 = 1 states. A discussion is presented on the different redns. of the
Hamiltonian employed. Improved values were obtained for equil. parameters of this mol.
18. A UV-Differential Optical Absorption Spectrometerfor the measurement of sulfur
dioxide emissions from vehicles. J. T. Pisano1, C. G. Sauer1, J. Robbins2, J. W. Miller1,
H. Gamble2 and T. D. Durbin1 (1)1 Bourns College of Engineering – Center for
Environmental Research and Technology (CE-CERT), University of California, Riverside,
CA 92521, (2) UNISEARCH Associates Inc., 96 Bradwick Drive, Concord, Ontario, Canada
L4K 1K8. J. Measurement and Science Technology. (2003), 14, 1-7

Abstract
A differential optical absorption spectrometer (DOAS) for measurements of low-level sulfur
dioxide emissions from vehicles in real time was developed and employed. With a time
resolution of 3 seconds and an optical path length of 19.6 m, a minimum detection limit of 75
ppbv sulfur dioxide (three times the standard deviation) was achieved. The wavelength
region covered, UVB, was chosen to include the major sulfur dioxide absorption feature at
300 nm, while avoiding wavelength regions in which other components of car exhaust are
known to absorb. However, formaldehyde was found to be present in vehicle exhaust in high
enough concentrations to be a major interferent. The analysis software was modified
accordingly. The application of the instrument to real vehicle exhaust demonstrated that it
was capable of measuring the contribution to sulfur dioxide emissions from lubricant oil
sulfur over the range of current and future oil sulfur levels.

6-6
Reference List From Literature Search

19. Determination of chemical species of selected trace elements in fly ash. Polyak, Klara;
Bodog, Ildiko; Hlavay, Jozsef. Dep. Analytical Chemistry, Univ. Veszprem, Veszprem,
Hung. Talanta (1994), 41(7), 1151-9.

Abstract
Complex anal. methods were developed to det. the chem. compn. of fly ash from coal-fired
power plants and municipal waste incinerators. Morphol. investigations and single particle
anal. were performed by SEM/EDAX method. A survey of mineralogical phases was made
by x-ray powder diffraction and IR spectrometry. Solvent-leaching expts. were conducted
for information on metal pollutant mobility under real environmental conditions. Cu, Ni, Co,
Cr, Pb, and Cd were studied, and of the toxic metals, Cd was found in exchangeable forms in
large amts. Mobile species of toxic metals may impact receiving water quality or soil organisms.
20. Sulfur trioxide generator and sulfur trioxide quantitative analysis method using
Fourier-transform IR spectrometer. Satake, Tsukasa; Yoshida, Makoto; Yamanuki,
Mikito. (Horiba, Ltd., Japan). Jpn. Kokai Tokkyo Koho (2001), 6 pp.

Abstract
The SO3 generator is used for generating SO3 by oxidn. of SO2; it includes an oxidn. furnace
for oxidn. of known concn. SO2 in an inert gas-based gas, using an FTIR spectrometer for
detecting SO2 concn. and SO3 concn. at the outlet of the oxidn. furnace, and controlling the
temp. condition of the oxidn. furnace according to the detected SO3 concn. signal. In the SO3
quant. anal. method is characterized by calcg. SO3 component gas concn. in gas sample
according to absorbance value in characteristic absorption regions of 1350-1420 cm-1 or
2400-2470 cm-1 of the absorption spectra. The app. and method can be used for measuring
SO2 oxidn.-formed SO3 concn. in boiler flue gas, etc.
21. Determination of oxides in work places by a new combustion technique. Sojka, Dariusz.
Huta Batory S.A., Chorzow, Pol. Hutn.--Wiad. Hutn. (1995), 62(10), 448-9.

Abstract
Two anal. instruments (Leco CS-244 and Leco TC-436) were used to det. sulfur trioxide and
iron oxide dusts in workplace air. The Leco CS-244 dets. C and S via an IR absorption
detection method via CO2 and SO2 released from the sample during combustion. The Leco
TS-436 dets. N and O; N by a thermal cond. TC cell which detects differences in the thermal
cond. of gases, whereas O is measured by IR detection of CO2.
22. Fourier transform infrared spectroscopy for process monitoring and control. Solomon,
Peter R.; Morrison, Philip W., Jr.; Serio, Michael A.; Carangelo, Robert M.; Markham,
James R.; Bates, Stephen C.; Cosgrove, Joseph E. Adv. Fuel Res., Inc., East Hartford, CT,
USA. Proc. SPIE-Int. Soc. Opt. Eng. (1993), 1717(Industrial, Municipal, and Medical
Waste Incineration Diagnostics and Control), 104-15.

Abstract
Tests of Fourier transform IR (FT-IR) spectrometric measurement of pollutant concns.,
temp., and particle size in a coal-fired power plant flue gas stream prior to the electrostatic
precipitator, in the smoke stack of a paper mill, and in a black liquor recovery boiler showed
that this technique has excellent potential for such applications. The hardware was reliable,
stable, and sensitive, and the in-situ gas species concns. were detd. at the few ppm level in

6-7
Reference List From Literature Search

spite of the presence of high concns. of water vapor. Particle opacity was also detd. FT-IR
emission and transmission tomog. was also done on a lab. C2H4 flame to det. species concns.
and temps.
23. Pollutant emission monitoring using QC laser-based mid-IR sensors. Sonnenfroh,
David M.; Allen, Mark G.; Rawlins, W. Terry; Gmachl, Claire G.; Capasso, Federico;
Hutchinson, Albert L.; Sivco, Deborah L.; Cho, Alfred Y. Physical Sciences Inc., Andover,
MA, USA. Proceedings of SPIE-The International Society for Optical Engineering (2001),
4199(Water, Ground, and Air Pollution Monitoring and Remediation), 86-97.

Abstract
Recent advances in mid-IR semiconductor laser technol. based on inter-subband transitions
in InGaAs quantum wells promise a dramatic impact on tunable diode laser-based sensors for
trace gases. Progress toward realizing room-temp., laser-based sensors for combustion-
generated pollutants such as NOx and SOx is discussed. Lab. measurements of SO2 at 8.6
.mu.m are presented with detection limits on the order of 1 ppm. Extensions of these
approaches for higher sensitivity measurements in exhaust gas conditions are described, as
well as SO3 measurements.
24. Laboratory and field measurements of H2SO4 and SO3 using FTIR spectroscopy
Spellicy, R. L. and Brewer, R. J. Industrial Monitor and Control Corporation 800 Paloma Dr.
Suite 100 Round Rock, TX 78664

Abstract
This internal report presents results of two tests to demonstrate H2SO4 and SO3 monitoring
with FTIR. The first set of tests were laboratory tests conducted at Imacc’s facilities in
Austin, Texas. These were controlled tests to demonstrate detection and to test compatibility
of material. Subsequently field tests were undertaken at a power plant in the US. These tests
attempted to reproduce the laboratory results and to perform real-time duct measurements at
the plant. The results of both tests indicate that FTIR can detect and quantitate
concentrations of H2SO4 and SO3.
25. Method and device for analysis of motor vehicle exhaust gases. Standt, Ulrich-Dieter.
(Volkswagenwerk A.-G., Germany). Ger. Offen. (1998)

Abstract
The concns. of pollutants such as CxHy, CO, CO2, NOx, N2O, HCN, NH3, SO2, SO3 and
exhaust particles or solid C, are analyzed by FTIR spectroscopy using polychromatic IR
radiation. The unfiltered exhaust gas is passed into a chamber with heated walls and
irradiated with IR radiation having an absorption spectrum in a range suitable for solid
hydrocarbons, esp. amorphous hydrocarbons or soot.

6-8
Reference List From Literature Search

26. Infrared-spectroscopic studies of the gas phase above the sulfur trioxide-water liquid
system. II. Composition of the the gas phase above the sulfuric acid-sulfur trioxide
liquid system as a function of temperature. Stopperka, Klaus; Kilz, Friedhelm. Tech.
Univ., Dresden, Ger. Z. Anorg. Allg. Chem. (1969), 370(1-2), 59-66.

Abstract
The ir absorption spectra (400-3600 cm-1) were recorded of the gas phase in equil. with liq.
H2SO4-SO3 systems at various concns. of SO3 and at temps. from 30° to the b.p. of the liqs.
(.apprx.200°). Assignments are discussed. Concns. of SO3 were detd. by anal. of the bands
at 2450 and 2780 cm-1; concns. of S3O9, by anal. of peaks of 865 and 1237 cm-1. H2S2O7 was
not detected in the gas phase. Domains of existence of H2SO4, SO3, and S3O9 in the gas
phase were detd. to be 0-30%, 0-100%, and 40-100% free SO3. Factors which govern the
equil. and the thermal decompn. of S3O9 to SO 3 are discussed. S3O9 decreased sharply in
concn. with increasing temp. >70° and was no l onger detected at >100° due to the dissocn.
to SO3.
27. Infrared synthesis of sulfur trioxide by homogeneous gas-phase carbon dioxide laser-
driven reactions. Voicu, I.; Alexandrescu, R.; Morjan, I.; Stoica, M.; Ursu, I.; Jianu, V.
Laser Dep., Inst. At. Phys., Bucharest, Rom. Infrared Phys. (1992), 33(6), 557-62.

Abstract
By using a CO2 laser operated at 10.6 .mu.m, a bimol. reaction was induced in a binary
SO2/O2 gas mixt., with a sensitizer gas (SF6) as photoabsorbing species. The IR prepn. of SO3
was performed in a flowing gas device equipped with a continuous trapping system of the
reaction product. Vibrational energy transfer processes presumably should play a specific
role in SO2 excitation and reaction in the presence of SF6.
28. Infrared tunable diode laser diagnostics for aircraft exhaust emissions characterization.
Wormhoudt, J.; Zahniser, M.S.; Nelson, D.D.; McManus, J.B.; Miake-Lye, R.C.; Kolb,
C.E. Cent. Chem. Environ. Phys., Aerodyne Res. Inc., Billerica, MA, USA. Proc. SPIE-
Int. Soc. Opt. Eng. (1994), 2122(Laser Applications in Combustion and Combustion
Diagnostics II), 49-

Abstract
The Atm. Effects of Stratospheric Aircraft (AESA) component of the NASA High Speed
Research Program (HSRP) will require measurements of trace gas concns. in the exhausts of
high speed civil transport engines. In parallel with the development of these engines by
NASA and its industrial partners, a portable IR tunable diode laser app. was assembled and
tested which is capable of both in-situ and extractive sampling of combustion gas flows. In
the present app., sensitive detection is achieved by rapid frequency scanning and real-time
nonlinear least squares fitting and background subtraction. Sensitivity is further increased
for extractive sampling by an advanced design multiple pass cell which gives longer path
lengths in smaller vols. Observations of a lab. flat flame burner are reported. These
observations and spectroscopic models are used to predict detection sensitivities in exhausts
and other combustion systems.

