-------
TABLE 1-5. Concluded
TEST
Thallium
Tin
Titanium
Uranium
Vanadium
Zinc
Zirconium
Chloride
Fluoride
Cyanide
Nitrate
Sulfate
Ammon i urn
Coal Analysis
C*
H*
0%
N%
S%
H2OX
Ash*
HHV, J/g
Btu/lb
BASELINE
<2.6
<6.4
6.1xl03
<3.9
2.6xl02
4.3x102
1.9xl02
2.7xl02
84
<1.3
<3.9
6.5xl03
<5.3
63.13
4.27
7.34
1.38
2.19
2.04
19.60
26288
11302
BIAS (Test 1)
<2.7
<6.7
5.7xl03
44
2.3xl02
5.9xl02
2.6xl02
4.1xl02
3.5xl02
0.3
24
3.9x103
7.2
63.46
4.24
7.97
1.13
1.75
2.34
19.09
26363
11334
BOOS (Test2)
<2.1
<5.1
3.6xl03
<2.1
1.6xl02
8.4xl02
6.8xl02
8.6xl02
1.2xl02
<1.3
7.7xl02
2.1xl03
1.4xl02
64
4.23
7.11
1.38
2.13
2.58
18.49
26521
11402
1-22
-------
baseline and low NO firing. Two levels of NO reduction were
A A
tested. Retrofit bias burner firing gave a 32 percent NO reduction,
A
and operation with the upper row of nozzles on air only gave a 38 percent
NO reduction. However, the percent NO reduction with bias firing
/\ A
should be tempered somewhat by the fact that there was a slight decrease
in fuel nitrogen content for that test. The furnace efficiency either
remained constant or increased slightly (due to lower excess air) under
low NO operation. There was no appreciable increase in
A
carbon-in-flyash with NO controls. It should be mentioned that these
A
tests were for short periods, so the long term operability under these low
NO conditions was not necessarily validated.
A
Unfortunately, due to limited coal supplies, the coal sulfur
contents were not constant throughout the test program, as noted in
Table 1-5. Nevertheless, the data do indicate that SO,, emissions are
not significantly affected by low NO firing. This is certainly the
A
case when comparing the BOOS test with baseline. And the drop in S02
emissions with the bias test can be attributed to the decrease in fuel
sulfur content, since 98 percent of the sulfur introduced into a utility
appears in flue gas as an oxide (Reference 1-28).
Comparing particulate emissions under bias firing with those under
baseline would indicate that low NO firing would have no significant
A
effect. However the observed decrease in particulate emissions under BOOS
firing cannot be fully explained by the lower fuel ash content or the
lower boiler firing rate. Nonetheless, the bias test when reinforced with
data from several other field test programs do show that particulate
emissions and particle size distribution are relatively unaffected by low
NO firing.
A
For the majority of elements listed in Table 1-5, the changes in
emission rates between baseline operation and low NO firing were within
the accuracy of the analysis and are not judged to be significant.
Notable exceptions are the Teachable nitrates and ammonium compounds.
Here it is possible that local fuel rich conditions under low NO
A
operation suppress reduced nitrogen compound oxidation normal to baseline
operation. Organic species analyses were inconclusive, though total
organic emissions increased with low NO firing. Reference 1-1 presents
A^
the analysis results for the other waste streams — cyclone ash, ESP ash,
1-23
-------
and bottom ash slurry. Table 1-6 lists the DS values for those inorganic
species or compounds where DS_>!. It is evident that the gaseous
pollutants, particularly S02 and NOX, dominate the potential toxicity
of the flue gas stream. Of the trace metals, arsenic shows the highest
DS, but none of the metals show any large change under low NO
A
conditions. As may be expected, S03 decreased under low NOX operation
and reduced N compounds increased.
The total weighted discharge severity for the inorganic component
of four waste streams of the boiler are compared in Table 1-7. Clearly
the flue gas stream dominates the TWOS with the solid streams 3 orders of
magnitude potentially less toxic, according to the model. With low NO
/\
firing, the flue gas stream TWOS is reduced, primarily due to the decrease
in NO concentration. The TWDS's for the other waste streams either
J\
decreased or were constant when going to low NO firing. As mentioned
A
earlier, more data are needed for waste stream organic composition before
the discharge severity for organic compounds, relative to inorganics, can
be estimated.
From the application of the source analysis model to the admittedly
sparse data base of a few short tests on a single coal-fired boiler, the
results indicate that NO controls are generally beneficial, reducing
rt
the overall adverse environmental impact of waste streams. These results,
along with the general indications from other reported tests, tend to
confirm that combustion modification NO controls are environmentally
A
sound, though work remains to confirm and correct any potential adverse
environmental impacts from incremental emissions.
1.8 CONCLUSIONS
Modifying the combustion process conditions is currently the most
cost-effective and best demonstrated method of effecting 20 to 60 percent
reductions in NO emissions from utility boilers. Table 1-8 summarizes
the capabilities of combustion modification NO controls. The methods
A
in the current control technology and advanced technology categories are
listed in preference of application. They were selected based on an
assessment of their effectiveness, operational, energy, cost, and
environmental impact, and commercial availability or R&D status.
In the current technology category, low NOX burners or off
stoichiometric combustion through overfire air addition (OFA) is the
1-24
-------
TABLE 1-6. FLUE GAS DISCHARGE SEVERITY — INORGANICS: 180 MW
TANGENTIAL COAL-FIRED UTILITY BOILER
N0y
X
so2
so3
CO
co2
Be
Ba
As
Ti
N (Mainly NH4)
so,
Chlorides
BASELINE
129
322
15
0.77
30
4.5
4.5
48
1
0.07
6.5
0.6B
BIAS
84
269
13
1.1
31
5.5
3.4
39
0.95
0.22
3.9
1
BOOS
73
324
9.6
0.80
32
3.6
3.0
41
0.60
6.1
2.1
2.1
TABLE 1-7. TOTAL WEIGHTED DISCHARGE SEVERITY (g/s) ~ INORGANICS:
180 MW TANGENTIAL COAL-FIRED UTILITY BOILER
Flue Gas
Cyclone Ash
ESP Ash
Bottom Ash Slurry
Total
BASELINE
4.3xl07
1.9x10"
6.1x103
5.7x10"
4.3xl07
BIAS
3.5xl07
1.6x10"
e.ixio3
5.3x10"
3.5xl07
BOOS
3.7xl07
1.6x10"
S.lxlO3
A.2xlD"
3.7xl07
1-25
-------
TABLE 1-8. COMBUSTION MODIFICATION NOX CONTROLS: BEST AVAILABLE
CONTROL TECHNOLOGY (BACT) AND ADVANCED TECHNOLOGY
BACT
Advanced
Technology
Fuel
Coal
Oil
Gas
Coal
Oil
Control Technique
Overfire aira
Low NOX burners
Low NOX burners plus
overfire air
Burners out of service or
overfire air
Flue gas recirculation plus
overfire air
Burners out of service or
overfire air
Flue gas recirculation plus
overfire air
Ammonia injection (1983)b
(combined with BACT
combustion modifications)
Advanced low NOX burners
(1985)
Advanced burner/furnace
concepts (1985)
Ammonia injection (1983)
(combined with BACT
combustion modifications)
NOX Control Level ,
ng/J (lb/106 Btu)
258 (0.6)
215 (0.5)
172 (0.4)
129 (0.3)
86 (0.2)
129 (0.3)
43 (0.1)
129 (0.3)
86 (0.2)
60 (0.15)
43 (0.1)
aAs manufacturers acquire more experience with LNB, they are now
recommending LNB over OFA.
^Estimated date of commercial availability of demonstrated technology.
1-26
-------
preferred technique for retrofit application to coal-fired units, with the
use of new low NO burners, or new burners in combination with OFA,
favored for new units. While current technology can achieve 172 ng/J (0.4
lb/10 Btu) for coal-firing, Table 1-8 indicates that advanced
techniques have the potential of reducing NO to 60 ng/0 (0.15 lb/10
A
Btu). However, ammonia injection, advanced low N0y burners, and
A
advanced burner/furnace concepts are several years away. Current
technology for oil- and gas-fired boilers can reduce NO to the
relatively low levels of 86 to 43 ng/J (0.2 to 0.1 lb/106 Btu),
respectively.
Potential problems with the use of conventional combustion
modifications have concerned possible adverse effects on boiler
efficiency, load capacity, furnace wall tube corrosion and slagging,
fouling, carbon loss, steam and tube temperatures, and flame stability and
vibration. However, recent field experience has shown that adverse
effects can be minimized to acceptable levels with proper care in design
for retrofit applications, and largely eliminated in new unit designs.
Another area of concern with combustion modification NOX controls
is possible increase in incremental emissions of other pollutants to the
environment. Recent test data with conventional techniques seem to
indicate that low NO firing has negligible effects on emissions of most
A
pollutants other than NO . Low NO firing does indeed lower the
X A
overall potential environment impact of the source. However, there are
areas of continued concern, such as possible increased organic emissions.
More extensive field testing will be required to identify and better
quantify these emissions, and compare these results with developing
information in the health effects area.
Finally, conventional combustion modifications are indeed
cost-effective means of control for NO , raising the cost of electricity
A
less than 1 percent in most cases. Furthermore, the initial capital
investment required should also only be of the order of 1 percent or less
of the installed cost of a boiler. With the exception of post combustion
NH, injection, advanced techniques (such as advanced low NO burners
O A
and advanced burner/furnace concepts) have projected costs in the same
range as conventional combustion modifications. Therefore, preferred
1-27
-------
current and projected combustion modification techniques are not expected
to have any significant adverse economic impact.
1.9 RECOMMENDATIONS
Preferred conventional combustion modifications are indeed
recommended for reducing NO emissions from utility boilers, with
A
minimal adverse environmental, operational, and cost impacts. However,
longterm testing and monitoring of field applications/demonstrations
should be continued. Although the issue of possible increased corrosion
with staged combustion has been largely resolved in short-term tests,
long-term corrosion testing, as under current EPA programs, should be
completed to definitively establish that low NO firing does not have
A
any adverse effects. Boiler efficiency should be closely monitored during
field applications to give guidance to control developers on minimizing or
eliminating efficiency losses. The current data base indicates that
efficiency losses of zero to 0.5 percent are possible. The exact number
is of significance; for example, a 0.25 percent loss in efficiency can
translate to one-third of the annualized cost to control.
Finally, the data gaps on the effect of NO controls on
A
incremental emissions are just now beginning to be addressed. Field
testing, with special emphasis on incremental emissions such as trace
metals and organics, on representative utility boiler/control applications
should continue.
Research and development efforts on new combustion modification
technology, such as advanced staged combustion, low NOX burners and
burner/furnace concepts, should continue; they have the potential of
further NO reduction capabilities with minimal adverse impacts.
1-28
-------
REFERENCES FOR SECTION 1
1-1. Water-land, L. R., et al., "Environmental Assessment of Stationary
Source NOX Control Technologies ~ Final Report," Acurex Draft
Report FR-80-57/EE, EPA Contract 68-02-2160, Acurex Corp., Mountain
View, CA, April 1980.
1-2. Mason, H. B., et al., "Utility Boiler NOX Emission
Characterization," in Proceedings: Second NOX Control Technology
Seminar. EPRI FP-1109-SR, Electric Power Research Institute, Palo
Alto, CA, July 1979.
1-3. "Powerplant and Industrial Fuel Use Act of 1978," Public Law
95-620, November 9, 1978.
1-4. "Proposed Rules to Implement the Powerplant and Industrial Fuel Use
Act," Federal Register 43-FR-53974, November 17, 1978.
1-5. "Standards of Performance for New Stationary Sources; Electric
Utility Steam Generating Units," Federal Register 44-FR-33580, June
11, 1979.
1-6. "Standards of Performance for New Stationary Sources (Lignite-Fired
Steam Generators)," Federal Register 43-FR-9276, March 7, 1978.
1-7. Salvesen, K. G., et al., "Emission Characterization of Stationary
NOX Sources. Volume I: Results," EPA-600/7-78-120a, NTIS PB 284
520, June 1978.
1-8. Burrington, R. L., et al., "Overfire Air Technology for
Tangentially Fired Utility Boilers Burning Western U.S. Coal,"
EPA-600/7-77-117 NTIS PB 277 012, October 1977.
1-9. Thompson, R. E., et al., "Effectiveness of Gas Recirculation and
Staged Combustion in Reducing NOX on a 560
MW Coal-Fired Boiler," EPRI FP-257, Electric Power Research
Institute, Palo Alto, CA, September 1976.
1-10. Castaldini, et al., "Technical Assessment of Thermal DeNOx
Process," EPA-600/7-79-117, NTIS-PB 297 947, May 1979.
1-11. Martin, G. B., "Field Evaluation of Low NOX Coal Burners on
Industrial and Utility Boilers," in Proceedings of the Third
Stationary Source Combustion Symposium. Volume I.
EPA-600/7-79-050a, NTIS PB 292 539, February 1979.
1-12. Johnson, S. A., et al., "The Primary Combustion Furnace System «
An Advanced Low-N0x Concept for Pulverized Coal Combustion," in
Proceedings: Second NOV Control Technology Seminar. EPRI
FP-1109-SR, Electric Power Research Institute, Palo Alto, CA, July
1979.
1-29
-------
1-13. Barsin, J. A., "Pulverized Coal Firing NOX Control," in
Proceedings: Second N0y Control Technology Seminar. EPRI
FP-1109-SR, Electric Power Research Institute, Palo Alto, CA,
July 1979.
1-14. Unpublished data supplied by G. A. Hollinden, Tennessee Valley
Authority, Chattanooga, TN, August 1977.
1-15. Crawford, A. R., et a!., "Field Testing: Application of Combustion
Modification to Power Generating Combustion Sources," in
Proceedings of the Second Stationary Source Combustion Symposium,
Volume II, EPA-600/7-77-073b,
NTIS PB 271 756, July 1977.
1-16. Bartok, W., et al., "Combustion Modification for the Control of Air
Pollutant Emissions from Coal-Fired Utility Boilers," ASME
78-WA/APC-7, December 1978.
1-17. Selker, A. P. "Program for Reduction of NOX from Tangential
Coal-Fired Boilers, Phases II and IIA," EPA-650/2-73-005a and b,
NTIS PB 245 162 and PB 246 889, June and August 1975.
1-18. Crawford, A. R., et al., "The Effect of Combustion Modification on
Pollutants and Equipment Performance of Power Generation
Equipment," in Proceedings of the Stationary Source Combustion
Symposium. Volume III. EPA-600/2-76-152C. NTIS PB 257 146. June
1-19. McGlamery, G. G., et al., "Detailed Cost Estimates for Advanced
Effluent Desulfurization Processes," EPA-600/2-75-006, NTIS PB 242
541, January 1975.
1-20. Waitzman, D. A., et al., "Evaluation of fixed-Bed Low-Btu Coal
Gasification Systems for Retrofitting Power Plants," EPRI Report
No. 203-1, Electric Power Research Institute, Palo Alto, CA,
February 1975.
1-21. Unpublished data supplied by E. J. Campobenedetto, Babcock & Wilcox
Co., Barberton, OH, November 1978.
1-22. Vatsky, J., "Effectiveness of NOX Emission Controls on Utility
Steam Generators" Foster Wheeler Energy Corp., Livingston, NJ,
Report to Acurex Corp., Mountain View, CA, November 1978.
1-23. Electrical World. Volume 180, No. 9, pp. 39-54, November 1973.
1-24. Chemical Engineering, Volume 85, No. 11, pp 189-190, May 1978.
1-25. Electrical World. Volume 184, No. 10, pp. 43-58, November 1975.
1-30
-------
1-26. Schalit, L. M., and K. J. Wolfe, "SAM/IA: A Rapid Screening Method
for Environmental Assessment of Fossil Fuel Process Effluents,"
EPA-600/7-78-015, NTIS PB 277 088, February 1978.
1-27. Waterland, L. R., and L. B. Anderson, "Source Analysis Models for
Environmental Assessment," presented at Fourth Symposium on
Environmental Aspects of Fuel Conversion Technology, Hollywood, FL,
April 17, 1979.
1-28. Mason, H. B. et al., "Preliminary Environmental Assessment of
Combustion Modification Techniques: Volume II. Technical
Results," EPA-600/7-77-119b, NTIS-PB 276 681/AS, October 1977.
1-3U
-------
SECTION 2
INTRODUCTION
This report assesses the operational, economic, and environmental
impacts from applying combustion modification NO controls to utility
/\
and large industrial boilers. With more NO controls being implemented
^
in the field and expanded control development anticipated for the future,
there is currently a need to: (1) ensure that the current and emerging
control techniques are technically and environmentally sound, and
compatible with efficient and economical operation of systems to which
they are applied, and (2) ensure that the scope and timing of the new
control development program are adequate to allow stationary sources of
NO to comply with potential air quality standards. The NO EA
A /\
program addresses these needs by (1) identifying the incremental
multimedia environmental impact of combustion modification controls, and
(2) identifying the most cost-effective source/control combinations to
achieve ambient NOp standards.
2.1 BACKGROUND
The 1970 Clean Air Act Amendments designated oxides of nitrogen
(NO ) as one of the criteria pollutants requiring regulatory controls to
A
prevent potential widespread adverse health and welfare effects.
Accordingly, in 1971, EPA set a primary and secondary National Ambient Air
Quality Standard (NAAQS) for N02 of 100 \±q/m (annual average), To
attain and maintain the standard, the Clean Air Act mandated control of new
mobile and stationary NO sources, each of which emits approximately half
A
of the manmade NO nationwide. Emissions from light duty vehicles (the
^
most significant mobile source) were to be reduced by 90 percent to a level
of 0.25 g N02/km (0.4 g/mile) by 1976. Stationary sources were to be
regulated by EPA New Source Performance Standards (NSPS), which are set as
control technology becomes available. Additional standards required to
2-1
-------
attain air quality in the Air Quality Control Regions (AQCR's) could be set
for new or existing sources through the State Implementation Plans (SIPs).
Since the Clean Air Act, techniques have been developed and implemented
that reduce NO emissions by a moderate amount (30 to 60 percent) for a
/\
variety of source/fuel combinations. In 1971, EPA set NSPS for large steam
generators burning gas, oil, and coal (except lignite). Recently, more
stringent standards for utility boilers burning all gaseous, liquid, and
solid fuels have been promulgated. In addition, NSPS have been promulgated
for stationary gas turbines and are currently being considered for stationary
internal combustion engines and intermediate size (industrial) steam
generators. Local standards also have seen set, primarily for new and
existing large steam generators and gas turbines, as parts of State
Implementation Plans in several areas with NO problems. This regulatory
^
activity has resulted in reducing NOX emissions from stationary sources by
30 to 60 percent. The number of controlled sources is increasing as new
units are installed with factory equipped NO controls.
A
Emissions have been reduced comparably for light duty vehicles.
Although the goal of 90 percent reduction (0.25 g N02/km) by 1976 has not
been achieved, emissions were reduced by about 25 percent (1.9 g/km) for the
1974 to 1976 model years and in 1979 were reduced to 50 percent to 1.25 g/km.
Achieving the 0.25 g/km goal has been deferred indefinitely because of
technical difficulties and fuel penalties. Initially, the 1974 Energy Supply
and Environmental Coordination Act deferred compliance to 1978. Recently, the
Clean Air Act Amendments of 1977 abolished the 0.25 g/km goal and replaced it
with an emission level of 0.62 g/km (1 g/mile) for the 1981 model year and
beyond. However, the EPA Administrator is required to review the 0.25 g/km
goal, considering the cost and technical capabilities, as well as the need of
such a standard to protect public health or welfare. A report to the
Congress is due July 1980.
Because the mobile source emission regulations have been relaxed,
stationary source NO control has become more important for maintaining air
^
quality. Several air quality planning studies have evaluated the need for
stationary source NO control in the 1980's and 1990's in view of recent
/\
developments (References 2-1 through 2-9). These studies all conclude that
relaxing mobile standards, coupled with the continuing growth rate of
stationary sources, will require more stringent stationary source controls
2-2
-------
than current and impending NSPS provide. This conclusion has been
reinforced by projected increases in the use of coal in stationary sources.
The studies also conclude that the most cost-effective way to achieve these
reductions is by using combustion modification NO controls in new sources.
A
It is also possible that separate NO control requirements will be
A
needed to attain and/or maintain additional NOp related standards. Recent
data on the health effects of N02 suggest that the current NAAQS should be
supplemented by limiting short term exposure (References 2-4 and 2-10
through 2-12). In fact, the Clean Air Act Amendments of 1977 require EPA to
set a short term N0~ standard for a period not to exceed 3 hours, unless
it can be shown that such a standard is not needed. EPA will probably
propose a short term standard in 1980 when update of the N02 air quality
criteria document (Reference 2-13) is completed (References 2-14 and 2-15).
EPA is continuing to evaluate the long range need for additional
NO regulation as part of strategies to control oxidants or pollutants for
/\
which NO is a precursor, e.g., nitrates, nitrosamines, and acid rain
(References 2-4, 208, 2-10, and 2-14 through 2-18). These regulations could
be source emission controls or additional ambient air quality standards. In
either case, additional stationary source control technology could be
required to assure compliance.
In summary, since the Clean Air Act, near term trends in NOX
control are toward reducing stationary source emissions by a moderate
amount, hardware modifications in existing units or new units of
conventional design will be stressed. For the far term, air quality
projections show that more stringent controls than originally anticipated
will be needed. To meet these standards, the preferred approach is to
control new sources by using low NO redesigns.
2.2 ROLE OF UTILITY BOILERS
Utility boilers produce the largest contribution of NO emissions
A
from stationary sources in the U.S. In fact, Figure 2-1 shows that utility
boilers were the origin of 52 percent of all stationary anthropogenic NO
emissions for the year 1977 (Reference 2-19). The problem of NO emissions
A
will continue unless adequate controls are developed (Reference 2-20). The
problem will become more severe as impending shortages of oil and gas fuels
force conversion to coal, which has the potential for higher NO emissions.
In fact, the Powerplant and Industrial Fuel Use Act of 1978 (Reference 2-22)
2-3
-------
Noncombustion 1.9%
Warm air furnaces 2.0%
Gas turbines 2.0%
- Incineration 0.4%
Others 4.1%
Industrial process
heaters 4.1%
Industrial
Boilers
14.4%
Reciprocating
1C Engines
18.9%
Total: 10.5 Tg/yr (11.6 x 10b tons/yr)
Figure 2-1.
Distribution of stationary anthropogenic NOX emissions for
the year 1974 (stationary fuel combustion: controlled NOX
levels).
2-4
-------
will prohibit all new utility boilers and other major fuel burning
installations with an aggregate heat input capacity >73 MW (250 x 10 Btu/hr)
from burning oil or natural gas, except under extraordinary circumstances.
Furthermore, conversion of existing units to coal may possibly be encouraged
through tax incentives.
Given this background and their potential for N0¥ control, utility
A
boilers were chosen as the first source category to be treated under the NO
A
EA program. The "Preliminary Environmental Assessment of Combustion
Modification Techniques" (Reference 2-8) concluded that modifying combustion
process conditions is the most effective and widely used technique for
achieving 20 to 70 percent reduction in oxides of nitrogen. Nearly all
current NO control applications use combustion modifications. Other
X
approaches, such as treating postcombustion flue gas, are being evaluated
in depth elsewhere (Reference 2-23) for potential future use.