6-9
Reference List From Literature Search

6.2 Measurement Technique References


1. Guidelines for troubleshooting utility stack opacity. Balfour W D, Rhudy R G, Keeth R J
In: Combined flue gas desulfurization and dry SO2 control symposium. St Louis, MO, USA,
25-28 Oct 1988. 5-267 – 5-281 (1988).
2. Flue gas SO3 determination – Importance of accurate measurements in light of recent
SCR market growth. Bionda J (2002) In: 2002 Conference on SCR and SNCR reduction
NOx control, Pittsburgh, PA, USA. 15-16 May 2002. 14pp (2002)
3. Dew points of simulated flue gases. Barham P, Tye F L, Vasanthakumar A L S, Warn J R
W (1993) Journal of the Institute of energy; 2-8 (Mar 1993)
4. Improvements to the controlled condensation measurement method for sulfuric acid.
Blythe G, Galloway C, Richardson C (1999) In: EPRI-DOE-EPA Combined utility air
pollution control symposium: The Mega Symposium, Atlanta, GA, USA. 16-20 Aug 1999,
5-13 - 5-27 (1999)
5. Flue gas sulfuric acid measurement improvement. Blythe G, Galloway C, Rhudy R
(2001a) In: EPRI-DOE-EPA Combined utility air pollution control symposium: The Mega
Symposium, Chicago, IL, USA. 20-23 Aug 2001, 20 pp (2001)
6. Measurements for the determination of acid dew points and SO3 concentration in the
flue gas utility boilers. Derichs W, Menden W, Ebel P K, (1999) In: VGB
Kraftwerkstechnik: 10, 865-868 (1991)
7. Controlled condensation method: new option for SO3 sampling. DeVito M S, Smith D L
(1991) In: Power: 41-44 (Feb 1991)
8. The sulfuric acid dew point in power station flue gases. Halstead W D, Talbot J R W
(1980) In: Journal of the Institute of Energy; 142-145 (Sept 1980)
9. Continuous measurement of sulfuric acid vapor in combustion gases using a portable
automatic monitor. Jackson P J, Hilton D A, Buddery J H (1981) In: Journal of the
Institute of Energy, 124-135 (Sept 1981)
10. The measurement of acid dew point temperature in power generation boilers. Johnston M
(1982) Sheffield, UK, Land Combustion 22pp (1982)
11. Dewpoint corrosion: mechanisims and solutions. Meadowcroft D B, Cox W M (1985) In:
Dewpoint corrosion Holmes d R (ed), Institute of corrosion science and technology,
Birmingham, UK, pp 17-34 (1985)
12. Personal Correspondence. Muzio L (2004) Fossil Energy Research Corporation, (Ferco)
23342-C South Pointe Drive, Laguna Hills, Ca 92653
13. The Severn Science Analyser to continuous in-stack monitoring of SO3 gas/acid mist.
Nowak R (1998) In: US DOE/FETC conference on formation, distribution, impact and fate
of SO3 in utility flue gas streams. Pittsburgh, PA, USA, 30-31 Mar 1998. 1pp (1998)
14. Controlled condensation for flue gas SO3 measurements. Oda R L, DeVito (1998) In: US
DOE/FETC conference on formation, distribution, impact and fate SO3 in utility flue gas
streams. Pittsburgh, PA, USA, 30-31 Mar 1998. 3pp (1998)

6-10
Reference List From Literature Search

6.3 SO3 Coal-Fired Power Plant Related References


1. Wet electrostatic precipitation demonstrating promise for fine particulate control –
part I. Altman R, Offen G, Buckley W, Ray I (2001a) Power Engineering; 37-39 (Jan 2001)
2. Wet electrostatic precipitation demonstrating promise for fine particulate control –
part II. Altman R, Buckley W, Ray I (2001b) Power Engineering; 42-44 (Feb 2001)
3. Effects of gas stream temperature on homogeneous SO2 to SO3 conversion via natural
gas reburning. Bayless D J, Khan A R (1998) In: 1998 International Joint Power
Conference, Baltimore, MD, USA. 23-26 Aug 1998, vol 1, pp 147-152 (1998)
4. The effects of cofiring natural gas and coal on sulfur retention in ash. Bayless D J,
Schroeder R, Olsen M G, Johnson D C, Peters J E, Krier H, Buckius R O (1996) Combustion
and Flame; 106, 231-240 (1996)
5. Control of sulfur dioxide and sulfur trioxide using by-product of a magnesium-
enhanced lime FGD system. Benson L B, Smith K J, Roden R A, Loch E, Potts J (2003)
In: Combined power plant air pollutant control Mega symposium. Washington DC, USA,
19-22 May 2003, 15 pp (2003)
6. Agent less flue gas conditioning for electrostatic precipitators. Bibbo P (1995) In: 57th
Annual American power conference, Chicago, IL, USA, 18020 Apr 1995. Pp32-37 (1995)
7. Furnace injection of alkaline sorbents for sulfuric acid control. Blythe G, McMIllian R,
Davis J, Lamison M, Rhudy R, Beeghly J, Benson L, Goetz E (2001) In: EPRI-DOE-EPA
Combined utility air pollution control symposium: The Mega Symposium, Chicago, IL,
USA. 20-23 Aug 2001, 20pp (2001)
8. Full-scale testing of the SBS process for flue gas SO3 control. Blythe G M, Dombrowski
K D, Miller S D, Meserole F B, Rhudy R G, Glancy D, Hall T (2003) In: Combined power
plant air pollutant control Mega symposium, Washington DC, USA, 19-22 May 2003. 23pp
(2003)
9. Acid mist causes problems for FGD systems. Buckley W P, Altshuler B (2002) In: Power
Engineering; 132-136 (Nov 2002)
10. Factors determining the effectiveness of magnesia based additives for SO3 control in oil-
fired boilers. Burdett N A, Gliddon B J, Hotchkiss RC, Squires R T (1983) In: Institute of
Energy conference on the effectiveness of fuel additives, Leatherhead, UK, 4-5 Oct 1983,
11pp (1983)
11. Sulfur trioxide issues at TVA’s Cumberland fossil plant. Burnett T A (1998) In: US
DOE/FETC conference on formation, distribution, impact and fate of SO3 in utility flue gas
streams, Pittsburgh, PA, USA, 30-31 Mar 1998. 2pp (1998)
12. Airheater operating experience on a utility boiler firing international market coals after
the retrofit of selective catalytic NOx reduction equipment. Cooper J, Reid D, Sorensen
H (1999) In: Power-Gen International 99, New Orleans, LA, USA, 30 Nov – 2 Dec 1999, 21
pp (1999)

6-11
Reference List From Literature Search

13. Operating experience with SCR’s and high sulfur coals & SO3 plumes. Duellman D M,
Erickson C A, Licata T (2002) In: ICAC FORUM – Cutting NOx, Houston, TX, USA, 12-
13 Feb 2002. 21pp (2002)
14. Current work on SO3 emissions from selective catalytic reduction systems.
Erickson C A, Jambhekar R (2002) In: 2002 Conference on SCR and SNCR reduction
NOx control, Pittsburgh, PA, USA. 15-16 May 2002. 14pp (2002)
15. Application of wet type electrostatic precipitator for utilities coal-fired boiler.
Fujishima H, Tsuchiya Y (1993) In: Tenth particulate control symposium and Fifth
international conference on electrostatic precipitation, Washington DC, USA, 5-8 Apr 1993,
38-1 - 38-13 (1993)
16. SO2 conversion rate of deNOx catalysts – effects on downstream plant components and
remedial measures. Gutberlet H, Hartenstein H, Licata A (1999) In: EPRI-DOE-EPA
Combined utility air pollution control symposium: The Mega Symposium, Atlanta, GA,
USA, 16-20 Aug 1999, 15-49 – 15-63 (1999)
17. The role of the combustion air preheater in an effective multipollutant control strategy.
Hamel B, Rhodes R, Benson S A, Mccollor D P (2003) In: Combined power plant air
pollutant control Mega symposium, Washington DC, USA, 19-22 May 2003. 19pp (2003)
18. Chemical reduction of SO3 particulates and NOx emissions. Kukin I, Bennett R P (1977)
In: Journal of the Institute of Fuel; 50; 41-46 (1977)
19. Effect of boiler operations on sulfuric acid emissions. Levy E K (1998) In: US
DOE/FETC conference on formation, distribution, impact and fate of SO3 in utility flue gas
streams. Pittsburgh, PA, USA. 30-31 Mar 1998, 5pp (1998)
20. Modern power station practice, chemistry and metallurgy, Littler D J, Davies E J,
Kirkby F, Myerscough P B, Wright W (eds) (1992) Oxford, UK, Pergamon Press, vp (1992)
21. Estimates of particle formation and growth in coal-fired boiler exhaust – 1
observations. Mueller S F, Imhoff R E (1994) Atmospheric environment; 28 (4); 595-602
(1994)
22. FGC as a means for cost-effective ESPs for low sulfur coals. Porle K, Bradburn K, Bader
P (1996) In: Sixth international conference on electrostatic precipitation. Budapest, Hungary
18-21 Jun 1996, 417-426 (1996)
23. Implication of the toxic release inventory for electric utilities. Rubin E S, Berkenpas M
B (1999) In: EPRI-DOE-EPA Combined utility air pollution control symposium, The Mega
Symposium, Atlanta, GA, USA, 16-20 Aug 1999, 19-1 - 19-9 (1999).
24. Tackling particulates. Sanyal A., Parker K. Power Engineering International; 98-102
(Jul/Aug 1999)
25. Factors affecting sulfuric acid emissions from boilers. Sarunac N, Levy E. In: EPRI-
DOE-EPA Combined utility air pollution control symposium, Atlanta, Georgia, USA, 16-20
Aug 1999, p. 5-1 – 5-11 (1999).