2.3 OBJECTIVE OF THIS REPORT
This report provides comprehensive, objective, and realistic
evaluations and comparisons of the important aspects of the available
combustion NO control techniques, using a common and uniform basis for
J\
comparison. The objective is to perform an environmental assessment of
NOV combustion modification techniques for utility and large industrial
/\
boilers to:
• Determine their impact on the achievement of selected
environmental goals, based on a comprehensive analysis from a
multimedia consideration
• Ascertain the effect of their application on boiler performance
and identify potential problem areas
• Estimate the economics of their operation
• Estimate the limits of control achievable by combustion
modification
• Identify further research and development and/or testing
required to optimize combustion modification techniques and to
upgrade their assessments
2.4 ORGANIZATION OF THIS REPORT
Evaluating the effectiveness and impacts of NO combustion
A
controls applied to utility and large industrial boilers requires
assessing their effects on both controlled source performance, especially
2-5
-------
as translated into changes in operating costs and energy consumption, and
on incremental emissions of other pollutants as well as NO . To perform
x\
such an evaluation, it is necessary to:
t Characterize the source category with regards to equipment and
emissions, including projected control requirements (Section 3)
t Identify current and potential NO control techniques
/\
available for implementation (Section 4)
• Identify key combustion parameters affecting NO formation by
A
correlating NO emissions with these parameters, thereby
rt
assessing the basis and effectiveness of control techniques
which modify these parameters (Section 5)
• Relate the application of preferred (major) NO controls to
A
demonstrated or expected impacts on controlled source
operations and performance (Section 6)
t Estimate the capital and operating costs, including energy
impacts of implementing NO control (Section 7)
X
• Evaluate the environmental impact of NO controls through the
/\
analysis of incremental emissions (Section 8)
Section 8 also summarizes the effectiveness of NO controls, their
y\
boiler operation/maintenance impact, and their economic impact. It
concludes with control technology and R&D recommendations.
Volume II of this report (Reference 2-24), printed under separate
cover, presents supporting data not listed in the present volume.
2-6
-------
REFERENCES FOR SECTION 2
2-1. Crenshaw, J. and Basala, A., "Analysis of Control Strategies to
Attain the National Ambient Air Quality Standard for Nitrogen
Dioxide," presented at the Washington Operation Research Council's
Third Cost Effectiveness Seminar, Gaithersburg, MD, March 1974.
2-2. "Air Quality, Noise and Health — Report of a Panel of the
Interagency Task Force on Motor Vehicle Goals Beyond 1980,"
Department of Transportation, March 1976.
2-3. McCutchen, G. D., "NOX Emission Trends and Federal Regulation,"
presented at AIChE 69th Annual Meeting, Chicago, November to
December 1976.
2-4. "Air Program Strategy for Attainment and Maintenance of Ambient Air
Quality Standards and Control of Other Pollutants," Draft Report,
U.S. EPA, Washington, D.C., October 1976.
2-5. "Annual Environmental Analysis Report, Volume 1 Technical Summary,"
The MITRE Corporation, MTR-7626, September 1977.
2-6. Personal coirmunication, Bauman, R., Strategies and Air Standards
Division, Office of Air Quality Planning and Standards, U.S. EPA,
October 1977.
2-7. "An Analysis of Alternative Motor Vehicle Emission Standards," U.S.
Department of Transportation/U.S. EPA/U.S. FEA, May 1977.
2-8. Mason, H. B., et al., "Preliminary Environmental Assessment of
Combustion Modification Techniques," EPA-600/7-77-119b, October
1977.
2-9. Greenfield, S. M., et al., "A Preliminary Evaluation of Potential
NOX Control Strategies for the Electric Power Industry," EPRI
TR-13300, April 1977.
2-10. French, J. G., "Health Effects from Exposure to Oxides of
Nitrogen," presented at the 69th Annual Meeting, AIChE, Chicago,
November 1976.
2-11. "Scientific and Technical Data Base for Criteria and Hazardous
Pollutants - 1975 EPA/RTP Review," EPA-600/1-76-023, NTIS-PB 253
942/AS, Health Effects Research Laboratory, U.S. EPA, January 1976.
2-12. Shy, C. M., "The Health Implications of an Non-Attainment Policy,
Mandated Auto Emission Standards, and a Non-Significant
Deterioration Policy," presented to Committee on Environment and
Public Works, Serial 95-H7, February 1977.
2-13. "Report on Air Quality Criteria for Nitrogen Oxides," AP-84,
Science Advisory Board, U.S. EPA, June 1976.
2-7
-------
2-14. "Report on Air Quality Criteria: General Comments and
Recommendations," Report to the U.S. EPA by the National Air
Quality Advisory Committee of the Science Advisory Board, June 1976.
2-15. "Air Quality Criteria Document for Oxides of Nitrogen; Availability
of External Review Draft," Federal Register, Vol. 43, pp. 58, 117-8,
December 12, 1978.
2-16. Personal communication, Jones, M., Strategies and Air Standards
Division, Pollutant Strategies Branch, September 1976.
2-17. "Control of Photochemical Oxidants — Technical Basis and
Implications of Recent Findings," EPA-450/2-75-005, Office of Air
and Waste Management, OAQPS, July 1975.
2-18. Waterland, L. R., et al., "Environmental Assessment of Stationary
Source NOX Control Technologies — Second Annual Report,"
EPA-600/7-79-147, June 1979.
2-19. Waterland, L. R., et al., "Environmental Assessment of Stationary
Source NOX Control Technologies — Final Report," Acurex Draft
Report, EPA Contract 68-02-2160, Acurex Corp., Mountain View, CA,
April 1980.
2-20. Mason, H. B., et al., "Utility Boiler NOX Emission
Characterization, in Proceedings: Second Annual NOX Control
Technology Seminar, EPRI FP-1109-SR, Electric Power Research
Institute, Palo Alto, CA, July 1979.
2-21. "Powerplant and Industrial Fuel Use Act of 1978," Public Law
95-620, November 9, 1978.
2-22. Proposed Rules to Implement the Powerplant and Industrial Fuel Use
Act," Federal Register, Vol. 43, pp. 53,974, November 17, 1978.
2-23. Faucett, H. L., et al., "Technical Assessment of NOX Removal
Processes for Utility Application," EPA 600/7-77-127 or EPRI
AF-568, March 1978.
2-24. Lim, K. J., et al., "Environmental Assessment of Utility Boiler
Combustion Modification NOX Controls: Volume II. Appendices,"
EPA-600/7-80-075b, April 1980.
2-8
-------
SECTION 3
SOURCE CHARACTERIZATION
This section presents a general characterization of the utility
boiler NO source category to aid in the process engineering and
A
evaluation of controls that follow in the subsequent sections of this
report. Utility boilers were categorized by equipment design and fuel fired
according to characteristics which affect the formation and/or control of
NO . The Preliminary Environmental Assessment of Combustion Modification
A
Techniques (PEAR, Reference 3-1) concluded that the number of equipment/fuel
classifications is too large to treat comprehensively at the same level of
detail. Accordingly, the PEAR completed a preliminary prioritization of
these classifications based on the quantification of source emissions and
the evaluation of the potential applications of NO control for various
A
equipment/fuel types. For example, there are a number of equipment designs
in the field which are no longer being manufactured. Since the bulk of
these sources are scheduled to be retired and have not been retrofitted with
NOX controls, they were accorded lesser priority in this study.
The PEAR (Reference 3-1) divided utility boilers into major design
types (tangential, single wall, and opposed wall) likely to be extensively
controlled for NO , and minor design types (cylcone, vertical, and stoker)
A
not likely to be extensively controlled due to dwindling use and/or lack of
control flexibility. It should be noted that minor design types are not
necessarily insignificant sources of NOX. For example, cyclone boilers
emit approximately 9 percent of stationary source NO and rank second
A
among all stationary source design/fuel classifications (Reference 3-2).
Yet, the cyclone combustion characteristics make them very difficult to
control for NOX. Their sale has been discontinued for other than high
sodium lignite applications, and it is unlikely many existing units will be
controlled for NO .
3-1
-------
Another basis used for source prioritization was fuel availability.
To date, gas- and oil-fired utility boilers have been the most extensively
controlled, but an increasing number of emissions standards have been set
recently for coal units. Few new gas- or oil-fired units are being sold, so
NO controls for coal units to meet Standards of Performance for New
A
Stationary Sources (NSPS) will dominate in the future. Consequently, this
study emphasizes coal-fired units; though NO control for gas- and
^
oil-fired units are also treated.
In the following subsections, the characteristics of the major and
minor utility boiler types are summarized with respect to: design
characteristics, fuels utilization, operational conditions, effluent
streams, and emissions. Current dominant designs and new trends are
considered. For the purposes of this study, the utility boiler category
encompasses all field erected watertube boilers with a heat input greater
than 73 MW (250 MBtu/hr) corresponding to an electrical generating capacity
of about 25 MW. For purposes of estimating emissions and evaluating the
applicability of NO controls, since large industrial boilers within the
/\
above capacity range are generally similar to the corresponding small
utility boilers, large industrial boilers can be effectively grouped with
the utility units.
3.1 COAL-FIRED BOILERS
In 1977, utility boilers consumed approximately 12 EJ of coal —
57 percent of all fossil fuels used by utility boilers (Reference 3-2).
According to the National Coal Association (NCA), coal consumption can be
expected to increase significantly (Reference 3-3). This projected rapid
increase in coal consumption is partly due to pressures on utilities by the
government to switch to coal as the primary fuel and a recognition by the
utilities themselves of the impending shortages of gas and oil.
The heavy dependence on coal will increase the environmental impact
of utility boilers on air quality. Coal is generally more polluting than
other conventional fossil fuels. The nitrogen, sulfur, and ash contents of
coal give rise to significant NO , SO^. and particulate emissions.
These emissions are almost always higher for coal than for gas or oil
combustion. In addition, trace elements in the coal account for other
3-2
-------
pollutants in the flue gases emitted to the atmosphere. Based on the
projected widespread use of coal-fired utility boilers in the 1980's and the
significant increase in NO emissions, these sources will be primary
n
candidates for NO controls.
n
The following sections characterize coal-fired utility boilers. A
brief description of each boiler design is presented in Section 3.1.1.
Current and projected coal consumption for each coal type and boiler firing
type is given in Section 3.1.2. Regional coal consumption is also presented
in this section. Section 3.1.3 describes gaseous, liquid, and solid
emission streams from coal-fired boilers. Then Section 3.1.4 presents an
overview of projected NSPS. NO emission inventories by equipment types
A
and geographical locations are also summarized in this section. Finally,
Section 3.1.5 describes pollutant control devices commonly installed on
these units.
3.1.1 Equipment Types
The major utility boiler designs are the following:
• Tangential
• Single wall
• Opposed wall (often termed horizontally opposed)
• Turbo furnace
• Cyclone
t Vertical
• Stoker
Tangential, single and opposed wall firing, and turbo furnaces are the
designs used by the four major utility boiler manufacturers, making up
approximately 87 percent of the total boiler population (Reference 3-4).
These primary design types are projected for widespread use in the 1980's.
Thus, they are candidates for application of NO controls and have been
extensively evaluated in this study. Since cyclone, vertical, and stoker
firing types are either diminishing in use or are unlikely to see widespread
use of NO controls in the near future, they are considered secondary
/\
designs. Table 3-1 describes the major design characteristics, fuel
consumption, and trends for each firing type.
The following subsections describe the major design characteristics
of each of these boiler types in more detail and preview typical NO
A
emissions from these boilers.
3-3
-------
TABLE 3-1. SUMMARY OF UTILITY AND LARGE INDUSTRIAL BOILER CHARACTERIZATION (Reference 3-4)
CO
1
-Pi
Design Type
Tangential
Single Mall
Design
Characteristics
Fuel and air nozzles
in each corner of
the combustion
chamber are directed
tangentially to a
small firing circle
in the chamber.
Resulting spin
of the flames mixes
the fuel and air in
the combustion zone.
Burners mounted
to single furnace
wall — up to
36 on single wall.
Typical Process
Values
Input Capacity:
73 MH to 3800 MW
Steam Pressure:
18.6 MPa (subcritical)
26.2 MPa (supercritical)
Steam Temperature:
75* to 840K
Furnace Volume: •,
Up to 38,000 m
Furnace Pressure
50 Pa to
1000 Pa
Furnace Heat Release:
Coal — 104 to 250
kW/m3
Oil, gas -- 208 to 518
kW/m5
Excess A1r
25X coal
10K oil
8X gas
Units typically limited
in capacity to about
400 MW (electric) because
of furnace area.
Fuel Consumption
(«)
67X coal fired
18X oil fired
15X gas fired
43X coal
22X oil fired
35X gas fired
Effluent Streams
Gaseous
Flue gas con-
taining flyash,
volatilized
trace elements.
S02, NO, other
pollutants.
Liquid
Scrubber streams,
ash sluicing
streams, wet
bottom slag
streams.
Solid
Solid ash removal
Flyash removal
Gaseous
Flue gas con-
taining flyash,
volatilized
trace elements,
SO?, NO, other
pollutants.
Liquid
Scrubber streams.
ash sluicing
streams, wet
bottom slag
streams.
Solid
Solid ash removal
Flyash removal
Operating
Modes
Soot blowing, on-
off transients,
load transients,
upsets, fuel
additives, rap-
ping, vibrating.
Soot blowing, on-
off transients,
load transients,
upsets, fuel
additives, rap-
ping, vibrating.
Effects of Transient,
Nonstandard
Operation
During startup.
NOX emissions are
low since flame
temperatures not
developed. Dur-
ing load reductions,
emissions of NOX
decrease because
of lower flame
temperatures.
NOX should de-
crease following
soot blow due to
improved heat
transfer.
During startup,
NOX emissions are
low since flame
temperatures not
developed. Dur-
ing load reductions,
emissions of NOX
decrease because
of lower flame
temperatures.
NOX should de-
crease following
soot blow due to
improved heat
transfer.
Trends
Trend toward
coal firing in
new units; con-
version to oil
and coal in
existing units.
19. 4X of current
installed units.
Trend toward
coal firing 1n
new units; wet
bottom units no
longer manufac-
tured due to
operational
problems with
low sulfur coals
and high combus-
tion tempera-
tures promoting
NO
X
59X of current
installed units.
Future
Importance
Primary
Primary
-------
TABLE 3-1. Continued
Design Type
Opposed Wall
Turbo
Furnace
Design
Characteristics
Burners are mounted
di opposite furnace
walls — up to 48
burners per wall.
Air and fuel fired
down toward furnace
bottom using burners
spaced across
opposed furnace
walls. Flame propo-
gates slowly passing
vertically to the
upper furnace.
Typical Process
Values
Units typically designed
in sizes greater than
400 MW (electric).
Units typically designed
In sizes greater than
400 MW (electric)
Fuel Consumption
(X)
32t coal
21* oil
47X gas
(includes turbo
furnace)
32X coal
21X oil
47X gas
(includes
opposed wall)
Effluent Streams
Gaseous
Flue gas con-
taining flyash,
volatilized
trace elements,
S02, NO, other
pollutants.
Liquid
Scrubber streams,
ash sluicing
streams, wet
bottom slag
streams.
Solid
Solid ash removal
Flyash removal
Gaseous
Flue gas con-
taining flyash,
volatilized
trace elements,
SO?, NO, other
pollutants.
Liquid
Scrubber streams,
ash sluicing
streams, wet
bottom slag
streams.
Solid
Solid ash removal
Flyash removal
Operating
Modes
Soot blowing, on-
off transients
load transients,
upsets, fuel
additives, rap-
ping, vibrating.
Soot blowing, on-
off transients,
load transients.
upsets, fuel
additives, rap-
ping, vibrating.
Effects of Transient,
Nonstandard
Operation
During startup,
NOX emissions are
low since flame
temperatures not
developed. Dur-
ing load reductions,
emissions of NOX
decrease because
of lower flame
temperatures .
NOX should de-
crease following
soot blow due to
improved heat
transfer.
During startup.
NOx emissions are
low since flame
temperatures not
developed. Dur-
ing load reductions,
emissions of NOX
decrease because
of lower flame
temperatures.
NOX should de-
crease following
soot blow due to
Improved heat
transfer.
Trends
Trend toward
coal firing and
conversions to
oil and coal
firing; again,
wet bottoms
being phased
out.
8.2X
of current
installed
un i ts .
Trend toward
coal firing —
(capacity in-
cluded with
opposed wall ).
;
Future
Importance
Primary
Primary
co
i
en
-------
TABLE 3-1. Concluded
I
oo
1
cr.
Design Type
Cyclone
Vertical and
Stoker
Design
Characteristics
Fuel and air intro-
duced circumferen-
tial ly into cooled
furnace to produce
swirling, high tem-
perature flame;
cyclone chamber
separate from main
furnace; cyclone
furnace must operate
at high temperatures
since it is a slag-
ging furnace.
Vertical firing re-
sults from downward
firing pattern.
Used to a limited
degree to fire
anthracite coal.
Stoker projects fuel
into the furnace
over the fire per-
mitting suspension
burning of fine
fuel particles.
Spreader stokers
are the primary
design type.
Typical Process
Values
Furnace Heat Release:
4.67 to 8.Z8 MW/m3
Furnace Heat Release:
1.1 to 1.9 MW/m2
Fuel Consumption
(X)
92% coal
4X oil
4t gas
100X coal
Effluent Streams
Gaseous
Flue gas contain-
ing flyash, vola-
tilized trace
elements, S02,
NO, and other
pollutants.
Liquid
Scrubber streams
Solid
Solid ash removal
Flyash removal
Gaseous
Flue gas contain-
ing flyash, vola-
tilized trace
elements, SO?,
NO, and other
pollutants.
Liquid
Scrubber streams
Solid
Solid ash removal
Flyash removal
Operating
Modes
Soot blowing, on-
off transients,
load transients.
upsets, fuel
additives, rap-
ping, vibrating.
Soot blowing, on-
off transients
load transients,
upsets, fuel
additives, rap-
ping, vibrating.
Effects of Transient
Nonstandard
Operation
During startup.
NOX emissions are
low since flame
temperatures not
developed. Dur-
ing load reductions,
emissions of NOX
decrease because
of lower flame
temperatures.
NOX should de-
crease following
soot blow due to
improved heat
transfer.
During startup,
NOx emissions are
low since flame
temperatures not
developed. Dur-
ing load reductions,
emissions of NOX
decrease because
of lower flame
temperatures.
NOx should de-
crease following
soot blow due to
improved heat
transfer.
Trends
Two cyclone
boilers sold
since 19/4
have not proven
adaptable to
emissions regu-
lations. Must
operate at high
temperatures re-
sulting in high
thermal NOX
fixation; also
operational
problems with
low sulfur coal.
3.3% of installed
units.
Since anthracite
usage has de-
clined, vertical
fired boilers are
no longer sold.
Design capacity
limitations and
high cost have
caused stokers
usage to diminish.
9.9X of current
installed units.
Future
report ance
Secondary
Secondary
-------
3.1.1.1 Tangential Boilers
Tangentially fired boilers are characterized by corner firing, with
arrays of burners and air nozzles located at the same elevation in each of
the four corners of the furnace. Each nozzle is directed tangentially to a
small firing circle in the center of the furnace, the actual combustion
zone. Figure 3-1 shows a typical tangential coal-fired boiler. In general,
tangentially fired boilers emit relatively lower NO than other
1\
uncontrolled boiler designs. The unique burner arrangement is a primary
cause of the lower NOX emissions.
In tangential boilers, the burners can tilt +30 degrees from their
horizontal setting. Burner tilt is used primarily as a method for
superheater steam temperature control. As the convective surfaces of the
furnace accumulate flue dust, the heat absorbed from the flue gas continues
to decrease. Burners are then tilted upwards to increase the temperature of
the flue gas entering the convective section of the boiler. When convective
tube fouling becomes severe, soot blowers are used to remove the coating on
the tubes. The sudden increase in heat absorption by the clean tubes
necessitates tilting the burners down to their original position. As the
fouling of the tubes resumes, the tilting burner cycle repeats itself.
Burner tilt also affects the level of NO emissions (as discussed in
A
Section 6). Optimum burner tilt settings depend on whether the furnace is
equipped with overfire air ports.
The twin furnace boiler is another design characteristic of
tangential boilers. Tangential boilers larger than 400 MW often include a
separate superheat and a reheat furnace. These two furnaces are identical
and physically joined side by side in a single unit. However, the flue gas
in one furnace does not interact with the gas in the other furnace, except
when both gas streams are joined at the stack.
Table 3-1 shows that tangentially fired boilers represent almost
20 percent of the entire boiler population. The majority of these units
burn coal as their primary fuel, but it is not uncommon for the furnace to
be retrofitted for oil or gas firing.
The average size of tangential coal-fired boilers investigated in
this study was 430 MW with a volumetric and surface heat release rate of
112 kW/m3 and :
from 16 to 64.
3 2
112 kW/m and 190 kW/m , respectively. The number of burners ranged
3-7
-------
-
' h I H
• rS -1
DRAWING FURNISHED THROUGH THE COURTESY OF
COMBUSTION ENGINEERING, INC.
Figure 3-1. Typical tangential fired boiler (Reference 3-5).
3-8
-------
3.1.1.2 Significant Design Changes for New Tangential Coal-Fired Boilers
The number of tangential coal-fired boilers in operation that were
designed to meet the 1971 NSPS is small compared to the number of pre-NSPS
units. Recent reports indicate that 5 to 10 such units were in operation or
scheduled to be online by the end of 1977 (References 3-6 and 3-7). If the
average unit size at an electrical output of 600 MW is assumed, this amounts
to less than 2 percent of installed conventional steam driven generating
capacity (References 3-8 and 3-9). The small number of units in operation
that were designed to meet 1971 NSPS is an indication of the length of time
required to design, fabricate, and install electric utility powerplant
components.
The tangential firing design is inherently a low NO producer. Of
A
28 pre-NSPS tangential coal-fired boilers, 23 units met NSPS under normal
operating conditions (Reference 3-7). Still, there are several significant
changes in the design of more recent tangential coal-fired units for
specifically meeting NSPS. These changes include the addition of overfire
air ports, and increased furnace height and plan area (References 3-7 and
3-10).
Overfire air (OFA) ports are included in the design of all new
tangential coal-fired boilers. The overfire air ports permit off
stoichiometric combustion by reducing the airflow to the burner zone and
adding air above the burner zone. For a normal overall operating level of
125 percent theoretical air, the burner zone theoretical air is reduced to
105 to 110 percent with the remainder of the combustion air introduced
through the OFA ports. Furnace slagging and tube wastage problems
associated with substoichiometric firing are reduced by firing with a small
amount of excess air in the burner zone and by the tangential firinq design,
which facilitates burning the fuel near the center of the furnace, away from
the furnace walls (References 3-7 and 3-10). Airflow to each fuel nozzle,
secondary air port, and overfire air port can be regulated by individual
dampers.
The furnace volumes of present designs are 15 to 20 percent larger
than in designs of the 1960's. This change was made to reduce slagging
problems associated with higher heat release ^ates (Reference 3-6). The
reduced heat release rate reduces thermal conversion of nitrogen to NO .
3-9
-------
3.1.1.3 Single and Opposed Wall Fired Boilers
Single and opposed wall fired boilers are essentially similar in
design. They only differ in the number cf furnace walls equipped with
burners ana in furnace depth. Single wall fired boilers have all burners on
the front or rear walls. The term front or rear wall fired boilers is often
used to make this distinction. Opposed wall fired boilers instead have
burners arranged on both the front and rear walls, horizontally facing each
other. Figures 3-2 and 3-3 show front wall and opposed wall fired boilers,
respectively.
A variation of the opposed wall design is the turbo furnace
manufactured exclusively by Riley Stoker. This design is unique because of
its venturi shaped cross section and directional flame burners. In the
Riley turbo coal-fired furnace, air and coal are injected downward toward
the furnace bottom below the venturi throat. According to Riley, this
furnace design will produce lower thermal NO emissions than uncontrolled
/\
conventional wall fired boilers (Reference 3-11). A schematic of a typical
coal-fired turbo furnace is shown in Figure 3-4.
Contrary to tangential boiler designs, the burners on these firing
types (except for turbo furnace designs) do not tilt. Superheater steam
temperatures are controlled by excess air level, heat input, flue gas
recirculation, and/or steam attemperation. Generally, the twin furnace
design is not found in either the single or opposed wall firing design.
Instead, division walls are occasionally installed in these boilers to
increase the heat transfer surface of the unit without enlarging the overall
size of the firebox. These walls divide the firebox from the furnace bottom
up to a distance of about 3 meters (10 feet) above the top burner level.
The flue gases from the two furnace boilers join before they enter the
convective section of the boiler. Division walls are not as popular on
current design coal-fired boilers as on gas- and oil-fired boilers because
coal ash deposits are difficult to clean from the wall surface.
Single and opposed wall fired boilers accounted for 67 percent of the
total installed utility boiler population in 1974 (see Table 3-1). However,
their combined coal consumption amounted to only 35 percent of the total
coal consumed by utility boilers in 1977 (Reference 3-2).
3-10
-------
£1 1724' 6"
DRAWING FURNISHED THROUGH THE COURTESY OF
THE POSTER WHEELER CORPORATION
Figure 3-2. Typical front wall fired boiler (Reference 3-5).
3-11
-------
223'-0"
DRAWING FURNISHED THROUGH THE COURTESY OF
THE BABCOCK AND WILCOX COMPANY
Figure 3-3. Typical opposed wall fired boiler (Reference 3-5).
3-12
-------
DRAWING FURNISHED THROUGH THE COURTESY
OF THE RILEY STOKER CORPORATION
Figure 3-4. Typical turbo furnace fired boiler (Reference 3-5)
3-13
-------
This consumption compares to 48 percent for tangential boilers. The main
reason for this difference is that single wall fired boilers are relatively
smaller in size. Single wall fired boilers are seldom greater than 400 MW
electrical output. Opposed wall boilers are often larger. Of all boilers
investigated in this study, the average size (electrical output) of the
single wall coal-fired boilers was 200 MW; the average opposed wall boiler
was 580 MW.
Inventory data on turbo fired furnaces are often reported in
conjunction with opposed wall boilers. The total number of turbo fired
furnaces currently supplying steam for the utilities is not widely known.
Therefore, for the purposes of fuel consumption and emissions inventory
discussions, turbo fired and opposed wall boilers are treated in this
section as a single equipment type.
3.1.1.4 Significant Design Changes for New Single and Opposed Wall
Coal-Fired Boilers
The number of coal fired single and opposed wall boilers (hereafter
collectively referred to as wall fired boilers) in operation that were
designed to meet 1971 NSPS is small compared to the number of pre-NSPS
units. A recent survey showed that nine such units were in operation or
scheduled to be online by the end of 1977 (Reference 3-6). This represents
an installed capacity of 5127 MW, less than 2 percent of installed
conventional steam driven generating capacity (Reference 3-8).
There are several significant changes in the design of coal-burning,
wall fired electric utility boilers for meeting NSPS. The primary changes
include new burner designs, addition of overfire air ports, improvement in
the control of air distribution to the burners, increased burner spacing,
and enlargment of the furnace plan area.
The new burner designs are of a limited turbulence design, as
discussed in Section 5. These burners control the mixing rates of coal and
air. This tends to delay combustion and thereby reduce the peak combustion
temperatures, limiting the thermal conversion of nitrogen to NO .
/\
Controlling the oxygen availability, by controlling the rate of mixing of
coal and air, in addition reduces the conversion of fuel nitrogen to NO .
A
Reduction in NO emissions of 45 to 60 percent due to burner design have
A
been indicated (References 3-12 and 3-13).
3-14
-------
Overfire air ports are included in many coal-burning, wall fired
units. The overfire air ports permit off stoichiometric combustion by
reducing the airflow to the burners and adding air above the burner zone.
Of course, effectively staging combustion in this manner raises the
potential for substoichiometric conditions to exist in the lower furnace.
While effective for NO control, substoichiometric combustion of coal
A
increases the potential for slagging of the furnace, corrosion, and
increased tube wastage. These problems have been attacked either by
reducing the degree of staging, or by introducing air at the furnace wall to
provide a local oxidizing atmosphere while retaining substoichiometric
conditions in the furnace.
The use of low NO burners and overfire air requires accurate
control of airflow to the burners and overfire air ports. The methods that
have been used to effect this are the compartmented windbox and the
perforated plate air hood. With the compartmented windbox, all burners
served by one pulverizer are served by one windbox compartment. Airflow to
the compartment is regulated. With the perforated plate air hood, airflow
can be regulated on an individual burner basis.
The burner spacing and the furnace plan area have also been increased
for wall fired NSPS units. These changes reduce the burner zone heat
release rates. The reduced heat release rate results in a lower level of
thermal conversion of air nitrogen to NO . Pre-NSPS designs generated
A
approximately 50 percent thermal NO and 50 percent fuel NO . With the
A A
reduced heat release rates, thermal NO generally accounts for only 25
A
percent of total NO generation (Reference 3-13). Although the increase
A
in furnace plan area reduces thermal NO , it was incorporated into unit
design by at least one manufacturer prior to NSPS implementation. A change
n 62
in heat input/furnace plan area from 6.6 kW/m (2.1 x 10 Btu/ft -hr)
to 5.7 kW/m2 (1.8 x 106 Btu/ft2-hr) was made to reduce the potential
for slag accumulation on the furnace walls (Reference 3-14).
The combined effects of the changes discussed above allow
manufacturers to guarantee that their coal, wall fired units will meet 1971
NSPS levels for NO emissions.
A
3-15
-------
3.1.1.5 "Minor" Design Boilers
Stokers, vertical, and cyclone units are categorized as "minor"
design types because relatively few boilers of this type are currently being
used by the utilities. The combined number of vertical and stoker boilers
in 1974 accounted for 9.9 percent of the entire utility and large industrial
boiler population. Cyclone boilers accounted for only 3.3 percent (see
Table 3-1). The combined utility population of stokers, vertical and
cyclone boilers is expected to even further decrease for the reasons
discussed below.
Stoker fired furnaces for utilities are seldom found in the field, as
past trends have been toward larger capacity boilers. Stoker sizes are
usually limited to 40 MW electric output. Stoker fired units also operate
at lower efficiency than pulverized coal units. Design capacity limitations
and high operating costs have made the stoker an uncommon utility equipment
type (Reference 3-15).
Vertical furnaces were developed for pulverized fuels before the
advent of water walled combustion chambers. They were also previously used
to a limited degree to fire anthracite coal. Anthracite is difficult to
burn in conventional boilers because of its low volatile content. The long
residence time resulting from the downward firing pattern in vertical
furnaces was effective in achieving ignition and char burnout for
anthracite. However, with the decline of anthracite as a utility fuel,
vertical furnaces are no longer sold and few are found in the field.
Cyclone furnaces were being sold as late as 1974, but because the
units have not proven adaptable for emissions control reasons, sales have
halted for all but high sodium lignite applications. These furnaces were
originally developed by B&W to burn low ash fusion temperature Illinois
coal, but they have recently been used successfully with lignite. In this
design, fuel and air are introduced circumferentially into the water-cooled
cyclone furnace to produce a high swirl, high temperature flame. The
cyclone furnace must operate at high combustion temperatures
(Reference 3-16), since it is designed to operate as a slagging furnace.
However, since high temperatures result in high thermal NO formation, the
A
cyclone furnace has lost much of its market. Figure 3-5 shows a schematic
of a typical cyclone fired boiler.
3-16
-------
SECONDARY SUPEtHEATER
OJTIET HEADER -+-f—
II I l||c:
SECONDARY SUPERHEATER '
INtfT H£ADER-f-,V£^-"r==
/-ATTEMPERATOR ,n
/ j it PRIMARY i
HEADER
!'..-'
i
i I;:
'
1
::ll
PRIMARY
SUPERHEATER Jjj
I
OUTLET
ii-l
^!!
•H^LH_J:: -
J Xi*tm*Bqf#\ |
I J^<^>^'< i~ r-7^_» , .. .
G^p^M^zgv^i j
DRAWING FURNISHED THROUGH THE COURTESY OF
THE BABCOCK & WILCOX COMPANY
Figure 3-5. Typical cyclone fired boiler (Reference 3-5)
3-17
-------
3.1.2 Coal Consumption
The data available on coal consumption in utility boilers is
summarized in this section. Since NO emissions from a utility boiler can
/\
vary significantly as a function of both coal type burned and equipment
design, data on both are discussed. Coal consumption by equipment type is
described in Section 3.1.2.1, while coal consumption by coal type is
described in Section 3.1.2.2. Regional coal consumption by the utility
industry is reviewed in Section 3.1.2.3.
3.1.2.1 Coal Consumption by Equipment Type
Table 3-2 lists the amount of coal burned for each utility equipment
type discussed in the previous section. In 1977 energy from coal reached
12"EJ (11 x 10 Btu), corresponding to 57 percent of the total fuel
consumed by all utility boilers. Coal consumption data in the utility
sector were obtained from References 3-16 through 3-27.
In spite of the environmental problems inherent in the recovery and
utilization of large quantities of coal, the trend toward increased coal use
is expected to continue. Table 3-2 also gives the projected coal
consumption for utility equipment types in the years 1985 and 2000. For the
year 2000, two energy scenarios -- high nuclear and low nuclear energy
contributions -- are used.
These energy scenarios were developed primarily from the DOE Midterm
Analysis Report and two EPRI documents (References 3-28 through 3-30). The
DOE report was used because of its recent analysis of the National Energy
Act. The two EPRI reports provide alternative energy growth scenarios. All
three studies were used because of the technical expertise, the high
visibility, and wide circulation of these results.
The growth in electric demand is high, with electric generation
capacity growing at between 5 and 6 percent per year. Coal and nuclear will
meet most of the electric demand, with synfuels, oil, and gas contributing a
small fraction.
In the low nuclear case, there is a heavy emphasis on utility coal
use, and no new nuclear capacity projected after 1985. In 2000, coal will
contribute 60 percent to total electricity generation and nuclear only
16"percent. For all sectors, this case is projected to use 41 percent more
fossil fuels than in the case of maximum conservation. This scenario
would occur if there were increased pressure to use our coal resources to
meet future energy demand and if construction of nuclear powerplants
continues to be slow.
3-18
-------
TABLE 3-2. UTILITY COAL CONSUMPTION, (EO)
Equipment Type
Tangential
Single Wall
Opposed Wall and
Turbo Furnace
Cyclone
Vertical and
Stoker
All Boilers
1977
5.8
(28)a
3.0
(14)
1.2
(5.7)
1.5
(7.2)
0.32
(1-5)
12
(57)
1985
10
(39)
2.7
(11)
4.6
(18)
1.3
(5.1)
0.27
(1.1)
19
(74)
2000
Low Nuclear
36
(52)
2.8
(4.0)
23
(33)
0.83
(1.1)
0.18
(0.26)
63
(91)
High Nuclear
23
(48)
2.4
(5.0)
15
(32)
0.83
(1.7)
0.18
(0.38)
41
(86)
aPercent of total utility fuel consumption is given in parentheses.
In the high nuclear scenario, nuclear powerplants are projected to
supply 40 percent of the utility sector's electric generation by the year
2000.
Table 3-2 shows that corner fired boilers will continue to be the
preferred coal combustion equipment. By the year 2000, tangential boilers
are predicted to produce over 50 percent of all electrical energy from
utilities by burning coal. Opposed wall fired boilers will be the second
most common coal firing equipment in the year 2000, with the remaining
boiler types decreasing coal consumption from 1977 levels.
3-19
-------
3.1.2.2 Coal Consumption by Coal Types
Coal is an extremely heterogeneous fuel whose chemical and physical
properties vary significantly between places of origin. Of these
properties, the two most routinely monitored by utilities are sulfur and ash
content. Sulfur (S) and ash (A) contents of the following coals are
considered representative of utility boiler consumption:
• Bituminous and sub-bituminous
— Interior province (high sulfur) — 2.8 percent S, 9.0 percent
A
— Eastern province (medium sulfur) ~ 2.2 percent S,
9.2 percent A
-- Western province (low sulfur) -- 1.6 percent S, 8.7 percent A
— North Dakota lignite, 0.4 percent S, 12.8 percent A
— Pennsylvania anthracite, 0.6 percent S, 11.9 percent A
The medium sulfur levels correspond to the average sulfur concentration of
coals used in U.S. utilities in 1974 (Reference 3-17).
Trace element content of individual coal samples is also highly
variable, typically varying within a single coal-producing region, and even
within a single seam (Reference 3-31). However, representative
concentration levels for coal have been determined and are listed in
Table 3-3 together with corresponding sulfur, ash, and heating value
contents.
Table 3-4 presents the trend in utility consumption of these coal
types. This table was drafted with information from Reference 3-2. The
data show that for both energy scenarios, low and high nuclear, described
above, the increase in consumption of medium and low sulfur bituminous and
sub-bituminous Western coals combined will be more significant than the high
sulfur Eastern coals. This conclusion was based on stringent sulfur oxide
regulations and economic tradeoffs — switching to low sulfur coals versus
implementation of scrubbing devices. Anthracite coal consumption is
expected to be substantially reduced.
3.1.2.3 Regional Coal Consumption
The distribution of fuel consumption for utility boilers by region is
given in Table 3-5. In compiling this table, regions were used to partition
national coal consumption geographically. This table was obtained from
3-20
-------
TABLE 3-3. PROPERTIES AND TRACE ELEMENTS OF REPRESENTATIVE U.S. COALS
(Reference 3-4)
Ash (percent)
Sulfur (percent)
Heating Value (kJ/kg)
Al (ppm)
Sb
As
Ba
Be
Bi
B
Cd
Co
Cr
Cu
Pb
Mn
Hg
Mo
N1
P
Se
V
Zn
Zr
Anthracite
Coal
11.9
0.6
30,238
—
0.1
9.3
54
2.8
0.1
1.0
0.1
84
112
70
8.3
169
0.3
9.3
47
—
0.2
12
31
45
Sub-bituminous & Bituminous
High S
9
2.8
27,912
12,
1.3
15
36
1.7
Medium S
9.2
2.2
27,912
240
1.0
114
2.9
9.1
14
40
14
53
O.J
>
8.0
22
63
2.0
33
312
72
Low S
8.7
1.6
23,260
10,200
1.1
13
30
1.5
0.8
95
2.4
7.6
12
33
12
45
0.2
6.7
19
53
1.7
28
260
60
Lignite
Coal
12.8
0.4
18,608
8,160
0.9
10
24
1.2
0.7
76
2.0
6.1
10
26
9.2
36
0.1
5.3
15
42
1.3
22
208
48
3-21
-------
TABLE 3-4. UTILITY COAL CONSUMPTION BY COAL TYPES, (EJ)
Coal Type
Medium Sulfur Bituminous
and Sub-bituminous
High Sulfur Bituminous
and Sub-bituminous
Low Sulfur Bituminous
and Sub-bituminous
Lignite
Anthracite
1977
5.3
(25)a
4.6
(22)
1.7
(8.1)
0.25
(1.2)
0.10
(0.48)
1985
8.9
(35)
6.9
(27)
2.9
(11)
0.32
(1.3)
0.088
(0.34)
2000
Low Nuclear
31
(44)
21
(31)
10
(15)
0.82
(1.2)
0.057
(0.082)
High Nuclear
20
(42)
14
(29)
6.5
(14)
0.55
(1.2)
0.057
(0.12)
aPercent of total utility fuel consumption is given in parentheses.
3-22
-------
TABLE 3-5. REGIONAL COAL CONSUMPTION BY EQUIPMENT TYPE IN 1974, (Percent) (Reference 3-32)
CO
I
CO
Equipment
Type
Tangential
Single Wall
Opposed Wall
Cyclone
Vertical
and Stoker
Total
New England
0.25
0.14
0.04
0.08
0.02
0.53
Middle Atlantic
5.4
3.1
0.89
1.7
0.36
11
E-N-Central
16
9.2
2.6
5.0
1.1
34
W-N-Central
4.5
2.6
0.73
1.4
0.29
9.4
South Atlantic
10
5.4
1.6
2.9
0.62
21
E-S-Central
6.5
4.2
1.2
2.3
0.49
15
W-S-Central
0.63
0.56
0.10
0.19
0.04
1.3
Mountain
3.3
1.9
0.54
1.0
0.22
6.9
Pacific
0.38
0.22
0.06
0.12
0.03
0.81
Total
47
27
7.8
15
3.2
100
-------
information in Reference 3-32. These regions are also used in data compiled
by the Federal Power Commission (FPC) and the Bureau of Mines. The
following sources were used to compile the regional coal consumption
estimates.
• Federal Power Commission — fuel consumption by type of fuel and
sulfur content (Reference 3-33)
• Bureau of Mines — data on domestic fossil fuel production and
end use by state (Reference 3-34)
• National Emissions Data System (NEDS) — fuel consumption by
region and end use (Reference 3-35)
• Battelle — analysis of boiler populations and fuels
(Reference 3-25)
As shown by this geographical fuel distribution, relatively little coal is
used the New England, Pacific, and West South-Central regions, where oil and
natural gas consumption prevails.
3.1.3 Utility Boiler Combustion Process and Effluent Streams
Utility boilers have several multimedia effluent streams which may be
affected by altering the combustion process to control NOX formation.
This section briefly discusses the combustion process in utility boilers and
identifies the multimedia effluent streams emitting from these units. It
also includes a listing of nonconventional operating practices which may
affect the makeup of these effluent streams. The following discussion
concentrates on coal since it is the main fuel now used in utility boilers,
and it requires more process equipment than other fuels.
Types of processes in utility boilers include fuel combustion, flue
gas cleaning, ash removal, and fireside boiler tube cleanup. Figure 3-6
gives a flow diagram for a typical pulverized coal-fueled boiler, showing
how these four processes relate to each other.
The fuel combustion process in utility boilers produces bottom or
hopper ash, combustion gases, volatilized noncombustible contaminants of the
fuel, and suspended ash entrained in the hot flue gases. Coal usually
contains between 5 and 15 percent ash and up to about 60 trace elements.*
Residual fuel oils contain less than 0.2 percent ash, but may have
*Trace defined as<1 percent by weight of coal
3-24
-------
( AIR EMISSIONS)
CO
I
ro
ui
WASTE WATER
1
1 STACK
WATER
BOILER TUBE
CLEANING
FIRESIDE
i
l
AIR EMISSION FLYASH */ SOLID WASTE ]
!-•. COLLECTION v S
SOOT Rl OWFR T AND/OR S02
_SpOT_B_LO_WER SCRUBBING Sw^r^W^
CTCflM A nrv/TPC / "Holt VNH 1 tl\ IU A
^IEAM t DEVICE __J ASH HANDUNG \
OR
AIR
FUEL »-
COMBUSTION AIR »
STEAM
GENERATING
BOILER
i
WATER +>
_J
FLUE GASES
V SYSTEM 1
, BOTTOM ASH
ASH
HANDLING
SYSTEM
*J AIR EMISSIONS ]
*
f
*\ SOLID
\
WASTES j
(WASTE WATER)
Figure 3-6. Coal-fired utility boiler combustion process flow diagram
(Reference 3-1).
-------
significant amounts of trace metallics, particularly vanadium. Natural gas
contains virtually no ash or trace element constituents.
Up to 65 percent of the ash in coal is entrained in the hot
combustion gases and either deposited on various boiler parts or carried out
of the boiler to the flyash collection system. Flue gas cleanup generally
consists of particulate removal equipment (cyclone, electrostatic
precipitator, or baghouse). Sulfur dioxide removal devices are employed on
less than 5 percent of current installations. The flyash collection
equipment usually produces a dry solid waste stream which is removed either
in the dry state or by a water sluicing stream which is diverted to an ash
settling pond. A recent analysis of powerplant data (Reference 3-22) shows
that about 80 percent of utility boilers remove ash by sluice water, and the
remaining 20 percent use dry removal.
The entrained ash deposited on furnace walls or other heat transfer
sections may reduce heat transfer efficiency and lead to severe slagging or
fouling if not removed. Soot blowing systems using steam or compressed air
are used to maintain fireside tube surfaces on a regular schedule depending
upon fuel and load. The soot blown off the boiler tubes becomes entrained
in the flue gases or settles in the superheater or economizer ash hoppers.
Coal ash which is not entrained in combustion gases either falls dry
to the furnace hopper (dry bottom) or melts and adheres to the furnace wall
and flows into a slag tank (wet bottom). Dry ash is removed by way of a
pnuematic conveyance system or by a water sluicing stream to an ash settling
pond. Superheater and economizer ash hoppers generally produce
insignificant amounts of ash compared to the furnace hopper and the flyash
collection system. Table 3-6 summarizes the effluent streams associated
with the combustion process in utility boilers.
Several periodic or nonstandard operating procedures can affect the
composition of the various effluent streams discussed above. Although
sootblowing was described above because it is so commonly used, it is also
included in the following periodic or nonstandard operations:
• Sootblowing
• Startup or shutdown transients
• Load changes
3-26
-------
TABLE 3-6. COMBUSTION RELATED EFFLUENT STREAMS FROM A UTILITY
BOILER (Reference 3-1)
Stream/Fuel
Gaseous effluent
streams
Liquid effluent
stream
Solid
Pulverized Coal
Flue gas containing
flyash, volatilized
trace elements,
S02, NO, other
pollutants
Scrubber streams
Ash sluicing
stream
Wet bottom slag
stream
Solid ash removal
Fuel Oil
Flue gas containing
volatilized trace
elements, flyash,
NO, S02, other
pollutants
Scrubber stream
Ash sluicing
stream (if any)
Solid ash removal
(if any)
Natural Gas
Flue gas con-
taining NO,
other
pollutants
None
None
3-27
-------
• Fuel additives
• Rapping or vibrating
• Flameout
• Upsets
• Equipment failure
Table 3-7 shows how often these operations take place and the
effluent streams which they may affect.
3.1.4 NO.. Emissions Inventory
X ™ " "
This section describes the contribution of coal-fired utility boilers
to total stationary source NO emissions, beginning with an estimate of
^
future NO control levels in Section 3.1.4.1. The fuel consumption data
rt
of Section 3.1.2 were then used to calculate total NOV emissions.
A
These
emissions are partitioned by equipment firing type in Section 3.1.4.2 and by
region in Section 3.1.4.3.
TABLE 3-7. EFFECT OF NONSTANDARD OPERATING PROCEDURES
ON THE EFFLUENT STREAMS FROM A DRY BOTTOM
PULVERIZED COAL-FIRED BOILER (Reference 3-5)
Procedure
Soot Blowing
Startup, Shutdown
Load Change
Fuel Additives
Rapping, Vibrating
Flameout
Upset
Equipment Failure
Frequency
3 to 4/day
12 to 50/yr
I/day
Continuous if used
3 to 4/day
l/yr
l/yr
Several /yr
Gaseous
•a
•
•
•
•
•
t
•
Liquid
•
•
•
0
•
•
Solid
•
•
•
•
•
Indicates possible affect on stream composition
3-28
-------
3.1.4.1
Estimated Future NO Control Levels
In projecting emissions, the effects of controls implementation must
be incorporated. For this reason, future emission control levels were
projected, based on estimated availability schedules of emerging near- and
far-term utility boiler NO control techniques (outlined in Section 4).
^
These projected control levels are listed in Table 3-8.
The recently promulgated standards (1979) for coal-fired boilers
break out specific coal types, specifically:
• Units firing subbituminous coal are limited to NO emissions of
215 ng/J (0.5 lb/106 Btu)
t Units firing bituminous and anthractie coals are limited to
258 ng/J (0.6 lb/106 Btu)
• Units firing coal containing greater than 25 percent North
Dakota, South Dakota, or Montana lignite in a cyclone furnace are
limited to 344 ng/J (0.8 lb/106 Btu)
The effects of dividing standards by coal type, though not included in this
report, will be factored into future studies.
TABLE 3-8. PROJECTED FUTURE NOX CONTROL LEVELS FOR UTILITY
BOILERS
Fuel
Estimated Implementation
Date
Control Level
ng/J (lb/106 Btu)
Coal
Oil
Gas
1971
(Promulgated standard)
1979
(Promulgated standard)
1983
1988
1971
(Promulgated standard)
1971
(Promulagted standard)
301 (0.7)
258 (0.6) to
215 (0.5)
129 (0.3)
86 (0.2)
129 (0.3)
86 (0.2)
3-29
-------
3.1.4.2 NO.. Emissions by Equipment Type
A ' " "™ " '
Table 3-9 lists the emissions from coal-fired units for all the
boiler design types. All entries in this table were obtained from
Reference 3-2. Projected NOX emissions for the year 1985 and 2000 were
compiled by using the projected coal consumption and the estimated NSPS
controls.
Even with the implementation of the projected NSPS, NO emissions
A
from coal-fired utility boilers in the year 2000 will increase by two-thirds
for the high nuclear scenario and more than double for the low nuclear
scenario. The contribution of tangential coal-fired boilers to the total
stationary NO emissions also increases significantly and will account for
^
one-third to one-half of the total NO emitted from stationary sources,
/\
depending on the contribution of nuclear power.