6-12
Reference List From Literature Search

26. Role of calcium and iron on SO3 formation in coal-fired power plants. Seniro C L,
Johnson S A. In: US/DOE/FETC conference on formation, distribution, impact and fate of
SO3 in utility flue gas streams, Pittsburgh, PA, USA, 30-31 Mar 1998. 2 pp (1998)
27. The kinetics of SO3 formation in oil-fired boilers. Squires R T. (1982). Journal of the
institute of energy; 41-46 (Mar 1982)
28. Emissions of sulfur trioxide from coal-fired power plants. Srivastava R K, Miller C A,
Erickson C, Jambhekar R (2002). In: Power-Gen mercury, trace elements and particulate
matter conference, Orlando, FL, USA, 10-12 Dec 2002, 26 pp (2002)

6-13
A
MEASUREMENT OF NH3 IN COAL FIRED POWER
PLANTS USING TUNABLE DIODE LASER
SPECTROSCOPY IN CONJUNCTION WITH A
SPECIALLY DESIGNED DUST ATTENUATING PROBE

AWMA Control Number C1-8492-CFD1AD22F032

John T. Pisano (1)


(2)
Richard M. Himes
Chuck Dene (2)
(1)
Matthew R. Smith
Dennis R. Fitz (1)
(1)
University of California at Riverside (UCR)
1084 Columbia Avenue, Riverside, CA, 92507
(2)
Electric Power Research Institute (EPRI)
3412 Hillview Avenue, Palo Alto, CA 94304

A-1
Measurement of NH3 in Coal Fired Power Plants Using Tunable Diode Laser Spectroscopy in Conjunction with a
Specially Designed Dust Attenuating Probe

Abstract

Fossil fueled electric power generating facilities have been incorporating SCR and SNCR
systems in their operations to reduce NOX emissions and in the process have developed a need to
measure NH3, a by-product from the associated NOX reduction reactions. Conventional
extractive measurements of NH3 using wet chemical methods cannot provide the response time
necessary to optimize these NOX reduction systems during dynamic operation of the facility.
Inherently a near real time response for NH3 is required to control the NOX reduction
technologies. Additionally coal fired power plants have significant dust loadings, which make
optical measurements difficult and pathlength limited. Tunable diode laser spectroscopy offers
the time response and specificity necessary required for in-situ measurements of NH3, but this
type of optical measurement has been previously limited by the high dust loadings in these
applications. EPRI and UCR have developed a probe that reduces the amount of dust in the
optical path by approximately 70%. This technology effectively removes the previous
impediments to cross duct sampling at path lengths of up to 40-50 feet (12-16 m). The probe is
designed to easily interface with a commercially available tunable diode laser controller to
provide measurements in both permanent and portable applications.

Introduction

Thermal electric power facilities using fossil fuels have historically been major point sources of
NOx emissions. Nitrogen oxides are formed through thermal fixation of nitrogen and oxygen in
air at high combustion temperatures. In coal-fired power plants, additional NOx is produced
from the large amounts of nitrogen in the coal itself, when combined with excess oxygen during
the coal devolatilization process. Coal fired utilities account for approximately 25% of the NOx
emissions in the USA.

Nitrogen oxides are classified as primary pollutants and are subject to emission regulations by
the Clean Air Act Amendments of 1990. Additional amendments have recently been proposed to
lower NOx emissions during the months of May to September based on tonnage caps equivalent
to 0.15 lbs NOx per million Btu of energy generated. Although facilities burning high volatile
content Powder River Basin coals have achieved these emission levels with combustion
modifications alone, facilities firing eastern bituminous coals have been limited to nominal 0.20
– 0.40 lb/MBtu NOx emission levels. These limitations are resultant from a higher fixed carbon
content, which requires longer furnace residence time for char burnout, as well as potential
limitations with respect to the lower furnace stoichiometry, due to corrosion concerns with
medium to high sulfur bearing coals. Post combustion options under consideration for facilities
firing eastern bituminous coals include selective catalytic reduction (SCR) or non-catalytic
reduction (SNCR). The process effectively converts NOx to water and elemental nitrogen by
reaction with NH3 .

4NO + 4NH3 + O2Æ4N2 + 6H2O Equation A-1

2NO2 + 4NH3 + O2Æ3N2 + 6H2O Equation A-2

A-2
Measurement of NH3 in Coal Fired Power Plants Using Tunable Diode Laser Spectroscopy in Conjunction with a
Specially Designed Dust Attenuating Probe

The ammonia may be added directly, or in the form of urea and water. In SNCR urea is
introduced through injectors at the furnace exit at flue gas temperatures between 1900 – 2400ºF
(1038 – 1316ºC). SCRs are usually located downstream of the boiler with the ammonia added
after the economizer. Reaction occurs at a temperature of about 650ºF (343ºC).

In both approaches, some NH3 escapes the reaction zone unreacted. This can result from the
addition of excess NH3, or from incomplete mixing of the gases. The excess unreacted NH3 is
referred to as NH3 slip, which can have negative balance-of-plant impacts, including:
1. In the case of fuels with appreciable sulfur content, excess NH3 slip can lead to the formation
of ammonium salts, which can foul the air heater leading to unit degradation or forced outage
to allow a water wash of the air heater.
2. Incorporation of ammonia into the fly ash from coals with acidic ash, rendering it unsaleable
to the cement industry. Disposal of the ash in ponds or in landfills, can lead to NH3 pollution
of ground water.
3. Ammonia slip emitted from the stack can react with chlorides from the coal, leading to a
white NH4Cl plume.

Continuous measurements of NH3 downstream of the reaction zone are required to minimize NH3
slip. These measurements must be made rapidly, since the concentration is not uniform across
the ducts or across the SCR. Measurements should be representative of the entire cross section.
The distribution across the duct should be known so that appropriate adjustments can be made at
the reagent injection nozzles. Extractive sampling methods cannot meet this requirement. Non-
extractive, optical methods are capable of providing more representative sampling and generally
require much less maintenance

Tunable diode lasers (TDLs) are fabricated on a tiny piece of semiconductor material typically
only 0.5 mm square. The diode lasers themselves produce laser light at specific wavelengths that
can be continuously tunable over narrow wavelength intervals. Adjusting the device temperature
enables tuning of the laser output, and even finer tuning can be achieved by adjustment of the
laser operating current (Schiff et al., 1992). The development of diode lasers themselves has
been rapid over the past few years, especially for application in communications transmission.
Diode lasers with outputs ranging from visible to mid-infrared wavelengths are commercially
available.

TDLs offer significant advantages when applied to spectroscopic systems. As a light source, they
provide the highest available power density in a spectrally narrow window. The fundamental,
vibrational absorption bands in the infrared region (IR) contain a number of discrete rotational
lines, the width and shape depending on temperature and pressure. For some gases the lines are
well separated; for others, pressure broadening results in overlapping lines. Larger molecules
often possess so many closely spaced lines that thermal or Doppler broadening is sufficient to
cause line overlapping (Fried et al., 1991). This inhibits most conventional forms of IR
spectroscopy. However, TDLs have spectral resolutions smaller than Doppler line widths, and
advantage can be taken of this inherent property (Cooper et al., 1987). The fraction of light
intensity transmitted through a gas is given by the familiar Beer-Lambert law:

P/Po = exp (-σ(ψ)NL) Equation A-3

A-3
Measurement of NH3 in Coal Fired Power Plants Using Tunable Diode Laser Spectroscopy in Conjunction with a
Specially Designed Dust Attenuating Probe

Where P and Po are the transmitted and incident powers respectively, L is the absorption path
length in centimeters and N is the concentration of the absorbing molecules in number of
molecules per cubic centimeter. σ(ψ) is the wavenumber dependent absorption cross section in
square centimeters per molecule.

The development of TDL spectroscopy is growing, especially by companies such as Unisearch


Associates Inc., Seimens and Norsk Electro Optics. Their fabrication of commercially available
instrumentation has led to many commercial applications. In addition, the potential of TDLs has
only recently begun to be investigated. Since TDLs are significantly more efficient than
conventional gas or solid-state lasers in converting electrical energy into light, they require less
input power for a given light intensity. They are miniature devices and require further development
in their thermal control and signal analysis to make them ultimately an inexpensive sensor.

What makes this application for tunable diode laser spectroscopy particularly challenging is the
very heavy particle loading in the gas stream. Coal fired power plants provide a very dirty
environment for open path measurements.

Dust can affect performance in two ways:


1. Dust or particulate matter in the beam path scatters or obscures the light.
This affects the signal-to-noise ratio, and, consequently, the detection limit.
2. Dust can accumulate on the optical components (windows), obscuring the beam and limiting
the light reaching the detector across the measurement path.

Experimental
TDL Instrumentation:

For the experiments reported herein, a Unisearch Associates Inc., LasIR® tunable diode laser
controller was used. The analyzer houses the laser, a computer, a power supply, and the signal
processing electronics. It can also house a calibration/audit system for routine calibration checks.
The LasIR analyser is normally placed in a relatively clean control room, remote from the
measurement location. Light from the laser is transmitted to the remote optics via a fibre optic
cable, which may be up to 1000 m long. The signal from the detector, mounted in the remote
optics, is returned to the control unit via a co-axial cable. The control unit can be multiplexed to
one or more optical paths depending on the application. Multiplexing also permits the
simultaneous measurements at a number of locations with a single instrument.

This system has been used in a number of industrial and environmental applications.
®
The instrument used to measure ammonia slip, the LasIR , contains for its laser source, a NIR
GaAs laser with center frequency at 1.512 which is known to be suitable for NH3 measurements.

A-4
Measurement of NH3 in Coal Fired Power Plants Using Tunable Diode Laser Spectroscopy in Conjunction with a
Specially Designed Dust Attenuating Probe

Site One: Gas-Fired Utility

The initial site chosen to demonstrate the viability of using tunable diode laser technology for
NH3 measurements was a natural gas fired utility. This application provided the combination of
a lower average NOx and particulate concentrations on which to investigate the LasIR
performance since higher instrument sensitivity is required for gas-fired utilities as compared to
coal fired equivalents. The lower dust loadings, permitted the use of a longer path length to
improve the detection limit.