TABLE 3-9. NOX EMISSIONS FROM COAL-FIRED UTILITY BOILERS, (Gg/yr)
Equipment Type
Tangential
Single Wall
Opposed Wall
Vertical
and Stoker
Cyclone
Total
1977
1500
(25)a
1500
(25)
600
(10)
100
(1.7)
950
(16)
4600
(78)
1985
2500
(34)
1300
(18)
1600
(22)
87
(1.2)
800
(11)
6300
(87)
2000
Low Nuclear
5400
(49)
1000
(9.1)
3200
(29)
56
(0.51)
520
(4.7)
10,000
(92)
High Nuclear
3800
(44)
1100
(13)
2200
(25)
56
(0.65)
520
(6.0)
7700
(89)
aPercent of total stationary sources
3-30
-------
The trend is not the same for the minor firing design types. NO
A
emissions from these boilers decrease steadily for all scenarios. Since few
new cyclone and no new vertical or stoker boilers are expected to be
purchased by the utilities in the future, the existing units will be slowly
phased out. Their combined contribution to NO emissions from stationary
A
sources is expected to be 5 to 7 percent by the year 2000. Emissions from
turbo furnace boilers are included with emissions from horizontally oppposed
boilers.
3.1.4.3 Regional NO Emissions Inventory
A
This section presents regional NO emissions for coal-fired utility
A
boilers. Table 3-10 summarizes the percent of emissions for each of the
nine regions addressed. These inventories result from the regional coal
consumption data for 1974 presented in Reference 3-32. Over 50 percent of
all NO emissions from coal-fired utility boilers are from the East-North
A
Central and South Atlantic regions. New England contributes less than
1 percent of the total NO emissions from coal-fired units. The
West-North Central, West-South Central, and Pacific regions combined
contribute only 12 percent of the total NO emissions from coal-fired
A
utility boilers. These Western regions will be strongly affected by fuel
switching to coal since they are heavily oil and gas dominated.
3.1.5 Emission Control Devices
The emission control devices most commonly applied to the flue gas
stream of a utility boiler burning coal are particulate collectors and
SO scrubbers. However, application of scrubbing is still very limited.
A
Flue gas denitrification devices have not been installed in this country.
Their application has been limited to oil-fired boilers in Japan where NO
A
emission standards are very stringent. This section discusses particulate
and flue gas desulfurization (FGD) systems only. A discussion of flue gas
denitrification systems under study for utility boilers is presented in
Section 4.3.3 of this report. Control devices are rarely used in utility
boiler liquid and solid effluents.
3.1.5.1 Particulate Emission Controls
Particulate emssions from coal fired utility boilers are generally
controlled with centrifugal mechanical collectors or electrostatic
precipitators (ESP). Centrifugal mechanical collectors, also called
cyclones, are common on small and medium size utility boilers. Being of
3-31
-------
TABLE 3-10. DISTRIBUTION OF REGIONAL UNCONTROLLED NOX EMISSIONS FROM COAL-FIRED
UTILITY BOILERS IN 1974, (Percent) (Reference 3-32)
GO
I
OJ
ro
Equipment
Type
Tangential
Single Wall
Opposed Wallb
Cyclone
Vertical and
Stoker
Total
New
England
0.21
0.14
0.039
0.042
0.014
0.44
Middle
Atlantic
4.5
3.0
0.87
2.8
0.29
12
E-N-
Central
13.0
9.0
2.6
8.2
0.86
34
W-N-
Central
3.7
2.5
0.71
2.3
0.24
9.4
South
Atlantic
7.9
5.3
1.5
4.8
0.51
20
E-S-
Central
6.2
4.1
1.2
3.8
0.4
16
W-S-
Central
0.52
0.35
0.10
0.32
0.033
1.3
Mountain
2.7
1.8
0.52
1.7
0.18
6.9
Pacific
0.32
0.22
0.062
0.20
0.022
0.82
Total
39
26
7.6
24
2.5
100
basis
^Includes turbo furnace
-------
relatively simple design their initial cost is generally much smaller than
ESPs. However, their collection efficiency is not as high as an ESP.
Cyclone efficiencies can vary from 50 to 97 percent, based on the design of
the device and the physical characteristics of the participates in the flue
gas.
Electrostatic precipitators (ESPs) of single stage design are
commonly found on large size coal-fired boilers. The efficiencies of these
ESPs can be as high as 99+ percent. However, the initial installation cost
of these control devices can be very significant.
Several recent particulate studies (References 3-36 through 3-38)
have provided information on the particulate controls installed on utility
boilers. Twelve percent of pulverized coal-fired boilers have no collection
devices. Table 3-11 lists the combined average collection efficiency of
these devices. The data show that 35 percent of the flyash from pulverized
coal-fired boilers, 25 percent of the flyash from cyclone boilers and
50 percent of the flyash from stokers are not collected.
TABLE 3-11. AVERAGE PARTICULATE COLLECTION
FROM UTILITY BOILERS
Equipment/Fuel
All /Pulverized Coal
Cyclone/Coal
Stoker /Coal
All /Residual Oil
Percent Collection
65
75
50
25
3.1.5.2 Sulfur Oxides Emission Controls
There are many options under development for controlling SO
A
emissions from coal-fired utility boilers. These include the use of solvent
refined coal, dry limestone injection, direct firing of low sulfur coal, and
flue gas desulfurization (FGD) via wet scrubbing. The latter two strategies
are currently in active use and development. Where low sulfur coal is not
available, flue gas desulfurization units may be needed to meet existing
3-33
-------
regulations. In addition, the recent promulgated (1979) S02 NSPS
virtually require the use of F6D in new units. By 1979, about 65 units had
been installed on U.S. electric utility boilers, serving an electric
generation capacity of about 24,000 MW. Another 40 were under construction
and about 75 were planned in utility plants producing a total of over
85,000 MW for all existing and planned installations. This is out of a
total coal fueled capacity of 230,000 MW (Reference 3-39). Thus, the
application of FGD is becoming more widespread.
Although a typical FGD system is expected to reduce SO emissions
A
from a utility boiler by 80 to 95 percent, major problems remain. These
include scaling, corrosion, and mist elimination, i.e. all problems of an
operational or reliability nature (References 3-39 and 3-40).
Another area of serious concern to the utilities is the high cost of
FGD systems. The most creditable cost estimates have been completed by the
Tennessee Valley Authority (TVA) (Reference 3-41). TVA has updated its
detailed cost estimates for EPA (Reference 3-42). Representative investment
costs for an FGD system to remove 90 percent of S02 from a new 500 MW
boiler fired with 3.5 percent sulfur in the coal range from $60/kW to $85/kW.
Average annual revenue requirements range from 3.4 to 5.4 mills/kWh
(1977 costs). For perspective, a new coal-fired powerplant, operated
without FGD is estimated to cost from $400/kW to $600/kW with a total cost
of power of about 30 mills/kWh. It is evident that an FGD system would
represent a significant portion of the cost of installing and operating a
controlled plant. Thus, utilities are hesitant to apply FGD, unless
absolutely necessary.
3.2 OIL-FIRED BOILERS
Oil accounted for 20 percent of the total fossil fuel consumed by
utility boilers in 1977. This represented 34 percent of all the oil
consumed by stationary combustion sources (Reference 3-2).
Although domestic oil production peaked in 1970, the demand for oil
has continued to increase leading to an increased reliance on imported, more
expensive petroleum (Reference 3-43). As a result, economic and political
pressures have caused utilities to switch all new installations to coal
firing. Thus, according to utility boiler manufacturers, no new oil-fired
units have been purchased for the past 2 years and many previously ordered
3-34
-------
oil-fired units have been converted to coal firing during the design phase
(References 3-44 through 3-48).
Since few new oil-fired units will be coming online in the future,
and since there will be increasing impetus to switch existing units to coal
firing, the present treatment of oil-fired utility boilers has been less
comprehensive than that offered coal-fired units. Of course, many aspects
of coal-fired utility boiler source characterization also hold true for
oil-fired sources.
The following subsections describe oil-fired utility boiler equipment
characterization, highlighting the differences between oil- and coal-fired
boilers; utility boiler oil consumption; and oil-fired utility boiler
emission factors. Expected trends in fuel consumption to the year 2000 are
discussed and projected NO emissions are presented.
A
3.2.1 Typical Oil-Fired Boilers
Major types of oil-fired boilers are similar to those firing coal.
However, vertical and stoker fired boilers are not used to burn oil. Thus,
tangential, single wall, horizontally opposed, turbo and cyclone furnaces
are the only equipment firing types burning petroleum fuels.
Oil-fired boilers are more compact than coal-fired boilers of the
same heat input. The principle reason is that coal particles require longer
residence times for complete combustion in the furnace. Furthermore,
because the relatively low ash content of oil precludes slagging on the
cooling walls, oil-fired boilers can have smaller fireboxes than coal-fired
boilers.
Similarly, since the combustion gases contain less flyash, the
convective section of oil-fired boilers can be more compact, with more
closely spaced tubes. Finally, oil-fired boilers operate at lower excess
air levels than coal-fired units; up to 20 percent less air volume per unit
heat input is required for oil firing (Reference 3-49).
The more compact design of oil-fired furnaces often can cause NO
A
emissions to be as high as those from coal combustion even though the
nitrogen content of the oil is generally lower than that of coal. The lower
heat flux to furnace walls creates a higher temperature flame which causes
large quantities of thermal N0¥ to be formed. The thermal NOV
A A
contribution more than offsets the lower fuel NO contribution of the
n
cleaner oil fuels.
3-35
-------
Single wall and tangentially fired boilers consume the most fuel oil
among the design types. Each consumes about 8 percent of the total fossil
fuel consumed by utility boilers. The burners on single wall and
horizontally opposed units are usually register burners with capacities in
the 22 MW (~75 MBtu/hr) to 48 MW (~165 MBtu/hr) heat input range. Up to 72
burners can be mounted on the furnace walls. Residual oil is preheated and
injected through atomizers, usually using high pressure steam, though air
and mechanical atomizers are occasionally used (Reference 3-50). Distillate
oils are not preheated.
Cyclone furnaces represent the only minor design type burning
petroleum fuel. Oil burned in cyclone boilers accounted for only 5 percent
of all fuels burned in this boiler type in 1977 (Reference 3-2). Since
cyclone boilers are high NO emitters and fuel oil is becoming
increasingly scarce, it is expected that cyclone oil-fired boilers will be
even less prevalent in the future.
3.2.2 Oil Consumption
Table 3-12 shows the percentage of oil consumed for each firing
type in 1977 (Reference 3-2) and projects consumption up to the year 2000.
A major shift from oil-firing to coal-firing is expected to decrease oil
consumption by the year 1985. The Powerplant and Industrial Fuel Use Act of
1978 will greatly limit growth of oil consumption in utility boilers
throughout the century . No difference in oil consumption is expected
between the low nuclear and high nuclear scenarios because coal will fill
the utilities' fossil fuel demand if nuclear power generation is
restricted. Percent values for Table 3-12 were calculated using the total
fuel consumed by the utilities for both energy scenarios.
Many uncertainties make oil consumption difficult to predict. For
example, changes in import prices and supply can cause major changes in oil
consumption. The development of oil from the Outer Continental Shelf and
Alaska will have national as well as regional effects on the oil supply.
Since domestic supplies of petroleum are limited, means are being sought to
reduce liquid fuel consumption and increase its synthesis from other
sources. The technical and economic feasibility of several of these
synthesis processes remains to be demonstrated.
3-36
-------
TABLE 3-12. UTILITY OIL CONSUMPTION BY EQUIPMENT TYPE, (EJ/yr)
(Reference 3-2)
Year
Equipment
Type
Tangential
Single Wall
Opposed Mall and
Turbo Furnace
Cyclone
1977
1.6
(7.7)*
1.8
(8.6)
0.63
(3.0)
0.08
(0.36)
1985
1.4
(5.5)
1.6
(6.3)
0.58
(2.3)
0.067
(0.26)
2000
High
Nuclear
1.4
(2.9)
1.6
(3.4)
0.58
(1.2)
0.044
(0.092)
Low
Nuclear
1.4
(2.0)
1.6
(2.3)
0.58
(0.83)
0.044
(0.063)
aPercent of total fuel used by utilities is given in parentheses
Petroleum fuels, like coals, are heterogeneous fuels whose chemical
contaminants, sulfur, nitrogen, and trace metals, vary significantly among
regions. Petroleum fuels for utility boilers can be distinguished as
follows:
t Residual fuel oil
— Interior Province (high sulfur) — 2.8 percent (S)
~ Eastern Province (medium sulfur) — 2.2 percent (S)
— Western Province (low sulfur) — 1.6 percent (S)
t Distillate fuel oil -- 0.25 percent (S)
Table 3-13 presents the trend in utility petroleum consumption of
these different oil types. Medium and low sulfur oils will continue to
dominate the utility market to the year 2000, followed by high sulfur
residual and distillate oils. The percentage consumption values for these
fuels are shown in parentheses in the table.
3-37
-------
TABLE 3-13. UTILITY OIL FUEL CONSUMPTION BY TYPE,
(EJ/yr) (Reference 3-2)
Year
Oil Type
High Sulfur
Residual and Crude
Medium Sulfur
Residual and Crude
Low Sulfur
Residual and Crude
Distillate
1977
0.54
(2.6)a
1.3
(6.4)
1.7
(8.3)
0.49
(2.4)
1985
0.48
(1.9)
1.2
(4.7)
1.5
(5.9)
0.42
(1.6)
2000
High
Nuclear
0.48
(2.1)
1.2
(2.5)
1.5
(3.2)
0.42
(0.88)
Low
Nuclear
0.48
(0.69)
1.2
(1.7)
1.5
(2.2)
0.42
(0.60)
aPercent of total fuel consumed by utilities is given in parentheses.
Table 3-14 shows the regional oil consumption for 1974 by the
different boiler firing types. The South Atlantic region shows the highest
oil consumed with Middle Atlantic, Pacific, and New England regions
following close behind.
3.2.3 NO,, Emissions Inventory
A ' —"^^^^M-^-^-^^^—•!• Jim
The emissions from oil-fired utility boilers are listed in
Table 3-15. Single wall and tangential fired boilers produce more NO
/\
than all other types of oil-fired utility boilers. In 1977, NO
/\
contribution by oil-fired single wall and tangential boilers amounted to
82 percent of the total NO emissions from all oil-fired boilers. Their
A
contribution is expected to remain relatively the same throughout the
remainder of the century.
3-38
-------
co
i
CO
ID
TABLE 3-14. REGIONAL OIL CONSUMPTION BY EQUIPMENT TYPE IN 1974,
(Percent) (Reference 3-32)
Equipment
Type
Tangential
Single Wall
Opposed Wall
Cycl one
Total
New
England
5.8
6.3
2.3
0.32
15
Middle
Atlantic
10.4
12
4.2
0.59
27
E-N-
Central
2.0
2.1
0.79
0.11
5.0
W-N-
Central
0.3
0.3
0.1
~
0.7
South
Atlantic
12
13
4.6
0.65
29
E-S-
Central
0.71
0.78
0.29
0.04
1.8
W-S-
Central
1.7
1.8
0.68
0.10
4.3
Mountain
0.84
0.92
0.34
0.049
2.2
Pacific
6.0
• 6.5
2.4
0.34
15
Total
39
43
16
2.2
100
-------
TABLE 3-15.
NOX EMISSIONS FROM OIL-FIRED UTILITY
BOILERS, (Gg/yr) (Reference 3-2)
Year
Equipment
Type
Tangential
Single Wall
Opposed Wall
Turbo Furnace
Cyclone
1977
200
(3.4)a
330
(5.6)
98
(1.6)
20
(0.34)
1985
180
(2.5)
280
(3.8)
90
(1.2)
17
(0.23)
2000
High
Nuclear
180
(2.1)
280
(3.2)
90
(1.0)
11
(0.12)
Low
Nuclear
180
(1.6)
280
(2.6)
90
(0.82)
11
(0.01)
aPercent of total NOX from all utility boilers is given in
parentheses
Table 3-16 shows how NO emissions are partitioned in each of the
nine Census Bureau regions. In conjunction with regional fuel consumption
data, NO emissions for oil combustion are highest in Pacific and Eastern
A
Continental regions (New England, Middle and South Atlantic).
3.3 GAS-FIRED BOILERS
Natural gas accounted for 23 percent of the total fossil fuel
consumed by utility boilers in 1977, or 28 percent of the total gas consumed
by all stationary sources. As in the case of petroleum fuels, natural gas
will diminish as a fuel for utility steam generators as utilities will
switch to oil or coal.
The following subsections highlight the main differences between
gas-fired boilers and boilers burning other fuels. Natural gas consumption
3-40
-------
TABLE 3-16. REGIONAL UNCONTROLLED NO/ EMISSIONS FROM OIL-FIRED UTILITY BOILERS IN 1974,
(Percent) (Reference 3-32)
Equipment
Type
Tangenti al
Single Wall
Opposed Wall
Cyclone
Total
New
England
3.5
8.0
7.0
0.28
19
Middle
Atlantic
6.3
15
4.2
0.51
26
E-N-
Central
1.2
2.7
0.99
0.09
5.0
W-N-
Central
0.16
0.35
0.14
0.011
0.66
South
Atlantic
6.9
16
4.2
0.56
27
E-S-
Central
0.34
0.98
0.36
0.034
1.7
W-S-
Central
1.0
2.4
0.87
0.079
4.4
Mountain
0.51
1.2
0.43
0.045
2.2
t
Pacific
3.6
" 8.3
1.9
0.29
14
Total
24
55
20
1.9
100
basis
CO
I
-------
is discussed in Section 3.3.2 and N0x emissions from gas-fired utility
boilers are summarized in Section 3.3.3.
3.3.1 Typical Gas-Fired Boilers
Gas-fired boilers are quite similar in design to oil-fired boilers.
In fact, most gas-fired boilers were designed to fire oil as a supplementary
fuel. Those designed strictly for gas firing differ mainly in size.
Primarily gas-fired boilers are the most compact of all steam generators due
to the rapid combustion of the gaseous fuel, the low flame luminosity, and
the ash free content of natural gas. Figure 3-7 illustrates the sizes of
two utility boilers -- one coal fired and one gas fired -- with the same
heat input.
Because a constant supply of natural gas is difficult to maintain,
gas-fired generators are usually equipped to burn oil too. The oil burning
equipment allows plants to switch to petroleum fuels anytime the natural gas
supply is curtailed. Thus, these steam generators are not designed as
compactly as they could be if only natural gas were burned.
Since natural gas contains no fuel bound nitrogen, no fuel NO is
A
produced by its combustion. However, the high volumetric heat release rates
caused by small furnaces in gas-fired boilers can result in high thermal
NOV formation. Section 4 will show that uncontrolled NO emissions from
* x
gas-fired boilers can be higher at times than emissions from oil- or
coal-fired boilers.
3.3.2 Natural Gas Consumption
Table 3-17 shows past and projected future natural gas consumption
for each type of boiler. Single wall fired boilers are the most common type
of gas burning equipment, followed by horizontally opposed and tangential
boilers. In 1977, single wall fired boilers burned 50 percent of all
natural gas used by the utilities. Tangential and horizontally opposed
fired boilers consumed most of the remaining 50 percent. It is estimated
that natural gas consumption by the utilities will decrease by over
one-third from 1977 to 1985 and remain roughly constant after that.
Overall, natural gas represented almost 23 percent of all the fuel consumed
by utility boilers in 1977. By 1985, natural gas will represent only
12 percent of all fuels.
3-42
-------
b
Coal fired
Gas fired
Figure 3-7. Size comparison between coal- and gas-fired steam generators
of the same rating (Reference 3-49)
-------
TABLE 3-17. UTILITY GAS CONSUMPTION, (EJ/yr) (Reference 3-2)
Year
Equipment
Type
Tangential
Single Wall
Opposed Wall and
Turbo Furnace
Cyclone
1977
1.1
(5.3)a
2.4
(12.0)
1.2
(5.9)
0.078
(0.37)
1985
0.68
(2.6)
1.5
(5.7)
0.75
(2.9)
0.066
(0.26)
2000
High
Nuclear
0.68
(1.4)
1.5
(3.2)
0.75
(1.6)
0.042
(0.088)
Low
Nuclear
0.68
(0.98)
1.5
(2.2)
0.75
(1.1)
0.042
(0.06)
aPercent of total fuel consumed by utilities is given in
parentheses
The accuracy of these fuel consumption data depends on numerous
factors. For example, although a proposed pipeline to deliver gas from
Alaska in the mid-1980's will increase production temporarily, production
will decline rapidly after this source is exhausted unless recovery and
extensive offshore development can be pursued. Unfortunately, development
of offshore gas fields is not considered to be economical at today's
regulated prices. However, if price controls on interstate natural gas are
eliminated, impetus for further development and gas production may result.
In addition to the uncertainty concerning deregulation, technology for the
development of alternative synthetic gas is questionable. This will affect
the supply of gas since the projected shortfall in gas supplies in the
1980's will most likely have to be made up by synthetic gas, primarily from
coal.
3-44
-------
Table 3-18 lists the regional distribution of natural gas consumed by
utility boilers in 1974. Natural gas consumed in the West South Central
region accounts for over 60 percent of all gas consumed nationally. Natural
gas consumption is significant also in West North Central, Pacific, South
Atlantic, and Mountain regions.
3.3.3 N(L Emissions Inventory
^ "^
NO emissions from gas-fired utility boilers by equipment type are
^
listed in Table 3-19. In 1977 emissions from gas-fired utility boilers
accounted for 12 percent of all NO produced by the utilities. By 1985
n
emissions will account for only 6 percent of the total NO produced by the
utilities. The predicted reduction of NO emitted from all gas-fired
A
boilers is estimated due primarily to a 38 percent decrease in gas
consumption. Table 3-20 shows how NOX emissions from gas-fired boilers
are partitioned between the nine Census Bureau regions. The West South
Central region accounted for the most NO from natural gas combustion.
New England accounted for the least.
3-45
-------
TABLE 3-18. REGIONAL NATURAL GAS CONSUMPTION BY UTILITY BOILERS IN 1974,
(Percent) (Reference 3-2)
Equipment
Type
Tangential
Single Wall
Opposed Wall
Cyclone
Total
New
England
0.065
0.14
0.15
0.004
0.36
Middle
Atlantic
0.30
0.65
0.33
0.016
1.3
E-N-
Central
0.81
1.8
0.90
0.045
3.5
W-N-
Central
2.4
5.1
2.6
0.13
10
South
Atlantic
1.5
3.2
1.6
0.079
6.4
E-S-
Central
0.34
0.74
0.38
0.018
1.5
W-S-
Central
14
31
16
0.77
62
Mountain
1.4
3.0
1.5
0.073
6.0
Pacific
t
2.1
4.5
'2.3
0.11
9.0
Total
23
50
26
1.2
100
CO
I
-------
TABLE 3-19. NOX EMISSIONS FROM GAS-FIRED UTILITY BOILERS, (Gg/yr)
(Reference 3-2)
Year
Equipment
Type
Tangential
Single Wall
Opposed Wall and
Turbo Furnace
Cyclone
1977
100
(1.7)a
310
(5.2)
260
(4.4)
19
(0.32)
1985
66
(0.91)
180
(2,5)
140
(1.9)
16
(0.22)
2000
High
Nuclear
66
(0.76)
180
(2.1)
140
(1-6)
10
(0.12)
Low
Nuclear
66
(0.60)
180
(1.6)
140
(1.3)
10
(0.091)
aPercent of total NOX from utility boilers is given in
parentheses
3-47
-------
TABLE 3-20. DISTRIBUTION OF REGIONAL UNCONTROLLED N0xa EMISSIONS FROM GAS-FIRED
UTILITY BOILERS IN 1974, (Percent) (Reference 3-32)
CO
i
CO
Equipment
Type
Tangential
Single Wall
Opposed Wall
Cyclone
Total
New
Engl and
0.03
0.16
0.17
0.008
0.37
Middle
Atlantic
0.15
0.75
0.38
0.016
1.3
E-N-
Central
0.40
2.0
1.0
0.039
3.4
W-N-
Central
1.2
5.9
3.0
0.12
10
South
Atlantic
0.74
3.7
1.9
0.07
6.4
E-S-
Central
0.17
0.85
0.44
0.016
1.5
W-S-
Central
7.1
36
18
0.71
62
Mountain
0.68
3.4
1.8
0.07
6.0
1
Pacific
1.0
' 5.2
2.7
0.1
9.0
Total
12
58
30
1.2
100
aN02 basis
T-845
-------
REFERENCES FOR SECTION 3
3-1. Mason, H. B., et al., "Preliminary Environmental Assessment of
Combustion Modification Techniques," Volume II Technical Results,
EPA-600/7-77-119D, NTIS-PB 276 681, October 1977.
3-2. Waterland, L. R., et al., "Environmental Assessment of Stationary
Source NOX Control Technologies — Final Report," Acurex Draft
Report FR-80-57/EE, EPA Contract 68-02-2160, Acurex Corp.,
Mountain View, CA, April 1980.
3-3. "Energy Daily," November 1977.
3.4. Salvesen, K. G., et al., "Emissions Characterization of Stationary
NOX Sources. Volume I. Results," EPA-600/7-78-120a,
NTIS PB 284 520, June 1978.
3-5. Crawford, A. R., et al., "Field Testing: Application of Combustion
Modifications to Control NOX Emissions for Utility Boilers,"
EPA-650/2-74-066, NTIS-PB 237 344/AS, June 1974.
3-6. "NOX Control Review," Volume 2, No. 2, EPA Industrial Environmental
Research Laboratory, RTP, NC, Spring 1977.
3-7. Habelt, W. W., and Howell, B. M., "Control of NOX Formation in
Tangentially Coal-Fired Steam Generators," in Proceedings of the
NOy Control Technology Seminar, EPRI SR-39, Electric Power Research
Institute, Palo Alto, CA, February 1976.
3-8. Edison Electric Institute, "Statistical Year Book of the Electric
Utility Industry for 1976," EEI, New York, NY, October 1977.
3-9. Power, "1977 Annual Plant Design Report," November 1977.
3-10. Copeland, J. 0., and Crane, G. B., "Trip Report — Meeting with
Combustion Engineering, Inc.," Windsor, CT, February 1977.
3-11. Rawdon, A.M., and Johnson, S.A., "Control of NOX Emissions from
Power Boilers," Riley Stoker Corporation, presented at the Annual
Meeting of the Institute of Fuel, Australia, November 1974.
3-12. Campobenedetto, E. J., "The Dual Register Pulverized Coal Burner —
Field Test Results," presented at the Engineering Foundation
Conference on Clean Combustion of Coal, Rindge, NH, August 1977.
3-49
-------
3-13. Vatsky, J., "Attaining Low NOX Emissions by Combining Low Emission
Burners and Off Stoichiometric Firing," presented at 70th Annual
Meeting AIChE, New York, November 1977.
3-14. Campobenedetto, E. J., Babcock & Wilcox Co., Letter to Acurex
Corporation, November 1977.
3-15. "Steam, Its Generation and Use," Babcock & Wilcox Co., 38th Edition
1972.
3-16. Ctvrtnicek, T. E., and Rusek, S. J., "Applicability of NOX
Combustion Modifications to Cyclone Boilers (Furnaces),"
EPA-600/7-77-006, NTIS-PB 263 960, January 1977.
3-17. FPC News, Volume 8, No. 13, March 1975.
3-18. "Steam Electric Plant Factors," 1978, National Coal Association,
Washington, D.C., 1978.
3-19. Power, Plant Design Issues, 1974-1977; Vol. 118, No. 11, November
1974; Vol. 119, No. 11, November 1975; Vol. 120, No. 11, November
1976, Vol. 121, No. 11, November 1977.
3-20. Locklin, D. W., et al., "Design Trends and Operating Problems in
Combustion Modification of Industrial Boilers," EPA-650/2-74-032,
NTIS-PB 235 712, April 1974.
3-21. "End Use Energy Consumption Data Base: Series 1 Tables,"
DOE/EIA-0014, June 1978.
3-22. Surprenant, N., et al., "Preliminary Emissions Assessment of
Conventional Stationary Combustion Systems," Volume II,
EPA-600/2-76-046b, NTIS-PB 252 175, March 1976.
3-23. "Standard Support and Environmental Impact Statement for Standards of
Performance: Lignite Fired Steam Generators," EPA-450/2-76-030a,
NTIS-PB 267 610, December 1976.
3-24. "Steel Power Boilers," 1968 through 1975, U.S. Department of
Commerce, Bureau of Census; MA-34G (1968)-!, May 1969; MA-34G
(1969)-!, July 1970; MA-34G (1970)-!, August 1971; MA-34G (1971)-!,
May 1972; MA-34G (1972)-!, June 1973; MA-34G (1973)-!, May 1974;
MA-34G (1974)-!, May 1975; MA-34G (1975)-!, June 1976.
3-25. Putman, A. A., et al., "Evaluation of National Boiler Inventory,"
EPA-600/2-75-067, NTIS-PB 248 100, October 1975.
3-26. "Annual Report to Congress, 1978," U.S. Department of Energy, Energy
Information Administration, DOE/EIA-0173/2, Volume 2, July 1979.
3-27. Devitt, T., et al., "The Population and Characteristics of
Industrial/Commercial Boilers," EPA-600/7-79-178a, NTIS-PB 80-150881,
August 1979.
3-50
-------
3-28. "Energy Supply and Demand in the Midterm: 1985, 1990, and 1995,"
U.S. Department of Energy, DOE/EIA-0102/52, April 1979.
3-29. Williams, L. J., et al., "Demand 77," EPRI EA-621-SR, March 1978.
3-30. Greenfield, S. M., et al., "Preliminary Evaluation of Potential NOX
Control Strategies for the Electric Power Industry, Volume 1,"
EPRI-FP-715, March 1978.
3-31. Ctvrtnicek, T. E., "Evaluation of Low Sulfur Western Coal
Characteristics, Utilization, and Combustion Experience,"
NTIS-PB 243 911, EPA-650/2-75-046, May 1975.
3-32. Salvesen, K. G., "Emissions Characterization of Stationary NOX
Sources, Volume II: Data Supplement," EPA-600/7-78-120b,
NTIS-PB 285 429, August 1978.
3-33. "Consumption of Fuel by Electric Utilities for Production of Electric
Energy by State, Kind of Fuel and Type of Prime Mover, Year of 1974,"
FPC News Release No. 22686, October 1976.
3-34. Crump, L. H., "Fuels and Energy Data: United States by States and
Census Divisions, 1974," Bureau of Mines Information Circular 8739,
1977.
3-35. "1973 National Emission Data System (NEDS) Fuel Use Report, 1973,"
EPA-450/2-76-004, NTIS-PB 253 908, April 1976.
3-36. McKnight, J. S., "Effects of Transient Operating Conditions on
Steam-Electric Generator Emissions," EPA-600/2-75-022, NTIS-PB 247
701/AS, August 1975.
3-37. Offen, G. R., et al., "A Sunmary of Fine Particle Control by
Conventional Collection Systems", Acurex Corporation, Final Report
No. 76-216, November 1976.
3-38. Offen, G. R., et al., "Control of Particulate Matter from Oil Burners
and Boilers," EPA-450/3-76-005, NTIS-PB 258 495, April 1976.
3-39. Smith, M., et al., "EPA Utility FGD Survey, January to March 1980,"
EPA-600/7-80-029b, May 1980.
3-40. Papamarcos, J., "Stack Gas Cleanup," Power Engineering, Volume 81,
No. 6, pp. 56-64, 1977.
3-41. McGlamery, G. G., et al., "Detailed Cost Estimates for Advanced
Effluent Desulfurization Processes," EPA-600/2-75-006,
NTIS-PB 242 541, January 1975.
3-42. Ponder, W. H., Stern, R. D., and McGlamery, G. G., "S02 Control
Methods Compared," The Oil and Gas Journal. Volume 74, No. 50, pp.
60-68, December
3-51
-------
3-43. "Statistical Abstract of the United States: 1977," 98th Edition,
Washington, D.C., 1977.
3-44. Personal communication with Melosh III, H. J., Foster Wheeler
Corporation, June 1977.
3-45. Personal communication with Bouton, 6., Babcock & Wilcox, June 1977.
3-46. Personal comnunication with Devine, G., Combustion Engineering, June
1977.
3-47. Personal communication with Walsh, F. and Sadowski, R., Riley Stoker
Corporation, November 1976.
3-48. Personal communication with Barush, S., Edison Electric Institute,
December 1976.
3-49. Frendberg, A. M., "Effects of Fuel Changes on Boiler Performance,"
Babcock & Wilcox Company, presented to Pacific Coast Electric
Association Engineering and Operating Conference, March 1976.
3-50. Breen, B. P., "Combustion in Large Boilers: Design and Operating
Effect on Efficiency and Emissions," in Proceedings of the 16th
Symposium International on Combustion, Cambridge, Massachusetts,
August 1976.
3-52
-------
SECTION 4
OVERVIEW OF N0¥ CONTROL TECHNOLOGY
A. >
Modifying the combustion process conditions is the most effective
and widely used technique for achieving moderate (20 to 60 percent)
reduction in combustion generated oxides of nitrogen. This section
reviews the combustion modification techniques either demonstrated or
currently under development. The review begins with a discussion of the
formation mechanisms of NO and the general principles for suppressing
/v
NO emissions by process modifications.
4.1 GENERAL CONCEPTS ON NO FORMATION AND CONTROL
x
Oxides of nitrogen formed in combustion processes are due either to
the thermal fixation of atmospheric nitrogen in the combustion air, which
produces "thermal NO ," or to the conversion of chemically bound nitrogen
X
in the fuel, which produces "fuel NO ." For natural gas and light
A
distillate oil firing, nearly all NO emissions result from thermal
fixation. With residual oil, crude oil, and coal, the contribution from
fuel bound nitrogen can be significant and, in certain cases, predominant.
4.1.1 Thermal NO
A
During combustion, nitrogen oxides are formed by the high temperature,
thermal fixation of ^ Nitric oxide (NO) is the major product, even
though NOg is thermodynamically favored at lower temperatures. The
residence time in most stationary combustion processes is too short for
significant NO to be oxidized to NOp.
The detailed chemical mechanism for thermal NO formation is not
fully understood. However, it is widely accepted that thermal fixation in
4-1
-------
the postcombustion zone occurs according to the extended form of the
Zeldovich chain mechanism (Reference 4-1):
N2 + 0 J NO + N (4-1)
N + 02 t NO + 0 (4-2)
N + OH J NO + H (4-3)
assuming that the combustion reactions have reached equilibrium. Reaction
(4-1) has a large activation energy (317 kJ/mol) and is generally believed
to be rate determining. Oxygen atom concentrations are assumed to have
reached equilibrium according to:
02+MJO + 0 + M (4-4)
where M denotes any third substance (usually N2).
In the flame zone itself, the Zeldovich mechanism with the
equilibrium oxygen assumption is not adequate to account for experimentally
observed NO formation rates. Several investigators have observed the
production of significant amounts of "prompt" NO, which is formed very
rapidly in the flame front (References 4-2 through 4-10), but there is no
general agreement on how it is produced. Prompt NO is believed to stem from
the existence of "superequilibrium" radical concentrations (References 4-10,
4-11, and 4-12) within the flame zonewhich result from hydrocarbon chemistry
and/or nitrogen specie reactions, such as suggested by Fenimore
(Reference 4-13). To date, prompt NO has only been explicitly measured in
carefully controlled laminar flames, but the mechanism almost certainly
exists in typical combustor flames as well. Of course, in an actual
combustor, both the hydrocarbon and NO kinetics are directly coupled to
J\
turbulent mixing in the flame zone.
Recent experiments at atmospheric pressure indicate that under
certain conditions the amount of NO formed in heated N^, 0?, and Ar
mixtures can be expressed as (Reference 4-14):
[NO] = k] exp(-k2/T)[N2][02]1/2t (4-5)
4-2
-------
where [ ] = mole fraction
k,, k2 = constants
T = temperature
t = time
Although this expression certainly will not adequately describe NO formation
in a turbulent flame, it does point out several features of thermal NOX
formation. It reflects the strong dependence of NO formation on
temperature. It also shows that NO formation is directly proportional to
N~ concentration and to residence time, and proportional to the square
root of oxygen concentration.
Based on the above relations, thermal- NO can theoretically be
/\
reduced using four tactics:
• Reduce local nitrogen concentrations at peak temperature
• Reduce local oxygen concentrations at peak temperature
• Reduce the residence time at peak temperature
• Reduce peak temperature
Since reducing N~ levels is quite difficult, efforts in the field
have focused on reducing oxygen levels, peak temperatures, and time of
exposure in the NO producing regions of a furnace. On a macroscopic
J\
scale, techniques such as lowered excess air and off stoichiometric (or
staged) combustion have been used to lower local 02 concentrations in
utility boilers. Similarly, flue gas recirculation and reduced air preheat
have been used in boilers to control thermal NO by lowering peak flame
J\
temperatures. Flue gas recirculation also reduces combustion gas residence
time, but its primary effect as a thermal NO control is through
n
temperature reduction.
It is important to recognize that the above-mentioned techniques for
thermal NO reduction alter combustion conditions on a macroscopic scale.
/\
Although these macroscopic techniques have all been relatively successful in
reducing thermal NO , local microscopic combustion conditions ultimately
/\
determine the amount of thermal NO formed. These conditions are in turn
A
intimately related to such variables as local combustion intensity, heat
removal rates, and internal mixing effects. Modifying these secondary
combustion variables at microscopic levels requires fundamental changes in
combustion equipment design.
4-3
-------
For example, recent studies on the formation of thermal NO in
A
gaseous flames have confirmed that internal mixing can have large effects on
the total amount of NO formed (References 4-15, 4-16). Burner swirl,
combustion air velocity, fuel injection angle and velocity, quarl angle, and
confinement ratio all affect the mixing between fuel, combustion air, and
recirculated products. Mixing, in turn, alters the local temperatures and
specie concentrations which control the rate of NOV formation.
A
Unfortunately, generalizing these effects is difficult, because the
interactions are complex. Increasing swirl, for example, may both increase
entrainment of cooled combustion products (hence lowering peak temperatures)
and increase fuel/air mixing (raising local combustion intensity). The net
effect of increasing swirl can be to either raise or lower NO emissions,
A
depending on other system parameters.
In summary, a hierarchy of effects depicted in Table 4-1 produces
local combustion conditions which promote thermal NO formation. Although
A
combustion modificiation technology seeks to affect the fundamental
parameters of combustion, modifications must be made by changing the primary
equipment and fuel parameters. Control of thermal NO , which began by
A
altering inlet conditions and external mass addition, has moved to more
fundamental changes in combustion equipment design.
4.1.2 Fuel NOV
/\
The role of fuel bound nitrogen as a source of NO emissions from
A
combustion sources has been recognized since 1968 (Reference 4-17).
Although the relative contribution of fuel and thermal NO to total NO
X X
emissions from sources firing nitrogen containing fuels has not been
definitively established, recent estimates indicate that fuel NO is
A
significant and may even predominate. In one laboratory study
(Reference 4-18), residual oil and pulverized coal were burned in an
argon/oxygen mixture to eliminate thermal NO effects. Results show that
A
fuel NO can account for over 50 percent of total NO production from
A X
residual oil firing and approximately 80 percent of total NO from coal
A
firing. Tests on a full scale system, a 560 MW coal-fired utility boiler,
confirm this prediction (Reference 4-19). Flue gas recirculation, which
controls primarily thermal NOX, was a relatively ineffective NO control
measure for the coal-fired boiler tested.
4-4
-------
TABLE 4-1. FACTORS CONTROLLING THE FORMATION OF THERMAL NOX
Primary Equipment
and Fuel Parameters
Secondary
Combustion Parameters
Fundamental
Parameters
Inlet temperature,
velocity
Firebox design
Fuel composition
Injection pattern
of fuel and air
Size of droplets
or particles
Burner swirl
External mass
addition
\
Combustion intensity
Heat removal rate
Mixing of combustion
products into flame
Local fuel/air ratio
Turbulent distortion
of flame zone
Oxygen level
Peak temp.
Exposure time
at peak temp.
Thermal
NO
Fuel bound nitrogen occurs in coal and petroleum fuels. However, the
nitrogen containing compounds in petroleum tend to concentrate in the heavy
resin and asphalt fractions upon distillation (Reference 4-20). Therefore
fuel NO is of importance primarily in residual oil and coal firing. The
A
nitrogen compounds found in petroleum include pyrroles, indoles,
isoquinolines, acridines, and porphyrins. Although the structure of coal
has not been defined with certainty, it is believed that coal-bound nitrogen
also occurs in aromatic ring structures such as pyridine, picoline,
quinoline, and nicotine (Reference 4-20).
The nitrogen content of residual oil varies from 0.1 to 0.5 percent.
Nitrogen content of most U.S. coals lies in the 0.5 to 2 percent range
(Reference 4-21); anthracite coals contain the least and bituminous coals
the most nitrogen. Figure 4-1 illustrates the nitrogen content of various
U.S. coals, expressed at ng N02 produced per joule for 100 percent
conversion of the fuel nitrogen (Reference 4-22). The figure clearly shows
that if all coal bound nitrogen were converted to NOX, emissions for all
4-5
-------
3000
e
o
'£ 2000
Ol
§ 4
O
O
^^
3
3
CO
"~ «3
2 2
**v,^
CM
O
Z
_o
•" 1
0
D«U From:
U.S 8 M. T«ch. Htpu. (1922-1MM
_^ Avef.nl1»491 • An*r«etw
V»ndavMf (19651. p. 45 O Bituminovt
Mark •! •! (1S64I *«•._•
i»i«f». vi «i. i IOTI*I ^ Sub'BttummoM
Perry 1 1963) _ ' .
Parks ft O'Oonnctl (18SA)
—
CX o ° °
— ' J?U ^ ^.^
^k rT D Q _ I
*» rv u f*
-------
coals would exceed even the 1971 Standards of Performance for Large Steam
Generators (NSPS). Fortunately, only a fraction of the fuel nitrogen is
converted to NO for both oil and coal firing, as shown in Figure 4-2
A
(Reference 4-23). Furthermore, the figure indicates that fuel nitrogen
conversion decreases as nitrogen content increases. Thus, although fuel
NO emissions undoubtedly increase with increasing fuel nitrogen content,
/\
the emissions increase is not proportional. In fact, recent data indicate
only a small increase in NO emissions as fuel nitrogen increases
A
(Reference 4-24). From observations such as these, the effectiveness of
partial fuel denitrification as a NO control method seems doubtful.
A
Although the precise mechanism by which fuel nitrogen is converted to
NO is not understood, certain aspects are clear, particularly for coal
^
combustion. In a large pulverized coal-fired utility boiler, the coal
particles are conveyed by an airstream into the hot combustion chamber,
where they are heated at a rate in excess of 10 K/s. Almost immediately
volatile species, containing some of the coal bound nitrogen, vaporize and
burn homogeneously, rapidly (~10 ms) and probably detached from the original
coal particle. Combustion of the remaining solid char is heterogeneous and
much slower (~300 ms).
Figure 4-3 summarizes what may happen to fuel nitrogen during this
process (Reference 4-25). In general, nitrogen evolution parallels
evolution of the total volatiles, except during the initial 10 to 15 percent
volatilization in which little nitrogen is released (Reference 4-26). Both
total mass volatilized and total nitrogen volatilized increase with higher
pyrolysis temperature; the nitrogen volatilization increases more rapidly
than that of the total mass. Total mass volatilized appears to be a
stronger function of coal composition than total nitrogen volatilized
(Reference 4-27). This supports the relatively small dependence of fuel
N0x on coal composition observed in small scale testing (References 4-18
and 4-28).
Although there is not absolute agreement on how the volatiles
separate into species, it appears that about half the total volatiles and 85
percent of the nitrogeneous species evolved react to form other reduced
species before being oxidized. Prior to oxidation, the devolatilized
nitrogen may be converted to a small number of common, reduced
4-7
-------
- 5
00
X
o
z:
f^
° 100-
c
o
W1
1 80-
c
o
o
« 60'
£
*4—
M 40
20
r ,.. .flipVnn, ILiqnite coal 1 1 BUumi
Calif i c «.- 1 1 1 i — j _i
i/ iSuh-hi tumi nous coal
Coal
0 Pereira, et al. (1974)
.£} ^Pershing, et al . (1973)
n A McCann (1970)
* Q O ^ Jonke (1970)
V ^ • Bituminous M.I.T. (1975)
»££)£>. ^ Lignite M.I.T. (1975)
^^^ i^^ ^^
tf j »\ 1*1
£°|a
0 0.2 0.4 0.6 0.8 1.0 1.2
nous coal ,
I Residual oil ^
fupl oi 1 #4 A R fupl ni 1 _
Data Sources Oil
^ Flagan & Appleton (1974)
A Hazard (1973)
Q Turner & Siegmund (1972
O Fenimore (1972)
O Turner et al. (1972)
V Martin & Berkau (1972)
I> Pershing, et al. (1973)
<«J Martin, et al. (1970)
^^
w ^
^k ^^L
V ^
1.4 1.6 1.8 2.0
Weight % N in fuel
Figure 4-2. Conversion of fuel N in practical combustors (Reference 4-22).
-------
Volatile fractions
(Hydrocarbons. RN etc
Ash
virtually
nitrogen
free
Figure 4-3.
Possible fate of fuel nitrogen contained in coal
particles during combustion (Reference 4-25).
4-9
-------
intermediates, such as HCN and NH-, in the fuel-rich regions of the flames.
The existence of a set of common reduced intermediates would explain the
observations that the form of the original fuel nitrogen compound does not
influence its conversion to NO (e.g., References 4-20, 4-29). More recent
experiments suggest that HCN is the predominant reduced intermediate
(Reference 4-30). The reduced intermediates are then either oxidized to NO,
or converted to N~ in the postcombustion zone. Although the mechanism for
these conversions is not presently known, one proposed mechanism postulates
a role for NCO (Reference 4-31).
Nitrogen retained in the char may also be oxidized to NO, or reduced
to N2 through heterogeneous reactions occurring in the postcombustion
zone. However, it is clear that the conversion of char nitrogen to NO
proceeds much more slowly than the conversion of devolatilized nitrogen. In
fact, based on a combination of experimental and empirical modeling studies,
it is now believed that 60 to 80 percent of the fuel NO results from
/\
volatile nitrogen oxidation (References 4-26, 4-32). Conversion of the char
nitrogen to NO is in general lower, by factors of two to three, than
conversion of total coal nitrogen (Reference 4-29).
Regardless of the precise mechanism of fuel NO formation, several
/\
general trends are evident, particularly for coal combustion. As expected,
fuel nitrogen conversion to NO is highly dependent on the fuel/air ratio for
the range existing in typical combustion equipment, as shown in Figure 4-4.
Oxidation of the char nitrogen is relatively insensitive to fuel/air
changes, but volatile NO formation is strongly affected by fuel/air ratio
changes.
In contrast to thermal NO , fuel NO production is relatively
A X
insensitive to small changes in combustion zone temperature (Reference 4-29).
Char nitrogen oxidation appears to be a very weak function of temperature,
and although the amount of nitrogen volatiles appears to increase as
temperature increases, this is believed to be partially offset by a decrease
in percentage conversion. Furthermore, operating restrictions severely
limit the magnitude of actual temperature changes attainable in current
systems.
As described above, fuel NO emissions are a strong function of
A
fuel/air mixing. In general, any change which increases the mixing between
the fuel and air during coal devolatilization will dramatically increase
4-10
-------
OJ
X
o
c
0)
C7l
o
o
V.
c
o
u
100-
90
80
70
60-
SO
40'
30-
20-
101
0-
Wall temp 1500 K
Flame temp 1600 K
A Lignite 75-90 tim
A Lignite 38-45 nm
Q Bituminous 75-90 um
• Bituminous 38-45 nm
A
0
234
Fuel equivalence ratio
(Inverse of stoichiometric ratio)
Figure 4-4. Conversion of nitrogen in coal to NO (Reference 4-23).
-------
volatile nitrogen conversion and increase fuel NO . In contrast, char NO
A
formation is only weakly dependent on initial mixing.