The unit was installed at a gas-fired power plant to continuously measure ammonia (NH3) post
SCR on the west duct. Since the laser is fiber coupled the instrument containing the laser and
associated electronics was placed in the instrument shed that contained the NOx measuring
instrumentation. The configuration employed a monostatic type optical system incorporating
single sided transmitting/receiving optics and a retro-reflector diametrically opposite.

As shown in Figure A-1, the optical configuration employed at the gas fired power plant. The
laser output is transmitted to the optical assembly via 9-micron fiber optic cable, connected via
FC-APC connectors. The beam diverges at f10 upon leaving the fiber optic connector and is
nearly collimated but still slightly diverging via a lens assembly. The still slightly diverging
beam is then returned to the transmitting side via a 2.0“ retro-reflector where it is collected and
focused to the detector via a 1.5“ off axis parabolic mirror. The detector signal is returned to the
controller via coaxial cable. In this mode the effective optical path length is double the duct
width, and since the duct width is approximately 10 meters, an optical path length of 20 meters is
achieved. With this optical path length, minimum detection limits (MDL’s) of better than 25
ppbV (parts per billion by volume) were routinely measured.

An important characteristic of a tunable diode laser is that the wavelength at which it emits
changes slightly with the electric current passing through it. This makes it possible to scan across
the entire selected absorption line of the target gas as well as to a region where the target gas
does not absorb. This allows for correction of any anomalies of the absorption signal from effects
other than the target gas absorption. By initially scanning the absorption feature prior to target
gas absorption, deviations in overall laser intensity can be measured.

This is significant for measuring optically over path lengths where there would be significant
deviations in optical intensity as well as some perturbation from dust, although this gas fired unit
had little dust when compared to the highly dust laden environment of a coal fired unit.
Therefore, the laser intensity that the detector sees will be dependent on the opacity of the
emission stream and will have to be monitored continuously.

A-5
Measurement of NH3 in Coal Fired Power Plants Using Tunable Diode Laser Spectroscopy in Conjunction with a
Specially Designed Dust Attenuating Probe

D etector DUCT R etroreflector


Eye-piece

Coaxial O ptical
output Fiber

Figure A-1
Optical Configuration Employed at Gas Fired Power Plant for Continuous
TDL Measurements of NH3

The laser contained in the system control unit is tuned across the spectral feature selected to
monitor the gas species of interest. It is mounted directly on a thermoelectric cooler and forms
the laser head assembly. Fixing the temperature of the laser is significant as temperature changes
yield coarse changes in the wavelength emission. Since the tuning range is small, diode lasers are
temperature-locked, and tuning is done, by varying the current through the laser crystal itself.
Fine-tuning is achieved by a cyclic current ramp that facilitates repetitive scanning of a given
laser frequency range.

Multi-scan averaging improves the sensitivity of the system. To maximize sensitivity the laser
current is further modulated by additional frequencies combined with two-tone or 2f detection.
Using two-tone FM modulation techniques approximating second derivative analysis, better
detection limits are possible (Muecke et al., 1991).

Initial results obtained on the gas-fired unit were encouraging, showing 100% operation when
the boiler was operating, and only losing alignment once the system went off-line due to loss of
alignment across the duct from differential thermal contraction. The system tracked the
ammonia injection rate as shown by the plot in Figure A-2.

A-6
Measurement of NH3 in Coal Fired Power Plants Using Tunable Diode Laser Spectroscopy in Conjunction with a
Specially Designed Dust Attenuating Probe

NH3 slip after SCR-Unit


at a Gas Fired Power Plant
25 1.0
NH3 Flow
0.5

Ammonia Flow (gpm)


20
NH3 Mixing Ratio

0.0
(ppmV)

15
NH3 Mixing Ratio -0.5
10 T corrected
-1.0

5
-1.5

0 -2.0
9/4/02 9/7/02 9/10/02 9/13/02 9/16/02 9/19/02
12:00 12:00 12:00 12:00 12:00 12:00
Date and Time (PDT)

Figure A-2
NH3 Slip Measured Compared to Reagent Flow at the Gas Fired Utility Over a Two-week
period

Note that when the unit was off-line, the system alignment deviated to a point where there was
no measurable signal. However, once the boiler came back on-line, alignment returned to a point
where measured total laser intensity levels were consistent with those measured prior to the
outage.

Site Two: 700 MW Coal-Fired Boiler

The application of tunable diode laser spectroscopy via the probe assembly to coal fired utilities
represents both a challenge and opportunity for this technology. Particulate matter emissions
make it difficult to traverse the entire duct with a conventional unshielded laser emission.

Fiber-coupled TDL technology is amenable to a multiplexed application where a single laser can
measure up to eight paths over a one-minute time frame.

In the application discussed, our objective for the coal-fired application of TDL technology was
to design a cross duct probe utilizing single laser multiplexing and also employing a design that
reduced the amount of dust in the optical path, so that we could measure a single path in discrete
intervals.

The assembly contained three separate probes, over the length of the diameter of the duct, in the
direction of flow, to permit the NH3 measurement over this path. This length per tube summed
would be like measuring over the whole diameter of the duct, but provided spatially resolved
line-of-sight averages.

A-7
Measurement of NH3 in Coal Fired Power Plants Using Tunable Diode Laser Spectroscopy in Conjunction with a
Specially Designed Dust Attenuating Probe

Details of the design are under patent review, and further information cannot be released at this
time. The following discussion will be limited to the results obtained with the multi-point probe.

The test site chosen was a 700 MW coal fired boiler, that has typical dust loadings of 5 grains/ft3.
The unit was undergoing an evaluation of SNCR technology.

The probes were installed on the flue gas duct at a location downstream of the SNCR injectors
and boiler economizer, and approximately 40 feet prior to the air heaters. At this location the
duct width was approximately 12 feet in depth. The LasIR controller was placed approximately
200 feet away from the probe assembly in a temperature controlled room, and was fibercoupled
to the probes.

The results for one of the sampling days are presented in Figure A-3. The four data plots
include(a) the NH3 slip in parts per million by volume (ppmV) as measured by the TDL 1-2
meter probe, (b) the urea injection rate in gallons per hour (gph), (c) the water flow mixed with
the urea in gallons per hour (gph), and (d) boiler load in megawatts. The ammonia slip numbers
represent 30-second averages.

Looking at the NH3 levels, they track the urea injection rate fairly well. There were three distinct
levels of urea injection during this days testing, with the highest slip numbers corresponding to
the time of highest urea injection rate, between 12:00 and 16:00. Note that when the urea
injection rate was lowered at 16:00, the average slip levels also decreased. The slip also was also
very unstable during boiler ramp up from 3:00 to 6:00 AM, indicating that changes in boiler load
will affect the resulting slip levels.

Over a four-day sampling period, measurements of NH3 were also being made by standard
impinger wet chemical techniques. The NH3 measurements by the wet chemical technique and
the data presented, were taken at times where there were concurrent TDL measurements.

It should be noted that the TDL data represents the integrated probe data over the first half of the
duct whereas the wet chemical measurements took data from a multi-point sampling probe over
the complete width of the duct. Also the TDL measurements are instantaneous whereas the wet
chemical measurements require a longer integration period.

A-8
Measurement of NH3 in Coal Fired Power Plants Using Tunable Diode Laser Spectroscopy in Conjunction with a
Specially Designed Dust Attenuating Probe

16
NH3 Slip (ppmV)
12

300
0
11 (GPH)

00

00

00

00

00
0

0
200

:0

:0

:0

:0
0:

3:

6:

9:

0:
12

15

18

21
2

3
/1

/1

/1

/1

/1
2

2
/1

/1

/1

/1
100
Urea

11

11

11

11
11

11

11

11
30
0
H21O (GPH)

00

00
00

00

00

0
:0 20

:0

:0

:0

0:
0:

3:

6:

9:

12

15

18

21
12

3
2

/1
/1

/1

/1

2
1/

/1

/1

/1

/1
10

11
11

11

11

11

11

11

11
0
800
11 (MW)

600
00

00

00

00

00
0

0
:0

:0

:0

:0
0:

3:

6:

9:

0:
12

15

18

21
400
2

3
/1

/1

/1

/1

/1
2

2
Load

/1

/1

/1

/1
11

11

11

11
11

11

11

11
200
0

00
00

00

00

00

0
:0

:0

:0

:0

0:
0:

3:

6:

9:

12

15

18

21

3
2

/1
/1

/1

/1

/1

2
/1

/1

/1

/1

11
11

11

11

11

11

11

11

11

Figure A-3
NH3 Slip Measured Compared to Reagent Flow, Load and Reagent Dilution (H2O) at a
700 MW Coal Fired Power Plant

Figure A-4 represents a linear regression plot showing the inter-comparison between the wet
chemical NH3 measurements to that as measured by the probe tunable diode laser NH3
measurements. The slope is nearly one except for an offset of approximately 1.4 ppmV. The
correlation coefficient R2 =0.738 for the 13 sample points. This shows overall good basic
agreement between the two techniques.

The slightly higher ammonia measurement obtained with wet chemical sampling techniques is
possibly attributable to HNCO slip from the urea reagent. As the ion-selective electrode is non-
specific, it detects all gas species released when the acidic sample solution is made basic during
the analysis procedure.

A-9
Measurement of NH3 in Coal Fired Power Plants Using Tunable Diode Laser Spectroscopy in Conjunction with a
Specially Designed Dust Attenuating Probe

10
y = (1.04±0.19) x - (1.4±1.3)
8 R 2 = 0.738
NH3 by TDL Measurement

0
0 2 4 6 8 10
NH3 by W et Chemical Measurement

Figure A-4
Linear Regression Plot Showing the Inter-Comparison Between the Wet Chemical
NH3 Measurements to That as Measured by the Probe Tunable Diode Laser NH3
Measurements.

Site 3: 375 MW Coal-Fired Boiler and 145 Coal-Fired MW Boiler

The next set of experiments, were conducted at another coal fired power plant over two distinct
boilers. This coal-fired power plant was also undergoing an evaluation of SNCR technology for
reduction of NOX emissions. In this application, project objectives were to identify conditions
for optimum NOX reduction where NH3 slip levels remained at the 5 ppmV level or less.