From the above modifications, it appears that, in principle, the best
strategy for fuel NO abatement combines low excess air (LEA) firing,
A
optimum burner design, and two stage combustion. Assuming suitable stage
separation, low excess air may have little effect on fuel N0x, but it
increases system efficiency. Before using LEA firing, the need to get good
carbon burnout and low CO emissions must be considered.
Optimum burner design ensures locally fuel-rich conditions during
devolatilization, which promotes reduction of devolatilized nitrogen to
Np. Two-stage combustion produces overall fuel-rich conditions during the
first 1 to 2 seconds and promotes the reduction of NO to N2 through
reburning reactions. High secondary air preheat may also be desirable,
because it promotes more complete nitrogen devolatilization in the fuel-rich
initial combustion stage. This leaves less char nitrogen to be subsequently
oxidized in the fuel-lean second stage. Unfortunately, it also tends to
favor thermal NO formation, and at present there is no general agreement on
which effect dominates.
4.1.3 Summary of Process Modification Concepts
In summary of the above discussion, both thermal and fuel NO are
A
kinetically or aerodynamically limited in that their emission rates are far
below the levels which would prevail at equilibrium. Thus, the rate of
formation of both thermal and fuel NO is dominated by combustion
A
conditions and is amenable to suppression through combustion process
modifications. Although the mechanisms are different, both thermal and fuel
NO are promoted by rapid mixing of oxygen with the fuel. Additionally,
A
thermal NOX is greatly increased by long residence time at high
temperature. The modified combustion conditions and control concepts which
have been tried or suggested to combat the formation mechanisms are as
follows:
t Decrease primary flame zone Op level by
— Decreased overall 0« level
— Controlled mixing of fuel and air
-- Use of fuel-rich primary flame zone
4-12
-------
• Decrease time of exposure at high temperature by
— Decreased peak temperature:
— Decreased adiabatic flame temperature through dilution
-- Decreased combustion intensity
-- Increased flame cooling
-- Controlled mixing of fuel and air or use of fuel-rich
primary flame zone
— Decreased primary flame zone residence time
• Chemically reduce NO in postflame region by
^
~ Injection of reducing agent (e.g., MM.,)
Table 4-2 relates these control concepts to combustion process
modifications applicable to utility boilers. The process modifications are
categorized according to their role in the control development sequence:
operational adjustments, hardware modifications of existing equipment or
through factory installed controls, and major redesigns of new equipment.
The controls for decreased 02 are also generally effective for peak
temperature reduction but have not been repeated. The following subsections
briefly review the status of each of the applicable control techniques
applied to utility boilers.
4.2 STATE-OF-THE-ART CONTROLS
Based on the general principles discussed above for the suppression
of NO emissions by process modifications, there are several control
A
techniques that may be used singly or conjunctively on utility boilers.
These techniques include low excess air firing, biased burner firing,
burners out of service, overfire air, low NO burners, flue gas
A
recirculation, and reduced firing rate. These methods for controlling NOX
may be used on existing boilers although modifications to the units may be
required. Tables 4-3 through 4-10 give the average NO reduction
A
achievable with the various control techniques, compiled from the data base
of test results and test selection procedures discussed in Section 5. It
should be noted that the data base is not complete in that only those tests
that were well characterized are included; i.e., such boiler design and
operating variables as number of burners, burner stoichiometry, direct input
per active burner, surface heat release rate, etc., were reported (see
Section 5.2.2).
4-13
-------
TABLE 4-2. SUMMARY OF COMBUSTION PROCESS MODIFICATION CONCEPTS
Combustion
Conditions
Decrease
primary
flame zone
0? level
(
Decrease
peak
flame
temperature
Chemically
reduce NOX
in post
flame region
Control
Concept
Decrease overall
02 level
Delayed mixing
of fuel and air
Primary fuel-
rich flame
zone
Decrease
adiabatic flame
temperature
Decrease
combustion
intensity
Increased flame
zone cooling/
reduce residence
time
Inject reducing
agent
Effect on
Thermal NOX
Reduces 0 rich,
high NO pockets
in the flame
Flame cooling and
dilution during
delayed mixing
reduces peak
temperature
Flame cooling in
low 0 , low
temperature primary
zone reduces peak
temperature
Direct suppression
of thermal NOX
mechanism
Increased flame
zone cooling yields
lower peak
temperature
Increased flame
zone cooling yields
lower peak
temperature
Decomposition
Effect on
Fuel NOX
Reduces exposure
of fuel nitrogen
intermediaries
to 02
Volatile fuel N
reduces to N? in
the absence of
oxygen
Volatile fuel N
reduces to N2 in
the absence of
oxygen
Ineffective
Minor direct
effect; indirect
effect on mixing
Ineffective
Decomposition
Primary Applicable Controls
Operational
Adjustments
Low excess air
firing
Burner
adjustments
Burners out of
service; biased
burner firing
Reduced air
preheat
Load reduction
Burner tilt
Hardware
Modification
Flue gas recirculation
(FOR)
Low NOX burners
Overfire air ports
Water injection, F6R
Ammonia injection
possible on some units
Major
Redesign
Optimum burner/
firebox design
Burner/firebox
design for two
stage combustion
Enlarged firebox,
increased burner
spacing
Redesign heat
transfer surfaces,
firebox
aerodynamics
Redesign convective
section for NH3
injection
-------
TABLE 4-3. AVERAGE N0x REDUCTION WITH LOW EXCESS AIR FIRING (LEA)
Equipment
Type
Tangential
Opposed Wall
Single
Wall
All Boilers
Fuel
Coal
Oil
Nat
Gas
Coal
Oil
Nat
Gas
Coal
Oil
Nat
Gas
Coal
Oil
Nat
Gas
All
Fuels
Number
of
Boilers
Tested
11
-
1
5
4
6
7
(2)
4
3
(1)
23
8
10
41
Baseline
Stoichionetry
to Active Burners
(percent)
124
~
117
126
120
115
123
(134)b
120
117
(124)
124
120
116
120
HOX Emissions
(ppm dry 9 3» 02)
459
--
340
746
357
717
624
(1338)
409
418
(992)
609
383
492
495
Low Excess Air (LEA)
Stoichiometry
to Active Burners
(percent)
116
-
113
118
113
110
114
(118)
112
108
(112)
116
115
110
114
NO Emissions
(ppm dry * 31 02)
373
-
245
660
290
600
522
(1325)
315
356
(931)
522
302
400
408
Average
NO. Reduction
(percent)
19
-
28
12
19
16
16
(1)
23
15
(6)
16
21
20
19
Maximum NOX
Reduction
Reported
(percent)
42
-
28
23
30
33
25
(3)
26
15
(6)
30
28
25
28
en
'Boiler load at or above 80 percent MCR. For individual tests, corresponding baseline and controlled loads were nearly identical.
lumbers in parentheses refer to boilers originally designed for coal firing with wet bottom furnaces.
-------
TABLE 4-4. AVERAGE NO REDUCTION WITH BURNER OUT OF SERVICE (BOOS)'
Equipment
Type
Tangential
Opposed Wall
Single
Wall
All
Boilers
Fuel
Coal
Oil
Nat
Gas
Coal
Oil
Nat
Gas
Coal
Oil
Nat
Gas
Coal
Oil
Nat
Gas
All
Fuels
Number
of
Boilers
Tested
7
--
1
4
1
4
8
(1)
3
3
19
4
8
31
Baseline
No. of
Burners
Firing
32 u
(16-56)b
--
8
40 u
(24-54)b
24
26b
(16-36 )b
16
(24 )C
16 .
(12-24)"
16
(12-16)b
28
(16-56)b
20
( 12-24 )b
16 K
(8-36)b
20
(8-56)b
Stoichiometry
to Active
Burners
(percent)
121
—
112
122
107
115
123
(134)C
119
117
122
113
115
117
NOX
Emissions
(ppm dry
9 sx o2)
462
--
146
670
442
674
618
(1196)°
425
418
583
433
412
4/6
Burners Out of Service (BOOS)
Percent
Burners
on Air
Only
17
--
NA
16
33
28
19
(33)C
18
22
17
25
25
22
Stoichiometry
to Active
Burners
(percent)
98
--
86
102
73
84
97
(89)C
95
89
99
84
86
90
NOX
Emissions
(ppm dry
9 3X 02)
293
--
146
522
292
290
412
(577)C
256
214
409
274
217
300
Average NOX
Reduction
(percent)
37
--
0
22
34
57
33
(52)C
40
49
31
37
35
34
Maximum NOX
Reduction
Reported
(percent)
56
-
0
46
34
61
48
(52)C
48
69
50
41
43
45
jjBoiler load at or above 80 percent MCR. For individual tests, corresponding baseline and controlled loads were nearly identical.
"Range in number of burners firing
lumbers in parentheses refer to boilers originally designed for coal firing with wet bottom furnaces.
-------
TABLE 4-5. AVERAGE N0¥ REDUCTION WITH OVERFIRE AIR (OFA)'
-p.
I
Equipment
Type
Tangential
Opposed Wall
Single
Wall
Fuel
Coal
Oil
Nat
Gas
Coal
Oil
Nat
Gas
Coal
Oil
Nat
Gas
Number
of
Boilers
Tested
6
-
—
-
S
2
-
-
—
Baseline
Stoichiometry
to Active
Burners
(percent)
129
-
—
—
118
114
~
—
--
NOX
Emissions
(ppm dry
33X02)
454
—
-
—
376
928
—
-
—
Overfire Air (OFA)
Stoichioroetry
to Active
Burners
(percent)
105
-
-
-
96
99
—
-
~
Furnace
Stoi chl erne try
(percent)
122
~
~
~
118
112
-
-
-
NOX Emissions
(ppm dry
9 3X02)
311
~
--
--
287
378
--
--
-
Average NOX
Reduction
(percent)
31
-
--
-
24
59
—
-
~
Maximum NOX
Reduction
Reported
(percent)
41
--
--
--
30
66
-
--
-
aBo11er load at or above 80 percent NCR. For individual tests, corresponding baseline and controlled loads were nearly identical.
-------
TABLE 4-6. AVERAGE N0x REDUCTION WITH FLUE GAS RECIRCULATION (FGR)'
Equipment
Type
Tangential
Opposed Wall
Single
Wall
Fuel
Coal
Oil
Nat
Gas
Coal
Oil
Nat
Gas
Coal
Oil
Nat
Gas
Number
of
Boilers
Tested
-
-
1
1
1
--
--
-
1
Baseline
Stoichiometry
to Active
Burners
(percent)
--
-
117
128
122
-
--
-
106
NOX
Emissions
(ppm dry
3 3* 02)
--
--
340
855
304
-
--
-
470
Overfire Air (OFA)
Stoichiometry
to Active
Burners
(percent)
--
-
115
127
126
-
--
-
107
FGR
( percent )
--
-
23
15
11
--
--
--
11
NOX Emissions
(ppm dry
S 3X 02)
-
-
135
735
263
—
--
-
307
Average NOX
Reduction
(percent)
—
—
60
17
13
—
—
--
35
Maximum NOX
Reduction
Reported
(percent)
-
--
60
17
13
—
—
—
35
oo
aBoiler load at or above 80 percent MCR. For individual tests, corresponding baseline and controlled loads were nearly identical.
-------
TABLE 4-7. AVERAGE N0x REDUCTION WITH REDUCED FIRING RATE
Equipment
Type
Tangential
Opposed Hall
Single
Wall
All
Boilers
Fuel
Coal
Oil
Nat
Gas
Coal
Oil
Nat
Gas
Coal
Oil
Nat
ias
Coal
011
Nat
Gas
All
uels
Nuaber
of
Boilers
Tested
7
-
1
4
4
5
2
(2)
3
2
(1)
13
7
8
28
Baseline (SOX HCR or Above)
Firing
Rate
(percent
MCR)
93
—
100
93
98
98
92
(90)
98
97
(98)
93
98
99
97
Stolchiometry
to Active
Burners
(percent)
112
—
117
131
118
115
125
(133)»
119
118
(115)
126
119
117
120
*>x
Emissions
(ppn dry
8 3*02)
462
~
340
825
362
651
651
(1338)
425
442
(992)
646
393
478
506
Reduced Load
Firing
Rate
(percent
MCR)
64
—
75
70
61
57
67
(54)
53
35
(59)
67
57
55
60
Stoichiooetry
to Active
Burners
(percent)
127
—
135
136
121
115
130
(138)
119
117
(131)
131
120
122
124
NOX
Emissions
(ppm dry
« 3X 0Z)
408
-
332
758
249
269
496
(990)
296
125
(522)
554
272
242
356
Average NOX
Reduction
(percent)
12
--
2
8
31
59
24
(26)
30
72
(47)
14
31
44
30
Maximum NOX
Reduction
Reported
(percent)
25
--
32
18
48
64
25
(33)
45
82
(47)
23
47
59
43
aNumbers In parentheses refer to boilers originally designed for coal firing with wet bottom furnaces.
-------
TABLE 4-8. AVERAGE NOX REDUCTION WITH OFF STOICHIOMETRIC COMBUSTION
AND FLUE GAS RECIRCULATION (OSC AND FGR)a
Equipment
Type
Tangential
Opposed Wall
Single
Wall
Fuel
Coal
Oil
Nat
Gas
Coal
Oil
Nat
Gas
Coal
Oil
Nat
Gas
Number
of
Boilers
Tested
-
-
1
1
1
-
-
2
1
Baseline
Stolen ioroetry
to Active
Burners
(percent)
—
—
117
128
122
~
~
118
106
NO,
Emissions
(ppm dry
9 3X 02)
-
—
340
781
304
--
-
355
470
OSC and FGR
Type of
OSC
-
-
BOOS
BOOS
OFA
—
-
BBF
BOOS
BOOS
Stoichiometry
to Active
Burners
(percent)
-
-
75
99
97
--
--
91
75
FGR
(percent)
--
—
21
19
11
—
—
14
12
NOX
Emissions
(ppm dry
S 3X 02)
-
--
105
453
247
-
-
154
115
Average NOX
Reduction
(percent)
—
—
69
42
19
-
—
57
76
Maximum NOX
Reduction
Reported
(percent)
-
--
69
42
19
—
--
59
76
-p*
o
aBoiler load at or above 80 percent MCR. For individual tests, corresponding baseline and controlled loads were nearly identical.
-------
TABLE 4-9. AVERAGE NOX REDUCTION WITH REDUCED FIRING RATE AND OFF STOICHIOMETRIC COMBUSTION
Equipment
Type
Tangential
Horizontally
Opposed Wall
Single
Wall
All
toilers
Fuels
Fuel
Coal
Oil
Nat
Gas
Coal
011
Nat
Gas
Coal
Oil
Nat
Gas
Coal
Oil
Nat
Gas
All
Number
of
Boilers
Tested
8
-
-
3
4
6
4
(2)
3
2
(1)
15
7
3
30
Baseline
Firing
Rate
(percent
NCR)
93
—
-
93
99
100
90
98
97
(98)
92
99
99
97
Stoichiometry
to Active
Burners
(percent)
122
-
--
129
118
115
124
(133)«
120
118
(125)
125
119
117
120
NOX
Emissions
(ppm dry
9 3t 02)
453
--
-
820
362
717
663
(1338)
426
442
(992)
645
394
579
539
Low Load and OSC
Firing
Rate
(percent
MCR)
61
-
—
73
64
58
73
(59)
56
35
(H)
69
60
31
1)3
Type of
OSC
BOOS
OFA
-
-
BOOS
BOOS
OFA
BOOS
OFA
BOOS
BBF
BOOS
BBF
BOOS
BOOS
OFA
BBF
BOOS
OFA
BOOS
OFA
BBF
BOOS
OFA
Stolen iometry
to Active
Burners
(percent)
95
—
—
102
117
88
99
(91)
97
93
(102)
99
107
91
99
NOX
Emissions
(ppn dry
9 31 02)
248
—
—
634
177
148
381
(887)
228
78
(641)
421
202
113
245
Average NOX
Reduction
(percent)
45
-
--
23
51
79
43
(34)
46
82
(35)
37
49
80
b5
Maximum
NOX
Reduction
Reported
(percent)
62
--
32
67
89 j
50
(55)
59
87
(35) !
46
!
63
i
88
66 j
!
ro
aNumbers in parentheses refer to boilers originally designed for coal firing with wet bottom furnaces.
-------
TABLE 4-10. AVERAGE NOX REDUCTION WITH LOAD REDUCTION, OFF STOICHIOMETRIC COMBUSTION
AND FLUE GAS RECIRCULATION
Equipment
Type
Tangential
Opposed Wall
Single
Wall
All
Boilers
Fuel
Coal
Oil
Nat
Gas
Coal
Oil
Nat
Gas
Coal
Oil
Nat
Gas
Coal
Oil
Nat
Gas
All
Fuels
Number
of
Boilers
Tested
-
-
-
-
3
2
-
2
I
-
5
4
9
Baseline
Firing
Rate
(percent
(CR)
-
-
-
-
99
100
-
98
100
--
99
100
100
Stoichiometry
to Active
Burners
(percent)
--
--
--
--
118
113
--
118
110
~
118
112
115
NOX
Emissions
(ppn dry
0 3X 02)
--
--
--
-
398
945
--
355
421
--
376
683
530
Controlled/Low Load and OSC and FGR
Firing
Rate
(percent
MCR)
--
--
--
--
46
43
--
62
65
--
54
54
54
Type of
OSC
--
--
--
--
BOOS
OFA
BOOS
OFA
--
BOOS
BOOS
--
BOOS
OFA
BOOS
OFA
BOOS
OFA
Stoichiometry
to Active
Burners
(percent)
--
--
--
--
87
90
—
92
81
--
90
86
88
FGR
(percent)
--
-
-
-
39
27
--
30
20
--
35
23
29
NOX
Emissions
(ppm dry
0 3* 02)
-
-
-
—
194
130
—
152
171
—
173
150
162
Average NOX
Reduction
(percent)
-
-
—
--
56
87
-
57
59
-
57
73
66
Maximum NOX
Reduction
Reported
(percent)
-
-
--
-
59
90
--
57
83
—
58
87
73
I
ro
ro
-------
The average NO reductions reported in Tables 4-3 through 4-10 were
y\
calculated in the following manner. First, reductions obtained from all
tests on each particular boiler were arithmetically averaged. Only tests
with the same NO control technique were used. Next, these average NO
A A
reductions were again arithmetically averaged using all the boilers within
the same firing type/fuel classification. All other numerical table
entries, such as burner stoichiometry, firing rate, etc., were calculated in
a similar manner.
It should be noted that baseline emission data vary occasionally
between tables for the same firing type/fuel classification. The reason for
this variation is that only the baseline tests and the corresponding
controlled tests with the particular control technique under consideration
were used in the averaging procedure described above. For example, if NOV
/\
emissions from boiler "A" firing coal were controlled only with the
technique of low excess air, then the baseline data from that boiler would
be used only to calculate the average emissions reported in Table 4-3, and
not used in deriving average baseline emissions in Tables 4-4 through 4-10,
for other techniques. NO reductions reported in these tables represent
A
values typical of what can be expected when control techniques are
implemented. Descriptions of these control techniques follow.
4.2.1 Low Excess Air (LEA)
Reducing the excess air level in the furnace has generally been found
to be an effective method of NO control. In this technique, the
/\
combustion air is reduced to the minimum amount required for complete
combustion, maintaining acceptable furnace cleanliness and steam temperature.
With less oxygen available in the flame zone, both thermal and fuel NO
A
formation are reduced (Reference 4-22). In addition, the reduced airflow
lowers the quantity of flue gas released resulting in an improvement in
boiler efficiency.
Low excess air firing is usually the first NO control technique
A
applied. It may be used with virtually all fuels and firing methods.
However, furnace slagging and tube wastage considerations may limit the
degree of application (Reference 4-22). Low excess air may also be employed
in combination with the other NO control methods (Reference 4-33).
X
4-23
-------
Many units use excess air for control of steam temperature,
especially at lower loads, often as an alternative to flue gas
recirculation. Reducing the excess air levels on these units would tend to
lower the outlet steam temperature and thus reduce cycle efficiency unless
the improvement in boiler efficiency is enough to compensate for the lower
steam temperature (Reference 4-34).
In low excess air firing, there is often a greater burden on
operating personnel. The attempt to optimize the excess air level requires
close monitoring of flue gas 0- and CO analyzers. In coal firing, the
operator must also check the furnace periodically for excessive slag
deposits. Accurate flue gas analyzers will often need to be purchased if
not already installed.
As shown in Table 4-3, Average N0¥ Reduction with Low Excess Air
/\
Firing, low excess air firing results in an average NO reduction of
/\
16 percent for coal, 21 percent for oil, and 20 percent for natural gas
firing.
4.2.2 Off Stoichiometric Combustion (OSC)
Off Stoichiometric, or staged combustion seeks to control NO by
A
carrying out initial combustion in a primary, fuel-rich, combustion zone,
then completing combustion, at lower temperatures, in a second, fuel lean
zone. In practice, OSC is implemented through biased burner firing (BBF),
burners out of service (BOOS), or overfire air injection (OFA).
4.2.2.1 Biased Burner Firing (BBF), Burners Out of Service (BOOS)
Biased burner firing consists of firing the lower rows of burners
more fuel rich than the upper rows of burners. This may be accomplished by
maintaining normal air distribution to the burners while adjusting fuel flow
so that a greater amount of fuel enters the furnace through the lower rows
of burners than through the upper rows of burners. Additional air required
for complete combustion enters through the upper rows of burners which are
firing air rich.
In the burners out of service mode, individual burners, or rows of
burners, admit air only. This reduces the airflow through the fuel
admitting or active burners. Thus the burners are firing more fuel rich
than normal, with the remaining air required for combustion being admitted
through the inactive burners.
4-24
-------
These methods reduce NO emissions by reducing the excess air
/v
available in the firing zone. This reduces fuel and thermal NO formation.
^
These techniques are applicable to all fuels and are particularly attractive
as control methods for existing units since few, if any, equipment
modfications are required (References 4-33 and 4-35). In some cases,
however, derating of the unit may be required if there is too limited extra
firing capability with the active burners. This is most likely to be a
problem with pulverized coal units without spare pulverizer capacity.
Monitoring flue gas composition, especially Op and CO concentrations,
is very important when employing these combustion modifications for N0x
control. Local reducing atmospheres may cause increased tube wastage when
firing coal and high sulfur oils. They may also cause increased furnace
slagging when burning coal because of the lower ash fusion temperature
associated with reducing atmospheres (References 4-34 and 4-36). In
addition, it is important to closely monitor flue gas, excess air, and CO to
avoid reducing boiler efficiency through flue gas heat and unburned
combustible losses, and to prevent unsafe operating conditions caused by
incomplete combustion. For these reasons, accurate flue gas monitoring
equipment and increased operator monitoring of furnace conditions are
required with these combustion modifications.
As shown in Table 4-4, burners out of service firing results in an
average NO reduction of 31 percent for coal, 37 percent for oil, and 35
X
percent for natural gas firing. A typical burners out of service pattern is
shown in Figure 4-5(a).
4.2.2.2 Overfire Air (OFA)
The overfire air technique for NO control involves firing the
burners more fuel rich than normal while admitting the remaining combustion
air through overfire air ports ,or an idle top row of burners.
Overfire air is very effective for NO reduction and may be used
/\
with all fuels. However, there is an increased potential for furnace tube
wastage due to local reducing conditions when firing coal or high sulfur
oil. There is also a greater tendency for slag accumulation in the furnace
when firing coal (References 4-22, 4-35 through 4-37). In addition, with
reduced airflow to the burners, there may be reduced mixing of the fuel and
4-25
-------
WINDIOX
IICONOAIT
All NOZ21IS
SICONDAlr All DAMMIS
IICONOAIT AM
OAMfll DIIV1 UNII
O Active burners
)9C Burners admitting air only
a. Typical burners out of
service arrangement
opposed fired unit
— COAL NOZ2HS
b. Typical overfire air system for
tangential fired unit (Reference 4-21)
Burners
Air
Forced draft fan
Apportioning
dampers
Flue gas recirculating
fan
c. Typical flue gas recirculation system for NO control
/\
Figure 4-5. Typical arrangements for (b) overfire air, (a) burners
out of service, and (c) flue gas recirculation.
4-26
-------
air. Thus, additional excess air may be required to ensure complete
combustion. This may result in a decrease in efficiency (References 4-35
and 4-37).
Overfire air is more attractive in original designs than in retrofit
applications for cost considerations. Additional duct work, furnace
penetrations, and extra fan capacity may be required. There may be physical
obstructions outside of the boiler setting making installation more costly.
Or, there may also be insufficient height between the top row of burners and
the furnace exit to permit the installation of overfireair ports and the
enlarged combustion zone created by the staged combustion technique
(Reference 4-35).
As shown in Table 4-5, the limited data indicate that with overfire
air, NO reductions of about 31 percent for coal, 24 percent for oil, and
A
59 percent for natural gas are possible. A typical overfire air system is
shown in Figure 4-5(b).
4.2.3 Low NOW Burners (LNB)
y\
Several utility boiler manufacturers have been active in the
development of new burners designed to reduce NO emissions from
X
coal-fired units. Although the techniques of low excess air and off
stoichiometric (staged) combustion have been shown to be effective in
reducing NO levels, there has been some concern as to potential increased
/\
slagging and corrosion with OSC operation. Furnaces fired with certain
Eastern U.S. bituminous coals with high sulfur contents may be especially
susceptible to corrosion attack under reducing atompsheres. Local reducing
atmosphere pockets may exist under off stoichiometric operation. The
problem may be further aggravated by slagging since slag generally fuses at
lower temperatures under reducing conditions. The sulfur in the molten slag
may then readily attack the tube walls. Faced with these potential problems
and stricter N0x NSPS, manufacturers are developing and marketing low
NOX burners which permit staging at the burners themselves, away from the
water wells, thus minimizing the potential corrosion and slagging problems
associated with OSC operation.
Most low NO burners designed for utility boilers control NO by
^ X
reducing flame turbulence, delaying fuel/air mixing, and establishing
fuel-rich zones where combustion initially takes place. This represents a
departure from the usual burner design procedures which promote high
4-27
-------
turbulence, high intensity, rapid combustion flames. The longer, less
intense flames produced with low NO burners result in lower flame
A
temperatures which reduce thermal NO generation. Moreover, the reduced
A
availability of oxygen in the initial combustion zone inhibits fuel NO
A
conversion. Thus, both thermal and fuel NO are controlled by the low
A
NOX burners.
The Babcock and Wilcox Company is currently installing the Dual
Register Pulverized Coal-Fired Burner in all its new utility boilers in
order to meet current NSPS (References 4-38 and 4-39). The limited
turbulence, controlled diffusion flame burner is designed to minimize fuel
and air mixing at the burner to that required to obtain ignition and sustain
stable combustion of the coal. A Venturi mixing device, located in the coal
nozzle, provides a uniform coal/primary air mixture at the burner.
Secondary air is introduced through two concentric zones surrounding the
coal nozzle, each of which is independently controlled by inner and outer
air zone registers. Adjustable spin vanes are located in the inner air zone
to provide varying degrees of swirl to the inner air to control coal/air
mixing during the combustion process. In addition, the windbox is
compartmented to provide airflow control on a per pulverizer basis, thus
permitting operation with lower excess air while maintaining an oxidizing
atmosphere around each burner.
To date seven dual register burner-equipped utility boilers have been
tested for NOX emissions (Reference 4-39). For the four bituminous
coal-fired units tested, NO emissions ranged from 194 to 258 ng/J (0.45
to 0.6 lb/106 Btu, 318 to 422 ppm) at or below the current 258 ng/J
(0.6 lb/106 Btu) NSPS for bituminous coal. The three subbituminous
coal-fired units exhibited NOX emissions in the range of 129 to 151 ng/J
(0.3 to 0.35 lb/106 Btu, 211 to 247 ppm), well below the current 215 ng/J
(0.5 lb/106 Btu) NSPS for subbituminous coal. It should be noted that
these low NOX burner-equipped boilers came onstream when the original 1971
NSPS of 301 ng/J (0.7 lb/106 Btu) was still in effect.
Comparisons with NO emissions from similar units equipped with the
/\
high turbulence older burners show reductions in NO levels from 40 to
A "*
60 percent due to the new burner design. As explained above, the majority
of the reduction is attributable to controlled air-coal mixing in the
furnace chamber. The resulting lower peak flame temperature and the
4-28
-------
decreased availability of oxygen in the primary flame zone tend to suppress
thermal NO generation and fuel nitrogen conversion.
J\
B&W claims that NO control through its Dual Register Burners is
^
superior to staging as it maintains the furnace in an oxidizing environment,
hence minimizing slagging and reducing the potential for furnace wall
corrosion when firing high sulfur bituminous coal. Also, more complete
carbon utilization can be achieved due to better coal-air mixing in the
furnace. Finally, lower oxygen levels are required with all the combustion
air admitted through the burners rather than having some of the total air
injected above the burner zone.
Although the Dual Register Burners were developed for use in new
boilers, they can also be retrofitted to older units. However, the new
boilers are also designed to provide airflow control on a per pulverizer
basis. This may not be possible in some of the older units, or the cost
involved in retrofitting a compartmented windbox and making the necessary
changes in pulverizer burner piping may be prohibitive. If careful control
of fuel and air to each burner is not feasible, the burners will not be as
effective in reducing NOX emissions. Nevertheless, the new burners should
reduce NO levels below those obtained with the older high turbulence
A
burners. They may still be considered for retrofit application, perhaps in
conjunction with other NO control techniques, but much development work
A
remains.
Foster Wheeler Energy Corporation has developed a dual register coal
burner for installation in its new boilers (References 4-36 and 4-40). The
new burner reduces turbulence as compared to the older designs and causes
controlled, gradual mixing of fuel and air at the burner. This is achieved
using a dual throat with two registers which splits the secondary air into
two concentric streams with independently variable swirl. The mixing rate
between the primary and secondary air streams and the rate of entrainment of
furnace gases can thus be varied. The primary air velocity can also be
varied by the use of a coal nozzle in the shape of a tapered annulus with an
axially movable inner sleeve tip. In addition, a perforated plate air hood
surrounds the burner and is used to measure airflow, improve burner
circumferential air distribution, and provide a discrete means for balancing
air on a burner to burner basis.
4-29
-------
New Foster Wheeler utility boilers are equipped with OFA ports in
addition to the new low NO burners. The OFA ports are installed for use
A
in cases where the new burners alone cannot reduce NO levels to meet the
/\
NSPS requirements. However, even when staging has to be employed, it is
expected that in most cases the total burner fuel/air ratios will be above
stoichiometric, as part of the NOV reduction burden is assumed by the
A
burners. Reducing atmospheres are therefore avoided for the most part, thus
minimizing associated slagging and corrosion problems. The new burners with
cooler, less intense flames, and the larger new furnace designs with lower
burner zone heat liberation rates also tend to reduce slagging while at the
same time decrease thermal NO formation.
A
Test results for the new Foster Wheeler burners are reported in
References 4-36 and 4-40. Reductions in NO emissions of about 40 percent
A
were observed on a four-burner steam generator when operated at full load
with the new burners. Three utility steam generators, two 265 MW opposed
fired units and one 75 MW front wall fired unit, have been retrofitted with
the new burners and tested for NOX emissions. Controlled NO emissions were
in the 172 ng/J (0.4 lb/106 Btu, 281 ppm) to 215 ng/J (0.5 lb/106 Btu,
352 ppm range. Test results on one of the 265 MW units are reported and
show a 48 percent drop in NO emissions due to the new burners. When the
A
boiler was operated with the new burners and overfire air in conjunction,
reductions in NO levels of 67 percent were achieved with the OFA ports
A
100 percent open. Under such conditions, however, slag began to accumulate
after about 24 hours of continuous full load operation and unburned carbon
in the flyash increased to 2.4 percent. Under normal operating procedure,
with OFA ports not more than 20 percent open, the NO reduction was
A
approximately 40 to 50 percent over the uncontrolled case. Carbon monoxide
was maintained below 50 ppm and unburned carbon in the flyash was less than
1 percent.
The general results and trends were found to be similar for the other
two units tested. The uncontrolled NO levels of all these units were in
A
the range of 367 to 397 ng/J (600 to 650 ppm). This is atypically low for
older units equipped with high turbulence burners. Installation of new
burners and adjusting them for low NOX operation combined with OFA
operation with ports open up to 20 percent reduced NO levels down to 183
/\
to 214 ng/J (300 to 350 ppm). This represents normal low NO operating
A
4-30
-------
procedure for these units. When OFA ports are opened 100 percent and the
burners are adjusted for minimum NOV emissions, NO levels of 122 to
A A
137 ng/J (200 to 225 ppm) were attained. Slagging, however, resulted under
these operating conditions. In all cases a good quality low sulfur coal was
used so that tube wastage problems did not occur. Since the test units were
not designed for staged combustion, the slagging effect was expected.
However, slagging when overfire air ports were open was significantly less
with the low NO burners than with the original high turbulence burners.
A
Although experience with the new burners, alone and in combination
with staging, has been successful it has been limited to a few boilers and a
particular type of coal. Minimum NO levels obtained with these fuels may
^
not be repeated with a higher nitrogen content, lower heating value coal.
In addition to NOV control in new units, the Foster Wheeler dual
A
register burner is well suited (technically) for retrofit application. The
airflow to the new burners is controlled individually at each burner by
means of the perforated hood. Hence, precise air/fuel control at each
burner is possible without incurring major hardware changes besides burner
replacement.
In the tests referred to above, NO levels were approximately
A
halved by the use of the new retrofitted burners alone. The burners can
also be used with oil. In fact the original patent for the dual register
burner was designed for and tested with oil. No detailed data on oil-fired
utility boilers fitted with the new burners have been released to date.
However, Babcock & Wilcox has reported the successful retrofit of an
oil-fired dual register burner, reducing NO emissions to below 129 ng/J
(0.3 lb/106 Btu, 225 ppm) (Reference 4-39).
Riley Stoker Corporation is currently modifying the burners used in
its turbo furnace to lower NO emissions (Reference 4-41). The new
A
burners are designed to be more flexible and to control fuel/air mixing to
reduce thermal and fuel NO . With the new burners and changes in furnace
J\
design Riley Stoker expects to meet current NSPS requirements without
increased carbon or unburned hydrocarbon losses. The new burners can be
used with coal, oil, and gas fuels but are not being considered for retrofit
application. No test data are available on the performance of the new
burners at present.
4-31
-------
In summary, low NO burners appear very attractive, with potential
/\
NO reductions of the order of 50 percent. Data from long term, full
A
scale demonstrations are imminent, and commercial application is well
underway. Indeed, LNB appears to be the preferred combustion modification
technique for coal-fired utility boilers.
4.2.4 Flue Gas Recirculation (FGR)
Flue gas recirculation for NO control consists of extracting a
/\
portion of the flue gas from the economizer outlet and returning it to the
furnace, admitting the flue gas through the furnace hopper or through the
burner windbox or both. Flue gas recirculation lowers the bulk furnace gas
temperature and reduces oxygen concentration in the combustion zone
(References 4-35 and 4-37).
Flue gas recirculation through the furnace hopper and near the
furnace exit has long been used for steam temperature control. Flue gas
recirculation through the windbox and, to a lesser degree, through the
furnace hopper is very effective for NO control on gas- and oil-fired
/\
units (References 4-33 and 4-37). However, it has been shown to be
relatively ineffective on coal fired units (Reference 4-19).
Flue gas recirculation for NO control is more attractive for new
/\
designs than as a retrofit application. Retrofit installation of flue gas
recirculation can be quite costly. The fan, flues, dampers, and controls as
well as possibly having to increase existing fan capacity due to increased
draft loss, can represent a large investment. In addition, the flue gas
recirculation system itself will require a substantial maintenance program
due to the high temperature environment experienced and potential erosion
from entrained ash. Thus the cost-effectiveness of this method of NO
/\
control has to be examined carefully when comparing it to other control
techniques.
As a new design feature, the furnace and convective surfaces can be
sized for the increase in mass flow and change the furnace temperatures. In
contrast in retrofit applications, the increased mass flow increases
turbulence and mixing in the burner zone, and alters the convective section
heat absorption. Erosion and vibration problems may result (References 4-37
and 4-38). Flame detection can also be difficult with flue gas
recirculation through the windbox. In addition, controls must be employed
4-32
-------
to regulate the proportion of flue gas to air so that sufficient
concentration of oxygen is available for combustion (Reference 4-43).
As shown in Table 4-6, the limited data indicate that with flue gas
recirculation alone, average NO reductions of about 17 percent for coal,
A
13 percent for oil, and 47 percent for gas have been achieved. It should be
noted that these values are based on very limited data comparing the effects
of flue gas recirculation alone to baseline conditions. (Additional data
are discussed in Section 6.) Data comparing the effects of flue gas
recirculation in combination with other control methods to baseline
conditions are more plentiful and briefly discussed below. A typical flue
gas recirculation system is shown in Figure 4-5(c).
4.2.5 Reduced Firing Rate
Thermal NO formation generally increases as the volumetric heat
A
release rate or combustion intensity increases. Thus, NO can be
A
controlled by reducing combustion intensity through load reduction, or
derating, in existing units and by enlarging the firebox in new units. The
reduced heat release rate lowers the bulk gas temperature which in turn
reduces thermal NO formation (Reference 4-44).
The heat release rate per unit volume is generally independent of
unit rated power output. However, the ratio of primary flame zone heat
release to heat removal increases as the unit capacity is increased. This
causes NO emissions for large units to be generally greater than for
X
small units of similar design, firing characteristics, and fuel.
The increase in NO emissions with increased capacity is especially
X .
evident for gas-fired boilers, since total NO emissions are due to
A
thermal NO . However, for coal-fired and oil-fired units the effects of
/\
increased capacity are less noticeable, since the conversion of fuel
nitrogen to NO for these fuels represent a major component of total NO
X X
formation. Still, a reduction in firing rate will affect firebox
aerodynamics which may, consequently, affect fuel NOX emissions. But such
effects on fuel NO production are less significant.
A
Table 4-7 presents a compilation of available data on NO reduction
A
as a result of reduced firing rate. For coal firing, an average of 15
percent reduction in NO resulted from a 28 percent reduction in firing
/\
rate. For oil firing, an average of 30 percent reduction in NO resulted
A
from a 42 percent reduction in firing rate. For gas firing, an average of
4-33
-------
44 percent reduction in NO resulted from a 44 percent reduction in firing
A
rate. Thus, reduction of NO with lowered firing rate is most evident
X
with gas-fired boilers.
Reduced firing rate often leads to several operating problems. Aside
from the limiting of capacity, low load operation usually requires higher
levels of excess air to maintain steam temperature and to control smoke and
CO emissions. The steam temperature control range is also reduced
substantially. This will reduce the operating flexibility of the unit and
its response to changes in load. The combined results are reduced operating
efficiency due to higher excess air and reduced load following capability
due to a reduction in control range.
When the unit is designed for a reduced heat release rate, the
problems associated with derating are largely avoided. The use of an
enlarged firebox produces NO reductions similar to load reduction on
A
existing units.
4.2.6 Combination of Controls
To achieve required NO emission levels, it is often necessary to
A
use a combination of control methods. Low excess air operation is common to
all combined control method strategies. Other control combinations that are
most effective are primarily fuel dependent. It is important in this
respect to distinguish retrofit controls from original design controls.
Unfortunately the test data available are primarily from retrofit control
applications.
Tables 4-8 through 4-10 represent compilations of test data from
employing combinations of controls to reduce NO emissions. Detailed data
/\
from combined controls tests are very limited. Table 4-11 extends the data
presented in previous tables and lists highest NO reductions attained
A
through combinations of controls as a function of boiler/fuel classification.
This table represents the best results achieved in specific applications and
should not be interpreted as generally achievable NO reductions. In
A
comparing Tables 4-3 through 4-7 to Tables 4-8 through 4-10, it is seen that
in combining control techniques, results are complementary but not additive
for NOX reduction.
4.3 ADVANCED CONTROLS
Several other combustion NO control techniques, which show promise
A
for future application, are in varying stages of development.
4-34
-------
TABLE 4-11. MAXIMUM REPORTED NOX REDUCTION ACHIEVED WITH
BOILER LOAD AT OR ABOVE 80 PERCENT MCR*
Equipment
Type
Tangential
Opposed Wall
Single
Wall
Average
All
Boilers
Fuel
Coal
Oil
Nat
Gas
Coal
Oil
Nat
Gas
Coal
Oil
Nat
Gas
Coal
Oil
Nat
Gas
All
Fuels
Control Techniques
Implemented
OFA
Reduced Firing Rate (RFR)
+ BOOS + FGR
RFR + FGR
BOOS
BOOS + OFA
BOOS + OFA
BOOS
BOOS + FGR
BOOS
(LEA)b
BOOS + OFA
BOOS + OFA + FGR
BOOS + OFA
BOOS + OFA <• FGR
Firing Rate
(percent NCR)
85
68
50
83
100
100
82
(81)t>
%
98
(100)
83
98
99
94
Stolen iometry
to Active Burners
(percent)
85
110
110
80
73
69
86
(80)
91
88
(123)
84
82
78
81
Furnace
Stoichiometry
(percent)
113
122
no
107
119
111
115
(120)
121
117
(123)
112
120
114
115
FGR
(percent)
--
NA
32
-
--
--
—
40
-
--
20
--
7
NOX Emissions
(ppm dry 9 3X 0?)
196
110
65
334
222
205
225
(386)
145
109
(931)
252
183
194
210
Maximum NOX
Reduction
(percent)
66
55
81
53
53
79
63
(68)
60
71
(6)
61
56
67
61
co
en
aFor individual tests, corresponding baseline and controlled loads were nearly identical.
bNumbers in parentheses refer to boilers originally designed for coal firing with wet bottom furnaces.
-------
advanced burner and furnace concepts, and non-catalytic homogeneous NO
A
reduction with ammonia injection in the boiler's convective section.
4.3.1 Advanced Burner/Furnace Designs
A number of advanced burner designs are being developed and tested
to reduce NO emissions from coal- and oil-fired utility and industrial
A
boilers. Advanced burners, as compared to low NO burners, are defined
A
as those devices still under experimental or pilot scale development for
lowering NO emissions. Burner modification has the potential of
A
lowering NO emissions well below levels attainable by conventional
A
combustion modification techniques. Burner modification also has the
advantage of requiring minimal changes in current boiler design and
operation and is suitable for retrofit application.
TRW, Incorporated is developing an advanced burner for oil- and
gas-fired commercial and industrial boilers with potential application to
utility boilers. The burner uses shaped fuel injection ports to control
fuel and air mixing and entrain combustion products into the flame zone
(Reference 4-45). In addition to reducing thermal NO , the burner is
A
effective in controlling fuel nitrogen conversion. In tests with residual
oils in a packaged boiler and a large industrial size boiler, the burner
was capable of reducing NO emissions by about 30 percent to values
A
below 200 ppm. A preliminary timetable for the industrial burner calls
for commercial application at the end of 1979 (Reference 4-46). An
EPA-sponsored field demonstration is underway and actual operating data
should soon be available (Reference 4-47).
Some manufacturers of oil-firing equipment are in the process of
developing burners capable of operating at very low levels of excess air.
The low excess air requirements increase boiler efficiency and reduce fan
power consumption while decreasing NO emissions. The low excess air
A
may also reduce SO^ conversion. The Peabody Engineering Company has
designed the Air Pressure Recovery (APR) burner designed to operate at
excess oxygen levels down to 1/2 percent without increase in particulate
and unburned hydrocarbon emissions. The Coen Company is developing the
LEA burner which uses a tip swirler to operate down to 0.1 percent excess
oxygen (Reference 4-48). Both burners are currently undergoing testing
and no data on NO emissions are available.
A
4-36
-------
For coal-fired utility boilers, Foster Wheeler is currently testing
an advanced dual register split frame burner design. A device added at the
burner nozzle splits the primary air-coal flow into several distinct
streams. Coal particles become concentrated within each stream and, hence
diffuse more slowly into the secondary air. This further inhibits NOX
formation by extending the slow-burning characteristics of the dual register
burner. Results from an industrial size test boiler are promising with a
NOV level of approximately 129 ng/0 (0.3 lb/106 Btu) for subbituminous
A
coal (Reference 4-40). However, the burner tested on a 375 MW electrical
output boiler produced approximately 215 ng/J (0.5 lb/10 Btu). A further
modification of this burner with a variable velocity split flame nozzle will
be installed, and a NOX level of 151 to 172 ng/J (0.35 to 0.4 lb/106 Btu)
is expected (Reference 4-49). The new design permits the velocity of the
primary air-coal stream to be optimized for minimum NO consistent with
A
flame stability and minimum CO.
Babcock & Mil cox and Energy and Environmental Research, under EPA
sponsorship, are developing an advanced utility coal burner for low N0x,
the distributed fuel/air mixing burner, for field testing (Reference 4-50).
The burner is designed to control both thermal and fuel NO . It is
A
estimated that in uncontrolled pulverized coal combustion, thermal NOX
represents approximately 15 percent of the total NOX, the volatile
component of fuel NO contributes 65 percent, and the char component about
A
20 percent (Reference 4-51). In the distributed mixing burner thermal NO
A
is reduced by minimizing peak flame temperature. Volatile NOX is reduced
by maintaining fuel-rich conditions in the flame zone. NO formation can
A
also be reduced by increasing residence times in the rich zone, thus
promoting reduction of NO by hydrocarbon and char fragments. For char
A
NO , no effective control measures are available, but the char component
A
can be reduced by maximizing evolution of nitrogen with the volatiles. This
can be accomplished by providing for adequate residence times in the rich
flame zone at high temperature.
The distributed fuel-air mixing burner design injects coal and
primary air from the center of the burner with a moderate axial component.
This stream is surrounded by a divided secondary airstream with a swirl
component for stabilization. Tertiary air for burnout is added axially
around the periphery of the burner. The arrangement results in a hot, rich
4-37
-------
recirculation zone at the center of the flame with stoichiometric ratios as
high as 2 or more. Adequate time at high temperature is also provided to
maximize evolution of nitrogen from the char. This time in the rich zone
helps reduce most of the NO that may be formed. Also, axial addition of
/\
the tertiary air leads to a large flame zone. Heat extraction prior to
completion of burnout along with dilution of the tertiary air by combustion
products lowers the peak flame temperature, thus reducing thermal NO .
A
Although experimental prototypes have achieved NO emissions below 86 ng/J
fi
(0.2 lb/10 Btu), actual field testing is not expected to be complete
until late 1982 (Reference 4-50).
Babcock and Wilcox Company is developing a primary combustion furnace
concept for coal-fired utility boilers in a program sponsored by the
Electric Power Research Institute (Reference 4-52). The fundamental process
to control NOX in this concept is conversion of fuel nitrogen to N~
through fuel-rich combustion. Pulverized coal is introduced into an
extended combustor with substoichiometric air, so that combustion occurs
under fuel-rich conditions isolated from the rest of the furnace. The
length of the combustor is sufficient to provide the necessary residence
time to partially oxidize the coal and permit the desirable N? producing
reactions to occur. Heat is removed along the combustion chamber to prevent
slagging. Secondary air is added at the exit of the primary combustion
furnace to bring the combustion products to oxidizing conditions before they
enter the furnace. Pilot scale testing of a 1 MW (4 x 10 Btu/hr) heat
input prototype has achieved the targeted NO level of below 86 ng/J
c X
(0.2 lb/10 Btu). Commercial offering of a full scale furnace is not
expected until at least 1983 (Reference 4-52).
In summary, advanced burner/furnace concepts though promising, still
require several years of development. It remains to be seen whether these
advanced burners may need to be combined with other combustion modification
techniques as well.
4.3.2 Ammonia Injection
The use of ammonia as a potential homogeneous NO reducing agent
/v
was first reported by Wendt, et al. (Reference 4-53). However, these
authors attributed their results to the pyrolysis of ammonia to hydrogen
with the hydrogen in turn reacting with NO. The postflame decomposition of
NO by reducing agents has more recently shown promise as a method for
A
4-38
-------
augmenting combustion modifications if stringent emission limits are to be
met. Lyon (Reference 4-54) has reported that selective homogeneous
reduction of NO in combustion effluents was possible with direct injection
of ammonia within a specific temperature range.