During the SNR evaluation, two separate boilers were tested. These included a 375 Mw
tangential design boiler firing a nominal 1% sulfur coal. Full load flue gas temperature was
nominally 2,300°F (1260 C), while reduced load flue gas temperatures at 200 MW were 1,900°F
(1038 C).A smaller capacity 145 MW tangential design boiler firing a similar coal was also
tested. Full load flue gas temperatures were nominally 2,200°F (1204 C), while reduced load
flue gas temperatures at 60 MW were 1,700°F (927 C).

The SNCR reduction technology requires that liquid urea (CO(NH2)2) be injected directly into
the furnace through injectors that are mounted directly onto the furnace wall. In this case for
both units, up to five injectors were placed on the front furnace wall with two additional injectors
placed on the sidewall. The reagent injection occurs at the flue gas temperatures between

A-10
Measurement of NH3 in Coal Fired Power Plants Using Tunable Diode Laser Spectroscopy in Conjunction with a
Specially Designed Dust Attenuating Probe
1,700°F – 2,300°F (927 – 1260 C), depending on load, and requires specific drop size
distribution and/or urea dilution to release the reagent at an appropriate process temperature

The experimental setup was designed to measure NH3 slip at two of the eight ports on the boiler
continuously. Two probes were manufactured for the project and the LasIR unit was a
multiplexed version configured for two-path sampling. Two ducts were sampled on both the 375
MW unit and the 145 MW unit, referenced as A duct and B duct. The actual experiments and
methodology behind these experiments are reported elsewhere

During the SNCR evaluation project, 195 experiments were conducted that required NH3 slip
measurements of approximately 10-20 minutes each. For each experiment, the maximum,
minimum and average slip numbers were taken over the line-of-sight of sample measurement.

Two of the results are presented here as they show the usefulness of real time NH3
measurements.

In the first case, one experiment was adjusting urea flow to reduce slip down to desired levels
when injectors were first operated during low load conditions. Initial urea flow was at 1.9 gph
per injector, at which time NH3 slip levels exceeded 30 ppmV.

Subsequent manipulations reduced the slip to just over 5 ppmV. Figure A-5 shows the complete
data set for this experiment.

The correlation coefficient is a measurement of how well the measured absorbance spectra
correlates to the stored calibration spectrum. Note, that the correlation coefficient, is always
higher than 0.8 at 5 ppmV or better.

Since the data presented is the raw data with 15-second integration times, longer integration
times, improve the signal to noise and also increase the correlation. Minimum detection limits,
better than 0.5 ppmv, were also achieved with the 15-second integrations.

One of the design parameters for the probe assembly is allowance for easy insertion and
extraction from any standard 4-inch port. This was shown to be useful should a traverse of the
whole duct be required. During one of the experiments a test for determining just how
representative the slip measurements from the single port selected for continuous sampling was
undertaken.

A-11
Measurement of NH3 in Coal Fired Power Plants Using Tunable Diode Laser Spectroscopy in Conjunction with a
Specially Designed Dust Attenuating Probe

40.0 1.00

35.0 0.75

64% NOx Reduction at this slip


30.0 0.50
NH3 Mixing Ratio (ppmV) (15-second averages)

NH3 Mixing Ratio


Correlation Coefficient
25.0 0.25

Correlation Coefficient
Adjust Urea Flow

20.0 0.00

Adjust Urea Flow

15.0 -0.25

10.0 -0.50

5 Front Wall 40-40 Injectors


Urea Flow Turned On
5.0 -0.75
Baseline

Baseline
0.0 -1.00
16:04 16:19 16:33 16:48 17:02 17:16 17:31 17:45
Time (hh:mm:ss)

Figure A-5
Shows the NH3 Slip Results and Correlation Coefficient of the TDL Data During the
Adjustment of Urea Flow With Newly Installed Injectors.

During this experiment a total of eight ports were sampled in less than a 2-hour period. While
traversing each of the ports, one of the probes was continuously monitoring NH3 slip at port 3 to
assess SNCR process stability during the measurements. The average slip for port 3 was 5.3
ppmV for the entire period the traverse was performed. Port 3 was shown to have the highest
slip values as the other ports were found to have average values between 1.8 and 4.8, a factor of
3 different.

This shows that measuring at one place is not truly indicative of the average slip, as the average
slip for duct A was found to be 2.8 ppmV and for duct B, 2.9 ppmV, both below that which was
measured at the port selected for continuous measurements. The results are presented in Table
A-1 below.

A-12
Measurement of NH3 in Coal Fired Power Plants Using Tunable Diode Laser Spectroscopy in Conjunction with a
Specially Designed Dust Attenuating Probe

Table A-1
Lists Results From the Traverse of All 8 Ports.

Traverse NH3 Slip (ppmV)


Results High Low Average A Duct 3

A Duct 1 4.1 2.5 3.0 5.5


A Duct 2 3.9 2.4 2.8 5.2
A Duct 3 6.4 4.5 5.3 5.3
A Duct 4 2.7 1.9 2.3 5.2
B Duct 5 2.5 2.0 2.3 5.5
B Duct 6 2.9 2.1 2.5 5.3
B Duct 7 5.2 4.3 4.8 5.5
B Duct 8 2.2 1.6 1.8 5.4

Site 4 635 MW Coal-Fired Boiler

The probe design developed for the portable applications was employed in designing a
permanent installation version, which would ultimately facilitate continuous monitoring of NH3
and other pertinent gases over longer paths.

A prototype version has been designed and installed across a 10-meter path in a duct at a 635
MW coal-fired boiler. Typical dust loadings at this boiler are 4-6 grains/ft3.

Tests were done to show the improvement in optical transmission that the probe yields. Refer to
Figure A-6, which shows the laser signal power (converted to volts) as measured with the probe
and without the probe over this path length.

At these dust loadings the signal power is reduced by 90% over just 2 meters, and reaches
extinction by 3 meters. Where as with the probe it was possible to measure over the entire 10-
meter path.

A-13
Measurement of NH3 in Coal Fired Power Plants Using Tunable Diode Laser Spectroscopy in Conjunction with a
Specially Designed Dust Attenuating Probe

Signal Pow er Com parison

1.20

1.00

0.80
Signal Power (volts)

0.60
No Probe
With Probe
0.40

0.20

0.00
0 2 4 6 8 10

-0.20
Optical Pathlength (m eters)

Figure A-6
Shows the Which Shows the Laser Signal Power as Measured With the Probe and Without
the Probe Over Pathlengths Up to 10 Meters.

Conclusion

The Unisearch Associates Inc. LasIR TDL performed satisfactorily and with the probe
configuration yielded MDL’s of better than 1.0 ppmV per meter of optical path measured. The
system was found to be easily integrated with the probe assembly design via fiber optic and
coaxial cable connection.

The probe/TDL configuration proved useful for the investigators in trying to determine the
optimal configuration for the SNCR system under evaluation by the respective power plants.

The probe design can be employed in both a portable version as well as a permanent installation.

For the portable version, the latest development of the probe permitted easy insertion and
extraction from a standard 4-inch port, this indicates that traverse measurements can be
undertaken in a short period of time.

A-14
Measurement of NH3 in Coal Fired Power Plants Using Tunable Diode Laser Spectroscopy in Conjunction with a
Specially Designed Dust Attenuating Probe
The probe effectively removes up to 70% of the dust in the optical path, allowing increased laser
return power to achieve optical path lengths of better than 10 meters.

With the NH3 slip measurement system the investigators were able to evaluate the NOX reduction
capabilities of the SNCR system under evaluation and achieve NOX reduction of better than 30%
with slip levels at or below 5 ppmV.

The cost for the amount of measurements taken is much less than what it would cost using
standard wet chemical methods.

For larger ducts where there is a need for spatial information instead of just an integrated one, the
multiplexing of the LasIR with the advantage of the probe having the capability of measuring
over the entire duct width, can ultimately lead to mapping the NH3 concentrations over the entire
duct. This would provide useful information in identifying problem areas such as individual
injector fouling or inequality reagent distribution. Such measurements could eventually be used
to control the rate of reagent injection.

For SCR applications, the enhanced pathlength, would provide better sensitivity with NH3
measurements reaching detection limits of 0.1 ppmV at 10 meters. This would help identify
catalyst perturbations and be used to monitor SCR performance. Presently some TDL systems
have been installed in SCR applications, but are limited in performance by the pathlength used.

Although, the work that has been done in this project has been for looking at NH3 slip, the TDL
technique can be applied to the measurement of many gases such as CO, CO2 and O2 for
combustion efficiency investigations.

EPRI is sponsoring additional research in this area with the goal of obtaining continuous,
spatially resolved measurements so as to enable the tuning of individual burners as well as post
combustion NOx control technologies such as SNCR or SCR.

References

Abbe, A.; Heger, R.; Pisano, J.T.; and Franz, M. (2000) Intercomparison of three separate
technologies for the measurement of HF stack emissions from the HAW primary aluminum
smelter. Light Metals, Annual Conference Proceedings.

Bartone, L.M.; Broek, S.; Fuller, A.M.; and Pisano, J.T. (2000) Measurement of HF in a high
temperature environment during the recycling of spent potliner using the Vortec smelting
system. Air and Waste Management Association, Annual Conference Proceedings.

Cooper, D.E., and Warren, R.E. (1987) 2-Tone optical heterodyne spectroscopy with diode
lasers-theory of line shapes and experimental results. J. Opt.Soc.Am., B4, 470-480.

Fried, A.; Drummond, J.R.; Henry, B.; and Fox, J.A (1991) Tunable diode laser spectrometer fpr
high precision concentration and ratio measurements of long lived atmospheric gases. In
Measurement of Atmospheric Gases. (H.I.Schiff, Ed.), SPIE 1433, 143-156.

A-15
Measurement of NH3 in Coal Fired Power Plants Using Tunable Diode Laser Spectroscopy in Conjunction with a
Specially Designed Dust Attenuating Probe

LasIR : A near infrared diode laser system for measurements of trace gas concentrations in stack
emissions and ambient air, general operating manual. Unisearch Associates Inc. March 2000

Muecke, R.; Werle, P.; Slemr, F.; and Pretl, W. (1991) Comparison of time and frequency
multiplexing techniques in multicomponent FM spectroscopy. In Measurement of Atmospheric
Gases. (H.I.Schiff, Ed.), SPIE 1433, 136-142.

Schiff, H.I.; Mackay, G.I.; Karecki, D.R.; Pisano, J.T.; and Nadler, S.D. (1992) Measurements of
H2O2 in the Grand Canyon by tunable diode laser spectrometry during the Navajo generating
station visibility study. In Monitoring of Gaseous Pollutants by Tunable Diode Lasers, pp 21-30.
Kluwer Academic, Dordrecht, The Netherlands.

A-16
B
LABORATORY AND FIELD MEASUREMENTS OF H2SO4
AND SO3 USING FTIR

B-1
Laboratory and Field Measurements of
H2SO4 and SO3 using FTIR

November 6, 2002

Industrial Monitor and


Control Corporation
800 Paloma Dr.
Suite 100
Round Rock, TX 78664
Industrial Monitor and Control Corporation

Introduction

This letter report presents results of two tests to demonstrate H2SO4 and SO3 monitoring
with FTIR. The first set of tests were laboratory tests conducted at Imacc’s facilities in
Austin, Texas. These were controlled tests to demonstrate detection and to test
compatibility of material. Subsequently field tests were undertaken at a power plant in
Ohio. These tests attempted to reproduce the laboratory results and to perform real-time
duct measurements at the plant. The results of both tests indicate that FTIR can detect
and quantitate concentrations of H2SO4 and SO3 while also monitoring other stack gases
of interest such as H2O, CO, CO2, NO, NO2, SO2, and HCl. The success of the tests
indicate that a second phase should be undertaken to look at longer term data establishing
the variations of concentration levels and to compare the FTIR results with other
monitoring techniques such as CEM systems and wet chemistry for the acids.

Laboratory Tests in Austin

The week of September 23 a controlled experiment was set up in the Imacc lab to:

a) Test the FTIR’s ability to see H2SO4 and SO3 and


b) Try to isolate problems with transmitting these gases through Teflon or
stainless lines as used at power plants in the past.

The setup fabricated for the tests is shown in Figure 1. It used a propylene glycol
temperature-controlled bath to hold a flask filled with H2SO4 at constant temperature.
Glass beads were added to the bath to aid in diffusion of the air flow through the acid
achieving a better equilibrium. A pump was placed at the output of the FTIR cell and this
was used to pull ambient air through the H2SO4 flask, through a 50 ft. heat trace line and
through the FTIR cell. All lines were heated to 150C and the cell was held at 185C to
avoid condensation of the H2SO4. The flask was filled with 96% H2SO4 and it was
heated to 80°C. At this temperature, the equilibrium vapor pressure of pure H2SO4 in
water should be 41.4 ppm and that of SO3 should be 0.13 ppm. The values reported by
the FTIR are shown in Figures 2 and 3 for various flow conditions and line temperatures
used throughout the day.
Heated Teflon
Line

Rotometer

Ambient Air
Input FTIR Cell
X
Heat Trace
Line

Pump
Exhaust
Constant
Temperature
H2SO4 Bath
Figure 1 Experimental setup used to test for H2SO4 and SO3 in the laboratory.

80

70

60

50
H2SO4 (ppm)

40

30

20

10

0
8:00 9:00 10:00 11:00 12:00 13:00 14:00 15:00 16:00
Time

Figure 2 Time-series plot of FTIR-reported concentrations of H2SO4 throughout the day


5

4.5

3.5

3
SO3 (ppm)

2.5

1.5

0.5

0
8:00 9:00 10:00 11:00 12:00 13:00 14:00 15:00 16:00
Time

Figure 3 Time-series plot of FTIR-reported concentrations of SO3 throughout the day

To confirm that the FTIR was seeing H2SO4/SO3 and not some contaminant, collected
spectra were viewed directly to identify the spectral signatures. Figures 4 through 6 show
these spectra. In Figure 4 the portion of the observed spectrum from 2400 to 2530 cm-1 is
shown on top (in red) and this can be compared to the library reference spectra for both
SO3 in the middle (in blue) and SO2 on the bottom (in purple). It is clear that both SO3
and SO2 are present. Figures 5 and 6 show two regions where H2SO4 absorbs. Figure 5
is the 1100 cm-1 to 1260 cm-1 region and figure 6 the 800 to 940 cm-1 region. In both of
these, the data is on top (in red) and the reference spectrum for H2SO4 on the bottom (in
blue). Except for a little water vapor in the 1200 cm-1 region, both measurements are
identical to the H2SO4 reference. Clearly H2SO4 was also observed.

Because both SO3 and H2SO4 were being observed, we attempted to alter flow and line
temperatures to see their effect on line transmission. Table 1 shows that the

Table 1
Observed H2SO4 Concentration as a Function of Flow Rate

Time Flow (SCFH) H2SO4 Observed


8:35 2.5 9
8:49 5.0 13
9:07 7.5 27
9:20 10.0 41.8
9:40 12.5 63.5
**Fri Sep 27 10:44:43 2002
0.052

0.050

Abs 0.048

0.046

*GAS5.040 SO3 7897.3 ppm 0.2m/760T/25C 8 mbar (not broadened p=6.08 Torr ) 20 cm 25 C 2 cm-1 => 0.5 cm-1

0.03
Abs

0.02

0.01

-0.00
SO2 800 ppm 5.2m/650t/185C Synthetic
0.03

0.02
Abs

0.01

-0.00
2520 2500 2480 2460 2440 2420
Wavenumbers (cm-1)

Figure 4 Plot of experimental data at 10:44 am (top-red plot) compared to the SO3
reference (middle – Blue) and the SO2 reference (bottom- purple).

0.40 **Fri Sep 27 10:44:43 2002

0.35

0.30

0.25
Abs

0.20

0.15

0.10

0.8 817ppm H2SO4 22.1m/650/185C ,10 ppm SO3 (SO3 = 5.0ppm cal against 7897 ppm SO3) Checks w/ Applied Optics vol17 No.7 p976.

0.7

0.6

0.5
Abs

0.4

0.3

0.2

0.1

1260 1240 1220 1200 1180 1160 1140 1120 1100 1


Wavenumbers (cm-1)

Figure 5 Plot of experimental data at 10:44 am (top-red plot) compared to the H2SO4
reference (bottom – Blue) in the 1100 to 1260 cm-1 region.
**Fri Sep 27 10:44:43 2002

0.30

0.25

0.20
Abs

0.15

0.10

0.05
817ppm H2SO4 22.1m/650/185C ,10 ppm SO3 (SO3 = 5.0ppm cal against 7897 ppm SO3) Checks w/ Applied Optics vol17 No.7 p976.
0.7

0.6

0.5
Abs

0.4

0.3

0.2

0.1

940 920 900 880 860 840 820 800


Wavenumbers (cm-1)

Figure 6 Plot of experimental data at 10:44 am (top-red plot) compared to the H2SO4
reference (bottom – Blue)in the 800 to 940 cm-1 region.

concentration of sulfuric increased with flow (as did SO3). This was originally thought to
be exclusively due to line equilibration effects which are known to exist for sulfuric in
most lines. However, the highest concentration of sulfuric observed actually exceeded
the equilibrium vapor pressure of the acid at 80°C. This cannot happen if the bath is in
equilibrium. It was found, that at the highest flow rates liquid was being splashed within
the acid bath and droplets were being pulled into the extraction line. Since the lines were
held at 120°C and the FTIR cell at 185°C this would allow a higher vapor pressure to be
established if liquid got into the line. At lower flow rates, it is believed that this
entrainment did not occur so the increase in concentrations up to about 41 ppm may be
indicative of increased line loss at lower flow. However, greater agitation of the bath
with increased flow also occurred and this could produce a similar effect if equilibrium
was not achieved in the bath. At about 10:45 the flow was dropped to 6.5 SCFH from the
high of 12.5 SCFH. An immediate drop in H2SO4 and SO3 was observed as shown in
Figures 2 and 3. This could again be line loss or agitation effects. Typically line loss is
seen only initially when gases line H2SO4 are first admitted to a line. Over time the line
is usually passivated and the losses cease. Total passivation would take longer at slow
flow and less time at high flow.

Losses caused by cold spots in the transfer lines could be a problem at a field site. To test
for these types of losses, the heated Teflon line between the acid bath and the heat trace
line (see figure 1) was cooled in successive steps. Starting about 12:15 this line was
cooled from its high of 167°C to about 120°C and this temperature was held from about
12:27 to 12:40. The temperature was then dropped to 110°C and held there from 1:00 to
1:15. Finally the line was dropped to 45°C achieving this temperature by 1:32. As
observed in figures 2 and 3, the concentrations during this period did not drop but in fact
increased slightly. This implied that loss in the Teflon lines is not very temperature
sensitive, a result we did not expect.

At 13:43 the flow was stopped and a 6” stainless steel line was added between the Teflon
line and the heat trace. The flow was held constant at 6.5 SCFH, as it had been before the
stainless was added, and the line heated to 200°C. At about 14:15 the temperature of the
stainless was stable and a concentration of H2SO4 of 51.0 ppm was observed. This was
essentially equal to what it had been before the insertion of the stainless line. The
stainless was then cooled, as had been done with the Teflon line. By 14:30 the stainless
had dropped to 94°C and the H2SO4 dropped steadily from 51 ppm to 45 ppm. This
indicated the stainless does have more severe losses than the Teflon at lower
temperatures. The stainless was heated again and by 14:37 it had reached a temperature
of 120°C with the H2SO4 rising again to 49 ppm, recovering almost back to its high value
at 200°C. At this point the cell was flushed and the concentrations of all gases fell to
essentially zero.

Field Tests

The week of October 21, tests were undertaken at a power plant in Ohio to:

a) Reproduce the laboratory tests discussed above; and


b) Monitor an actual duct exhaust in the plant.

Controlled Bath Tests

Initially the FTIR cell was disassembled, the mirrors checked, and the system aligned.
The apparatus used in Austin (Figure 1) had been shipped to Ohio. This apparatus was
set up in the FTIR trailer to provide a known source of H2SO4 and SO3 for injection into
the lines and the FTIR cell. Two 90 foot lines were in place at unit 2 going from the
FTIR-trailer to the duct. These were connected together at the duct so the H2SO4/SO3
from the bath could be passed up the lines and back down to the FTIR through a line 180
feet long. This would maximize any line effects if they existed. The bath was then filled
with 96% H2SO4 and heated to 90°C. Initially, outside air was used to feed the bath
through a short 50 foot heat-traced line to preheat the air. During this testing, little
H2SO4 was seen in the FTIR. In tracing the problem it was found that some Swagelok
connections were loose and needed to be redone. It was also suspected that the 50 foot
line preheating the air was loosing sample, it was therefore removed. After these changes
SO3 and H2SO4 were seen in the FTIR. The system was then run over night with the bath
feeding the 180 feet of line. It was thought that this would condition the lines minimizing
any line effects. As seen in Figures 7 and 8, the water vapor levels steadily dropped
throughout the night while the H2SO4 and SO3 levels rose. The maximum reached by the
SO3 was 0.6 ± 0.1 ppm while for H2SO4 it was 12.5 ± 0.4. At about 4:15 am the detector
ran out of liquid nitrogen so the signal started dropping and acquisition stopped.
16000

14000

12000

10000
H2O (ppm)

8000

6000

4000

2000

0
22:00 23:00 0:00 1:00 2:00 3:00 4:00 5:00
Time

Figure 7 Water vapor levels observed during the over-night run of the bath.

16

14

12
H2SO4
H2SO4 - SO3 (ppm)

10

SO3
2

0
22:00 23:00 0:00 1:00 2:00 3:00 4:00 5:00
Time

Figure 8 H2SO4 and SO3 observed during the over-night run of the bath.
On Thursday morning, the FTIR detector was re-filled with liquid nitrogen and the
reported levels were close to what they had been at 4:00 am. Apparently, no further
increase in concentration occurred with the longer gas flow. Figures 9 through 11 show
the reported concentrations of water, H2SO4, and SO3 both initially and as adjustments
were made to the system.

The period up to 8:30 am was run under the same conditions as the over-night runs. At
8:30 the air flow to the bath was increased from 15 scfh to 25 scfh. This resulted in an
immediate increase in both H2SO4, and SO3. Equilibrium was established at about 9:00
am and the peaks observed were 2.13±0.2 ppm for SO3 and 26.2±0.6 ppm for H2SO4.
The flow was then dropped again to 15 scfh at 9:02 am and as expected the
concentrations again fell but they did not reach as low a level as they had initially. In
stead the H2SO4 stopped at about 19±0.5 ppm and the SO3 1.6±0.7 ppm. A further
reduction in flow to 7.5 scfh at 9:30 further reduced the concentrations. This effect is
believed to be a result of the heating of the flask in the bath with higher air flow. There
was no heater on the upper portion of the flask so its temperature was dictated by the gas
flow and turbulence in the flask. Higher air temperature in the flask would increase the
air dew point and allow it to carry more H2SO4 /SO3 away.

300000

250000

200000
H2O (ppm)

3 Sigma Error
150000

100000

50000

0
8:00 9:00 10:00 11:00 12:00 13:00
Time

Figure 9 Water vapor observed during H2SO4 bath injection of sample lines and
FTIR cell. The error bars are 3 σ errors returned by the FTIR.
30

25

20
H2SO4 (ppm)

3 Sigma Error
15

10

0
8:00 9:00 10:00 11:00 12:00 13:00
Time

Figure 10 H2SO4 observed during bath injection of sample lines and FTIR cell. The
error bars are 3 σ errors returned by the FTIR.

4.5

3.5

3
3 Sigma Error
SO3 (ppm)

2.5

1.5

0.5

0
8:00 9:00 10:00 11:00 12:00 13:00
Time

Figure 11 SO3 observed during bath injection of sample lines and FTIR cell. The
error bars are 3 σ errors returned by the FTIR.
Because these flows were at very low water levels both H2SO4 and SO3 were seen. In the
duct the water levels will be higher and it is expected that the SO3 will disappear, being
converted to H2SO4. To simulate higher water content runs, the flow from the bath was
stopped at 10:16 am and sufficient water added to the bath to produce roughly a 76% acid
solution. One interesting result of stopping the flow was the slow decrease in H2SO4, and
to a lesser degree in SO3, during this period. This is indicative of adsorption of the acid
on the walls of the cell reducing its concentration with time. This type of effect is seen
with many polar compounds and is not unexpected. It is for this reason that continuous
flow is required to properly sample these gases.

After addition of water to the bath, its temperature was increased to 120ºC. At this
temperature, equilibrium tables indicated that about 92800 ppm of water should be
produced with about 12 ppm of H2SO4 and very low levels of SO3. At 10:57 am the bath
was up to temperature and flow to the lines and the FTIR cell was resumed. It can be
seen that the concentration of all constituents rose rapidly. After the initial rise, water
started dropping while H2SO4 leveled off at about 2.4 ppm while the SO3 seemed to level
off at <1 ppm although the error bars on this data made it hard to determine a definitive
behavior. The H2SO4 stabilized at a lower level than predicted by the tables. This was
believed to be due to the cool upper portion of the flask, as in the previous tests, which
would hold the air dew point low. As a result, the flow to the bath was again increased at
11:42 am from 15 scfh to 25 scfh. It can be seen that the sulfuric acid rose immediately
equilibrating at a level of about 8 ppm while water and SO3 continued to drop. While this
was not meant to be a calibrated run, the 8 ppm was believed to be reasonable agreement
with the tables because the exact concentration of the acid was not known.

The plots also show that the error bars on H2SO4 and SO3 increased after water was
added to the bath. This was expected. Spectrally water is an interferrant to both H2SO4
and SO3. The error bars displayed are based on the spectral residual (left over “stuff”)
after the analysis software has accounted for all the gases it can in the spectrum. If the
software cannot perfectly cancel out strong water vapor features in a region of analysis,
the error bars will increase for all compounds analyzed for in this region. The error bars
indicate the worst possible error on the concentrations of the analytes, but in reality errors
are usually much less than this.

Duct Tests

Based upon the success of seeing both H2SO4 and SO3 under controlled conditions, the
FTIR was moved to a unit in the plant that was in operation. The instrument was placed
on the back of a small trailer and was located on the ground under the unit inlet duct to an
absorber module. A 50 foot heat-trace line connected the FTIR to the duct.

The 50 foot line originally used for preheating air to the water bath had to be used at the
duct. Because of suspicions that this line was absorbing H2SO4, prior to use it was steam
cleaned using high purity steam available in the plant. The lab work in Texas also
indicated that stainless steel was not a good material to use with sulfuric, consequently a
probe was constructed with a copper tube as the outer sheath with a nylon inner tube for
extraction of the gas sample. Nylon is not an ideal material either but it was the only
material readily available. Later a Teflon line was obtained and substituted for the nylon.

Figures 12 through 19 show the results of the testing. The system was set up about
6:00 pm and was intended to run all night. However, a power failure occurred after the
system was left and, as a result, data was obtained only to 8:20 pm. The drop in
concentrations seen at 8:00 pm was the switching of the probe from Nylon to Teflon.
During this time the flow was stopped and ambient air leaked into the cell. After
switching to Teflon, the concentrations did not change significantly except for H2SO4 and
HCl. Initially, the sulfuric acid concentration rose significantly but then fell back down
appearing to equilibrate at about the same level as before the switch. HCL, on the other
hand, showed a slight increase in concentrations which seemed to persist. Both the
laboratory and these results seem to indicate that Teflon should be used throughout the
sampling train and that all other materials should be avoided. It may be worthwhile to
consider Teflon coating of the FTIR cell because of the slow losses seen while static
filled.

Table 2 outlines the concentrations reported by the FTIR for each compound observed in
the duct. The water, CO2, NOx, and SO2 levels are close to what was expected for this
duct. H2SO4 seems to be higher than previously assumed and the HCl is comparable to it.
The values reported here were validated by manually analyzing FTIR spectra and all
values seemed correct. However, this type of validation does depend on the accuracy of
the library references. The only reference that has not been validated several times in the
past is that of H2SO4. Sulfuric acid is hard to generate and equally hard to maintain at an
accurate concentration in test systems. The reference used here was generated using a
constant temperature bath of H2SO4 in a circulating system that forced thermal and
concentration equilibrium on the entire system. This reference has been checked against
available literature data and seems to be accurate. Further validation would require
generation of comparative references.

Table 2
Observed gas concentrations in the plant at an output duct

Compound Concentration Nylon


(ppm)
Water 60,000
Carbon Dioxide 114,000
Carbon Monoxide 2.1 to 2.4
Nitric Oxide 328 to 349
Nitrogen Dioxide 1.7 to 1.9
Sulfur Dioxide 2330 to 2370
Sulfuric Acid 36.3 to 37.5
Hydrochloric Acid 31.1 to 32.2
80000

70000

60000

50000
H2O (ppm)

40000
3 Sigma Error
30000

20000

10000

0
18:00 18:15 18:30 18:45 19:00 19:15 19:30 19:45 20:00 20:15 20:30
Time

Figure 12 Observed Water Vapor in the unit duct on 24 October 2002. The error bars
are the 3 σ errors returned by the FTIR.

140000

120000

100000
CO2 (ppm)

80000
3 Sigma Error
60000

40000

20000

0
18:00 18:15 18:30 18:45 19:00 19:15 19:30 19:45 20:00 20:15 20:30
Time

Figure 13 Observed CO2 in the unit on 24 October 2002. The error bars are the 3 σ
errors returned by the FTIR.
7

5
3 Sigma Error
CO (ppm)

0
18:00 18:15 18:30 18:45 19:00 19:15 19:30 19:45 20:00 20:15 20:30
Time

Figure 14 Observed CO in the unit duct on 24 October 2002. The error bars are the 3 σ
errors returned by the FTIR.

500

450

400

350

300
NO (ppm)

250

200
3 Sigma Error
150

100

50

0
18:00 18:15 18:30 18:45 19:00 19:15 19:30 19:45 20:00 20:15 20:30
-50
Time

Figure 15 Observed NO in the unit duct on 24 October 2002. The error bars are the 3 σ
errors returned by the FTIR.
4.5
3 Sigma Error
4

3.5

3
NO2 (ppm)

2.5

1.5

0.5

0
18:00 18:15 18:30 18:45 19:00 19:15 19:30 19:45 20:00 20:15 20:30
Time

Figure 16 Observed NO2 in the unit duct on 24 October 2002. The error bars are the 3
σ errors returned by the FTIR, these are large because of H2O interference.

3000

2500

2000
SO2 (ppm)

3 Sigma Error
1500

1000

500

0
18:00 18:15 18:30 18:45 19:00 19:15 19:30 19:45 20:00 20:15 20:30
Time

Figure 17 Observed SO2 in the unit duct on 24 October 2002. The error bars are the 3 σ
errors returned by the FTIR.
70

60
3 Sigma Error
50

40
H2SO4 (ppm)

30

20

10

0
18:00 18:15 18:30 18:45 19:00 19:15 19:30 19:45 20:00 20:15 20:30
-10
Time

Figure 18 Observed H2SO4 in the unit duct on 24 October 2002. The error bars are the
3 σ errors returned by the FTIR.

40

35

30

25
HCl (ppm)

3 Sigma Error
20

15

10

0
18:00 18:15 18:30 18:45 19:00 19:15 19:30 19:45 20:00 20:15 20:30
Time

Figure 19 Observed HCl in the unit duct on 24 October 2002. The error bars are the 3 σ
errors returned by the FTIR.
Conclusions

These tests indicate that:

a) Both H2SO4 and SO3 can be detected and quantified by FTIR in power plant
ducts;
b) Teflon is the best material to use in lines and couplings for these compounds;
c) Stainless seems to show significant losses for H2SO4 and SO3 particularly
below 120C to 150C, it should be avoided throughout the sampling system;
d) A probe utilizing only Teflon is probably necessary for low less transfer;
e) Conditioning of the lines through prolonged exposure to H2SO4 appears to be
necessary to avoid line losses;
f) Water vapor, at the levels observed in the plant ducts, does not appear to
restrict measurements at the concentrations needed, except for possibly NO2.
The NO2 analysis procedures need to be further refined;
g) Validation of the H2SO4 references should be undertaken to assure accuracy
of the current library spectra. This would require generation of spectra under
carefully controlled conditions in which the concentrations of the acid would
be well known.

Because the FTIR can see all compounds of concern at power plants, long term
monitoring of the full suite of compounds should be initiated to establish statistics on
the maximum, minimum, and variance of the gas concentrations and to develop a data
base of comparative data between the FTIR and the CEM systems.
C
EPRI CONTINUOUS SO3 MEASUREMENT
DEMONSTRATION PROJECT OPPORTUNITY

C-1
project opportunity

Continuous SO3 Measurement Demonstration

Participants in this project will gain:


• First hand experience in the application
and integration of a continuous SO3
monitor to minimize additive cost for
SO3 mitigation and ensure proper
ESP operation
• Design and cost-benefit information from
the use of a continuous SO3 monitor on
1–2 plants in their system

Visible Plume from High Sulfur Coal-Fired Boiler with SCR Even when additives are not utilized, accurate measurement
of SO3 emissions can serve to monitor and detect an
increase in SO3 levels resulting from a change in fuel
NOx emissions from coal-fired boilers in ozone non-
specifications, combustion considerations, SCR operation,
compliant regions are being restricted to levels that often
etc., so that appropriate corrective actions can be taken.
require the application of Selective Catalytic Reduction
(SCR) technology. One consequence of the catalyst within A continuous SO3 monitor would also help plants that burn
SCR systems is the oxidation of a fraction of the SO2 in the a low-sulfur coal to minimize the amount of SO3 injected to
flue gas to SO3. A number of balance-of-plant impacts can condition fly ash for high ESP removal efficiencies.
result from the increased SO3 emissions, which include (a)
increased propensity to form ammonium bisulfate within the PROJECT SUMMARY The current project scope entails
air heater, (b) corrosion of downstream surfaces at the combination of EPRI’s phase discrimination probe
temperatures below the acid dew point, and (c) increased (developed in conjunction with the University of California
opacity due to the emission of acid aerosols from the stack. – Riverside) with a Fourier Transform Infrared (FTIR)
Control of SO3 often entails the use of an additive injected analyzer. EPRI believes this combination could enable the
into the flue gas. Control of the additive feed rate, however, in-situ measurement of SO3 over a line-of-sight within the
is based on manually collected SO3 measurements during flue gas duct anywhere downstream of the economizer
initial testing of additive use, with no current method outlet. Specific project tasks would include:
available to optimize its use over the normal range and • Host site visit to identify sampling location, physical
variability of unit operation. The availability of a access constraints, and overall line-of-sight dimensions
continuous SO3 monitoring method would provide to enable the fabrication of a probe for installation
significant cost savings by enabling the minimization of within the duct.
additive feed rates used to control SO3 emissions. This • Design and fabrication of probe and optics assemblies
minimum is achieved when the resulting SO3 concentrations to be used at host site.
are reduced enough to avoid a visible plume while still
• Installation of ports at identified duct location to
being high enough to condition the fly ash for high
provide probe access, as well as simultaneous wet
efficiency collection by the ESP.
chemical sampling access for monitor validation.

Post-Combustion NOx Control and Continuous Emissions Monitoring March 2004


• Field test documenting FTIR monitor performance PRICE OF PROJECT The total project cost is estimated to be
for SO3, based on a comparison against wet chemical $100,000–$200,000, which will be offset with $30,000 of
reference methods. base funding at each of the first two test sites. The estimated
• Optional task can be incorporated upon successful project costs are inclusive of monitor rental, site specific
demonstration of continuous SO3 measurement phase discrimination probe design and fabrication,
capability that assesses the variability in SO3 emissions development of optics for SO3 measurement across the
as a function of unit load, SCR operating conditions, probe, reference SO3 measurements using controlled
and possibly coal. condensate methodology, and all reporting. The range in
costs takes into account a range of site-specific parameters,
DELIVERABLES A report summarizing the results of the as well as a potential demonstration at two plants.
demonstration and its performance relative to current
reference SO3 measurement methods. Based on monitor PROJECT STATUS AND SCHEDULE This project is scheduled

performance, recommendations will be made as to how a to begin in 2004 and will take approximately six months to
system might be implemented successfully to continuously complete, depending upon unit access for installation of the
monitor SO3 at 1–2 plants belonging to each participating phase discrimination probe at the demonstration site.
company, along with a cost-benefit analysis for those
CONTACT INFORMATION For more information, contact
applications.
the EPRI Customer Assistance Center (EPRI CAC) at
BENEFITS OF PARTICIPATION Participating organizations 800.313.3774 (askepri@epri.com).
will receive design and performance information before
TECHNICAL CONTACT Richard Himes, 949.766.8470,
they are made available to other program members,
(rhimes@epri.com), or Chuck Dene, 650.855.2425,
enabling them to design additive systems and begin
(cdene@epri.com).
benefiting from the associated cost savings sooner. Such
savings would be proportional to the reduction in additive
use. With additive costs for a 500 MW medium sulfur coal-
fired plant with an SCR estimated to be $500,000–
$1,000,000 per year to achieve a clear stack, a 20%
reduction in reagent use would save $100,000–$200,000
per year.

© 2004 Electric Power Research Institute (EPRI), Inc. All rights reserved.
Electric Power Research Institute and EPRI are registered service marks of
the Electric Power Research Institute, Inc. EPRI. ELECTRIFY THE WORLD
is a service mark of the Electric Power Research Institute, Inc.

Printed on recycled paper in the United States of America

1009582

EPRI • 3412 Hillview Avenue, Palo Alto, California 94304 • PO Box 10412, Palo Alto, California 94303 USA
800.313.3774 • 650.855.2121 • askepri@epri.com • www.epri.com
Export Control Restrictions
Access to and use of EPRI Intellectual Property is granted
with the specific understanding and requirement that
responsibility for ensuring full compliance with all applicable
U.S. and foreign export laws and regulations is being
undertaken by you and your company. This includes an
obligation to ensure that any individual receiving access
hereunder who is not a U.S. citizen or permanent U.S.
resident is permitted access under applicable U.S. and foreign
export laws and regulations. In the event you are uncertain
whether you or your company may lawfully obtain access to
this EPRI Intellectual Property, you acknowledge that it is
your obligation to consult with your company’s legal counsel
to determine whether this access is lawful. Although EPRI
may make available on a case by case basis an informal
assessment of the applicable U.S. export classification for
specific EPRI Intellectual Property, you and your company
acknowledge that this assessment is solely for informational
purposes and not for reliance purposes. You and your
company acknowledge that it is still the obligation of you and
your company to make your own assessment of the
applicable U.S. export classification and ensure compliance
accordingly. You and your company understand and
acknowledge your obligations to make a prompt report to
EPRI and the appropriate authorities regarding any access to
or use of EPRI Intellectual Property hereunder that may be
in violation of applicable U.S. or foreign export laws or
regulations.

About EPRI
EPRI creates science and technology solutions for
the global energy and energy services industry.
U.S. electric utilities established the Electric Power
Research Institute in 1973 as a nonprofit research
consortium for the benefit of utility members, their
customers, and society. Now known simply as EPRI,
the company provides a wide range of innovative Programs: 1009812

products and services to more than 1000 energy- Post-Combustion NOx Control
related organizations in 40 countries. EPRI’s Continuous Emissions Monitoring
multidisciplinary team of scientists and engineers
draws on a worldwide network of technical and
© 2004 Electric Power Research Institute (EPRI), Inc. All rights reserved. Electric Power Research
business expertise to help solve today’s toughest
Institute and EPRI are registered service marks of the Electric Power Research Institute, Inc.
energy and environmental problems. EPRI. ELECTRIFY THE WORLD is a service mark of the Electric Power Research Institute, Inc.

EPRI. Electrify the World Printed on recycled paper in the United States of America

EPRI • 3412 Hillview Avenue, Palo Alto, California 94304 • PO Box 10412, Palo Alto, California 94303 • USA
800.313.3774 • 650.855.2121 • askepri@epri.com • www.epri.com

Potrebbero piacerti anche