The gas phase reaction in the temperature range of 1090K (1500 F)
to 1310K (1900°F) converts nitric oxide, in the presence of oxygen and
ammonia, into nitrogen and water according to the following chain reaction
(Reference 4-55):
NH2 + NO N2 + H + OH (4-6)
NH2 + NO N2 + H20 (4-7)
H + 02 OH + 0 (4-8)
0 + NH3 OH + NH2 (4-9)
OH + NH3 H20 + NH2 (4-10)
H + NH3 H2 + NH2 (4-11)
Oxygen acts as a catalyst in reducing ammonia to the intermediate NH2
compound which in turn reacts selectively with NO, reducing it to N2 and
water. Based on this discovery a patent under the name Thermal De-N0x was
issued to Exxon Research and Engineering for this NO reduction technique.
Results of lab scale tests show that the level of NOX reduction
depends on the combustion product temperature, initial NO concentration,
/\
and quantity of ammonia injected. The data shown in Figures 4-6 through 4-8
was obtained by Muzio, et al. (Reference 4-56) during pilot-scale tests
using a 59 kW (200,000 Btu/hr) heat input plug flow combustion tunnel
burning natural gas. Figure 4-6 shows the effect of temperature and
ammonia/nitric oxide ratio on the reduction of NO for an initial NO level of
300 ppm and an excess oxygen level of 4 percent. It dramatically
illustrates the narrow temperature window for optimal NO reduction. This
optimal temperature range is near 1240K (1780°F), where reductions of 30
to 90 percent were achieved with ammonia injections of 0.3 to 1.6 times the
initial concentration of NO. Although increasing the ratio of ammonia to
nitric oxide reduced more NO, Figure 4-7(a) shows that for NH3/NO ratios
of greater than 2, essentially no further reduction of NO was achieved. The
additional ammonia injected at 1240K leaves unreacted. However, when the
reaction temperature is greater than 1300K, the ammonia reacts with oxygen
to form additional NO, an undesirable situation. Thus, at these higher
4-39
-------
1.0
0.8
0.6
Excess oxygen: 4:
Initial NO: 300 ppm
0.4
0.2
(NH3)/(NO)
1.6
1000
1100 1200 1300
Temperature, K
1400
Figure 4-6.
Effect of temperature on NO reduction with
ammonia injection (Reference 4-56).
4-40
-------
600
O.
o.
2% excess oxygen
(NH3)/(NO)
a. Nitric oxide reduction
(NH3/(NO)
b. Ammonia carryover
Figure 4-7. Nitric oxide reductions and ammonia carryover with ammonia
injection at 2 percent excess oxygen (Reference 4-56).
4-41
-------
temperatures essentially no ammonia leaves unreacted (see Figure 4-7(b)).
Figure 4-8 shows that for a given (NH3)/(NO) ratio, ammonia injection is
more effective at higher initial NO levels. However, this trend is only
signficiant at initial NO levels of less than 400 ppm. These factors are
important considerations in assessing the tradeoffs between implementing
only ammonia injection as a NO control or in combination with combustion
J\
modification techniques.
Byproduct pollutants from ammonia injection have been analyzed by
Lyon and Longwell (Reference 4-57) who measured emissions of N20, CO, HCN,
S03 and NH.HSO. from gas- and oil-fired pilot scale combustion
facilities. Emissions of N20 were found to be limited to 2 moles NpO
for every 100 moles of NO reduced. Ammonia was not found to react with
COp to form CO. However, the presence of ammonia in the combustion
effluent was found to inhibit the conversion of CO to COp. Therefore,
this technique becomes a problem only if the concentration of CO in the flue
gas is significant. In utility boilers this may not be the case, especially
for gas- and oil-fired units. However, coal-fired boilers with higher CO
levels may present a problem.
Cyanide can only form if unburned hydrocarbons are present in the
flue gas. For normal operation of utility boilers, only a few ppm of HCN
can be found.
Careful laboratory work has shown that ammonia injection does not
produce additional SOo emissions (Reference 4-58). In fact, S0? levels
remained unchanged during the injection stage. The main byproduct pollutant
of concern is ammonium bisulfate. The unreacted NH7 leaving De-NO
«5 X
reaction was found to combine with S03 and HpO to form ammonium
bisulfate. This substance forms a very corrosive liquid at 480K to 530K
(400°F to 500°F). Thus it could potentially corrode sections of the
boilers such as the air preheater and flue gas ducts if the concentration of
ammonia bisulfate is significantly high. During full scale studies of an
oil-fired boiler however, no evidence of additional corrosion was found
(Reference 4-59).
Full scale application of the Exxon Thermal De-NO process has been
A
reported for six gas- and oil-fired combustion sources in Japan affiliated
with Exxon Corporation. Figure 4-9 shows that the average NO reduction
A
reported with the ammonia injection technique is of the order of 50 percent.
4-42
-------
i
-p»
CO
0.6 -
Excess oxygen: 2"
Temperature: 1233K
Initial NO level (ppm)
100
200
400
0 680
1050
.= 0.4 -
0.2 -
0
(NH,)/(NO)
Figure 4-8. Effect of initial nitric oxide concentrations on NO
reduction with ammonia injection (Reference 4-56).
-------
60
:.
50
o 40
30
20
I
Unit Description
70 t/hr steam boiler
150 kbbl/d crude heater
A 430 t/hr steam boiler
^430 t/hr steam boiler
O 120 t/hr steam boiler
O]50 kbbl/d crude heater
1000 1050 1100 1150 1200 1250
Injection zone flue gas temperature, K
1300
Figure 4-9. Performance of Thermal De-NO systems in commercial applications
(Reference 4-55).
-------
One other full scale combustion facility in the United States has been
retrofitted with ammonia injection. NO emissions were reduced from
A
270 ppm to 80 to 120 ppm (Reference 4-59). Details of the process are not
available.
The Thermal De-NO process for NO emission reduction shows a
f^ A
promising application for utility boilers, with potential NO reductions
A
of 40 to 60 percent. However, full scale studies have been limited to gas
and oil which are becoming less available in the utility fuel market. Pilot
scale tests using ammonia injection on coal-fired furnaces have been
completed by KVB under EPRI sponsorship (Reference 4-60). Basically the
study confirmed the effectiveness of the technique as well as its potential
limitations such as the narrow temperature window and possible ammonia
byproduct emissions.
EPA has assessed the applicability and effectiveness of ammonia
injection in two recent studies (References 4-61 and 4-62). The studies
conclude that ammonia injection holds promise for additional NOV
X
reductions, 40 to 60 percent, in those air quality regions where stringent
NO controls may be required. Ammonia injection could be applied as an
A
add-on technology, in combination with conventional combustion modification
techniques. However, a number of limitations need to be considered and
evaluated before the process is retrofitted, especially for coal-fired
boilers:
• Performance is very sensitive to flue gas temperature, and is
maximized only within a 50K temperature gradient from the optimum
temperature of about 1240K. This temperature sensitivity may
require special procedures for load following boilers, such as
multiple NH3 injection grids.
• Performance is very sensitive to flue gas residence time at
optimum temperatures. High flue gas quench rates are expected to
reduce process performance.
t Costs of the process can be much higher than for other combustion
controls
• Successful retrofit application is highly dependent on the
geometry of convective section
0 Byproduct emissions such as ammonium bisulfate might cause
operational problems, such as air preheater fouling, especially
in coal-fired boilers
4-45
-------
• Ammonia emissions may be an environmental problem if the process
is not carefully controlled
Exxon is currently investigating possible solutions to these
potential problems. Ammonia injection can potentially offer a near term
control option for achieving NO emission levels not obtainable with
A
current state-of-the-art controls (Reference 4-62).
4.4 MINOR EMPHASIS CONTROLS
In this section, controls which were given minor emphasis in the
present study are briefly discussed. These were treated in less detail
because they were considered to have less promise for widespread application
than those described above, for such reasons as energy penalties, high cost,
or technical difficulties. However, flue gas treatment techniques are
included here largely because they are being studied in greater depth in
other efforts (References 4-63 and 4-64).
4.4,1 Reduced Air Preheat
Thermal NO production is strongly influenced by the effective peak
A
temperatures in the combustion zone. Thus, any modification that lowers
these temperatures, such as reducing the combustion air temperature, should
lower NO emissions. Theory indicates that a 56K (100°F) decrease in
air preheat temperature will result in an approximately 28K (50°F)
reduction in the adiabatic combustion temperature, which in turn will
decrease thermal NO formation by 27 percent (References 4-44 and 4-65).
A
Since reduced air preheat does not significantly suppress fuel nitrogen
conversion (Reference 4-66), it is expected that this control technique
would be most effective on fuels, such as natural gas and distillate oil,
which have low nitrogen content.
Reduced air preheat is potentially applicable to most utility boilers
because these sources are equipped with regenerative air heaters which
preheat combustion air. This method for controlling NO usually greatly
A
lowers fuel economy, however. New designs to reduce stack gas temperatures,
for example, and redesign of the convective section of a boiler for more
heat absorption would be necessary to maintain efficiency.
Only limited field test data are available on the effect of reduced
air preheat in utility boilers due to the severe efficiency penalty incurred
4-46
-------
with this method. Some field test results and discussions on reduced air
preheat for utility boilers are available in References 4-67 through 4-69.
The data reported for coal firing showed varying trends, although a
maximum reduction of 75 ppm (at zero percent 02) per 56K reduction in air
temperature was reported in one case (Reference 4-66). In general, NOX
reductions of about 50 percent for gas-fired boilers and 40 percent for
oil-fired boilers can be expected with reduced air preheat, in contrast to
the relatively small reductions in coal-fired boilers (Reference 4-77).
In summary, reduced air preheat reduces efficiency, and is therefore
not considered a practical control technique for existing units. Design
changes in new units, such as installing or enlarging an economizer, would
be required to regain the waste heat which would otherwise be lost through
the stack.
4.4.2 Water Injection
Water injection has been shown to reduce flame temperature and is
widely used in gas turbines. Only recently has water injection been tried
on utility boilers.
The Ormond Beach, steam generating units operated by Southern
California Edison were tested with water injection to reduce NO
A
(Reference 4-33). The boilers operating at 75 percent of full load (design
capacity 800 MW) with 10 percent tertiary air, were emitting 400 ppm of NO
when 0.6 kg of water per kg of oil was injected, the emissions were reduced
to 228 ppm, a 43 percent reduction. Higher reductions were obtained with
flue gas recirculation and water injection combined. For example, with
15 percent gas recirculation and injection of 0.2 kg of water/kg of oil, NO
reduction of nearly 50 percent was achieved. Compared to flue gas
recirculation, water injection imposes a large energy penalty. Water
injection increased the minimum 0~ requirement and reduced boiler
efficiency by 10 percent in the Ormond Beach case. The large efficiency
loss due to water injection makes this technique unattractive to the utility
sector.
In summary, water injection is not seen as a feasible NO reduction
A
technique for utility boilers based on the large energy penalty incurred.
Thus little current work with this technique is being performed on large
steam generators.
4-47
-------
4.4.3 Flue Gas Treatment
While combustion modification techniques seek to lower NO
A
emissions by minimizing NO formation, flue gas treatment (FGT) processes
involve post-combustion NO removal from the flue gas. Flue gas treatment
A
has potential for use combined with combustion modifications when very high
removal efficiencies are required (References 4-63, 4-64, and 4-76).
FGT has been applied to only a few commercial oil- and gas-fired
boilers in Japan. No FGT installation for NO control on utility boilers
A
exists in the United States as combustion modifications represent the most
cost effective approach to achieving moderate NO reductions. However,
A
combustion modifications alone may not be able to provide the degree of
control necessary to meet future N02 ambient air quality standards. Thus
EPA has initiated several demonstration projects to investigate the use of
FGT in the U.S. (Reference 4-64).
FGT processes can be divided into two main categories: dry processes
and wet processes. Dry processes reduce NO by catalytic reduction and
A
operate at temperatures between 570 to 700K (570 to 800 F). Wet systems
are generally either oxidation/absorption or absorption/reduction processes,
both operating in the 310K to 320K (100 to 120°F) range.
Among the many dry process variations, selective catalytic reduction
(SCR) using ammonia has been perhaps the most successful. Over 50 percent
NO , and often up to 90 percent reductions have been claimed using such
A
processes. However, plugging of the catalyst bed and fouling of the
catalyst itself are major operational concerns, especially with coal
firing. Moreover, use of SCR has raised concerns in that any ammonia left
in the flue gase may combine with existing SOg/SOp to produce a visible
plume, and byproducts, such as ammonium bisulfate, which are corrosive to
boiler equipment.
Wet FGT processes utilize more complex chemistry than dry processes.
In the oxidation/absorption processes, strong oxidants such as ozone or
chlorine dioxide are used to convert the relatively inactive NO in the flue
gas to N02 or N205 for subsequent absorption. In the absorption/
reduction processes, chelating compounds, such as ferrous ethylenediamine-
tetracetic acid are required in the scrubbing solution to trap the NO.
However, because wet processes rely on absorption, most of them create
troublesome byproducts such as nitric acid, potasium nitrate, ammonium
4-48
-------
sulfate, calcium nitrate, and gypsum which may have little commercial
value. In addition, the high cost of an absorber and an oxidant or
chelating agent is likely to be prohibitive for flue gases with high NO
/\
concentrations.
In general, the dry F6T techniques used in Japan can probably be
applied to gas- and oil-fired sources in the U.S. However, the
applicability of dry processes to coal-fired boilers remains to be
demonstrated. Wet processes are less well developed and costlier than dry
FGT processes. However, wet simultaneous, as well as dry simultaneous
NO /SO processes warrant further investigation. In any case, more
A A
field tests are needed to determine the costs, secondary effects,
reliability, and waste disposal problems. Flue gas treatment holds some
promise as a control technique for use when high NO removal efficiencies
A
are necessitated by stringent emission standards. However, compared to
combustion modifications FGT is considerably more expensive.
4-49
-------
REFERENCES FOR SECTION 4
4-1. Zeldovich, J., "The Oxidation of Nitrogen in Combustion and
Explosions," Acta Physiochem URSS. (Moscow), Vol. 21, p. 4, 1946.
4-2. Bowman, C. T. and Seery D. J. , "Investigation of NO Formation
Kinetics in Combustion Processes: The Methane-Oxygen-Nitrogen
Reaction," in Emissions from Continuous Combustion Systems.
Cornelius, W. and Agnew, W. G., eds., Plenum, 1972.
4-3. Bartok, W., et al., "Basic Kinetic Studies and Modeling of NO
Formation in Combustion Processes," AIChE Symposium Series No. 126.
Vol. 68, 1972.
4-4. Halstead, C. J. and Munro, A. J. E., "The Sampling, Analysis, and
Study of the Nitrogen Oxides Formed in Natural Gas/Air Flames,"
Company Report, Shell Research, Egham, Surrey, U.K., 1971.
4-5. Thompson, D., et al., "The Formation of Oxides of Nitrogen in a
Combustion System," presented at the 70th National AIChE Meeting,
Atlantic City, 1971.
4-6. Lange, H. B., "NOX Formation in Premixed Combustion: A Kinetics
Model and Experimental Data," presented at the 64th Annual AIChE
Meeting, San Francisco, 1971.
4-7. Sarofim, A. F. and Pohl, J. H., "Kinetics of Nitric Oxide Formation
in Premixed Laminar Flames," 14th Symposium (International) on
Combustion, The Combustion Institute, Pittsburg, 1973.
4-8. Iverach, D., et al., "Formations of Nitric Oxide in Fuel-Lean and
Fuel-Rich Flames," ibid., 1973.
4-9. Wendt, J. 0. L. and Ekmann, J. M., "Effect of Fuel Sulfur Species
on Nitrogen Oxide Emissions from Premixed Flames," Comb. Flame.
Vol. 25, 1975.
4-10. Malte, P. C. and Pratt, D. T., "Measurement of Atomic Oxygen and
Nitrogen Oxides in Jet-Stirred Combustion," 15th Symposium
(International) on Combustion, The Combustion Institute,
Pittsburgh, 1975.
4-11. Mitchell, R. E. and Sarofim, A. F., "Nitrogen Oxide Formation in
Laminar Methane Air Diffusion Flames," presented at the Fall
Meeting, Western States Section, The Combustion Institute, Palo
Alto, California, 1975.
4-12. Bowman, C. T., "Non-Equilibrium Radical Concentrations in Shock
Initiated Methane Oxidation," 15th Symposium (International) on
Combustion, The Combustion Institute, Pittsburg, 1975.
4-50
-------
4-13. Fem'more, C. P., "Formation of Nitric Oxide in Premixed Hydrocarbon
Flames," 13th Symposium (International) on Combustion, The
Combustion Institute, Pittsburgh, 1971.
4-14. MacKinnon, D. J., "Nitric Oxide Formation at High Temperatures,"
Journal of the Air Pollution Control Association, Vol. 24, No. 3,
pp. 237 to 239, March 1974.
4-15. Heap, M. P., et al., "Burner Criteria for NOX Control; Volume I
-- Influence of Burner Variables on NOX in Pulverized Coal
Flames," EPA 600/2-76-061a, NTIS-PB 259 911/AS, March 1976.
4-16. Bowman, C. T., et al., "Effects of Interaction Between Fluid
Dynamics on Chemistry or Pollutant Formation in Combustion," in
Proceedings of the Stationary Source Combustion Symposium; Volume I
— Fundamental Research, EPA 600/2-76-152a, NTIS-PB 256 320/AS,
June 1976.
4-17. Shaw, J. T. and Thomas, A. C., "Oxides of Nitrogen in Relation to
the Combustion of Coal," presented at the 7th International
Conference on Coal Science, Prague, June 1968.
4-18. Pershing, D. W., et al., "Influence of Design Variables on the
Production of Thermal and Fuel NO from Residual Oil and Coal
Combustion," AIChE Symposium Series, No. 148, Vol. 71, pp. 19 to
29, 1975.
4-19. Thompson, R. E. and McElroy, M. W., "Effectiveness of Gas
Recirculation and Staged Combustion in Reducing NOX in a 560-MW
Coal-Fired Boiler," EPRI FP-257, NTIS-PB 260 582, September 1976.
4-20. Sarofim, A. F., et al., "Mechanisms and Kinetics of NOX
Formation: Recent Developments," presented at the 65th Annual
AIChE Meeting, Chicago, November 1976.
4-21. Martin. G. B. and Berkau, E. E., "An Investigation of the
Conversion of Various Fuel Nitrogen Compounds to Nitrogen Oxides in
Oil Combustion," presented at the 70th National AIChE Meeting,
Atlantic City, August 1971.
4-22. Habelt, W. W. and Howell, B. M., "Control of NOX Formation in
Tangentially Coal-Fired Steam Generators," in Proceedings of the NO
Control Technology Seminar. EPRI SR-39, NTIS-PB 253 661, February
4-23. "Air Quality and Stationary Source Emission Control," U.S. Senate,
Committee on Public Works, Serial No. 94-4, March 1975.
4-24. Pohl, J. H. and Sarofim, A. F., "Fate of Coal Nitrogen During
Pyrolysis and Oxidation," in Proceedings of the Stationary Source
Combustion Symposium; Volume i — hunaamentai Research, EHA
600/2-76-152a, NTIS-PB 256 320/AS, June 1976.
4-51
-------
4-25. Heap, M. P., et al., "The Optimization of Burner Design Parameters
to Control NOX Formation in Pulverized Coal and Heavy Oil
Flames," in Proceedings of the Stationary Source Combustion
Symposium; Volume II — Fuels and Process Research and Development.
EPA 600/2-76-1525, NTIS-PB 256 321/AS, June 1976.
4-26. Pohl, J. H. and Sarofim, A. F., "Devolatilization and Oxidation of
Coal Nitrogen," presented at the 16th Symposium (International) on
Combustion, Cambridge, Massachusetts, August 1976.
4-27. Blair, D. W., et al., "Devolatilization and Pyrolysis of Fuel
Nitrogen from Single Coal Particle Combustion," 16th Symposium
(International) on Combustion, Cambridge, Massachusetts, August
1976.
4-28. Brown, R. A., et al., "Investigation of Staging Parameters for
NOX Control in Both Wall and Tangentailly Coal-fired Boilers," in
Proceedings of the Second Stationary Source Combustion Symposium:
Volume III, New Orleans, EPA-600/7-77-073c, NTIS-PB 271 75//AS.
July 1977.
4-29. Pershing, D. W., "Nitrogen Oxide Formation in Pulverized Coal
Flames," Ph.D. Dissertation, University of Arizona, 1976.
4-30. Axworthy, A. E., Jr., "Chemistry and Kinetics of Fuel Nitrogen
Conversion to Nitric Oxide," AIChE Symposium Series. No. 148, Vol.
71, pp. 43 to 50, 1975.
4-31. Axworthy, A. E., et al., "Chemical Reactions in the Conversion of
Fuel Nitrogen to NOX," in Proceedings of the Stationary Source
Combustion Symposium, Volume I, EPA 600/2-76-152a, NTIS-PB 256
320/AS, June 1976.
4-32. Pershing, D. W. and Wendt, J. 0. L., "The Effect of Coal Combustion
on Thermal and Fuel NOX Production from Pulverized Coal
Combustion," presented at Central States Section, The Combustion
Institute, Columbus, Ohio, April 1976.
4-33. Norton, D. M., et al., "Status of Oil-Fired NOX Control
Technology," in Proceedings of the NOX Control Technology
Seminar. EPRI SR-39, NTIS-PB 253 661, February 1976.
4-34. Durrant, 0. W., "Pulverized Coal — New Requirements and
Challenges," presented to ISA Power Instrumentation Symposium,
Houston, Texas, May 1975.
4-35. Campobenedetto, E. J., Babcock & Mil cock Co., letter to Acurex
Corp., November 15, 1977.
4-36. Vatsky, J., "Attaining Low NOX Emissions by Combining Low
Emission Burners and Off-Stoichiometric Firing," presented at the
70th Annual AIChE Meeting, New York, November 1977.
4-52
-------
4-37. Rawdon, A. H. and Johnson, S. A., "Control of NOX Emissions from
Power Boilers," presented at Annual Meeting of the Institute of
Fuel, Adelaide, Australia, November 1974.
4-38. Campobenedetto, E. J., "The Dual Register Pulverized Coal Burner --
Field Test Results," presented to Engineering Foundation Conference
on Clean Combustion of Coal, New Hampshire, August 1977.
4-39. Barsin, J. A., "Pulverized Coal Firing NOX Control," in
Proceedings; Second NOX Control Technology Seminar, Electric
Power Research Institute, Report No. FP-1109-SR, Palo Alto,
California, July 1979.
4-40. Vatsky, J., "Experience in Reducing NOX Emissions on Operating
Steam Generators," in Proceedings; Second NOX Control Technology
Seminar, Electric Power Research Institute, Report No. FP-1109-SR,
Palo Alto, CA, July 1979.
4-41. "NOX Control Review," Vol. 2, No. 4, EPA Industrial Environmental
Research Laboratory, RTP, North Carolina, Fall 1977.
4-42. Krippene, B. C., "Burner and Boiler Alterations for NOX Control,"
presented to Central States Section, The Combustion Institute,
Madison, Wisconsin, March 1974.
4-43. Barr, W. H., et al., "Modifying Large Boilers to Reduce Nitric
Oxide Emissions," Chemical Engineering Progress, Vol. 73, pp. 59 to
68, July 1977.
4-44. Bell, A. W., et al., "Nitric Oxide Reduction by Controlled
Combustion Processes," KVB, Inc., Western States Section/Combustion
Institute, April 20-21, 1970.
4-45. Koppang, R. R., "A Status Report on the Commercialization and
Recent Development History of the TRW Low NOX Burner," TRW Energy
Systems Group Publication, Redondo Beach, CA, 1977.
4-46. Personal communication, Boughton, M., TRW, Inc., Redondo Beach, CA,
May 1979.
4-47. Matthews, B. J., TRW, Inc., Redondo Beach, CA, letter to Peters,
W., EPA, IERL-RTP, NC, March 1979.
4-48. Stavern, D. V., "The Coen Low Excess Air Burner," presented at the
NOX Control Technology Workshop, Pacific Grove, California,
October 1977.
4-49. Vatsky, J., "Larger Burners and Low NOX," Heat Engineering,
Vol. 49, No. 2, pp. 17-25, April-June 1979.
4-53
-------
4-50. Martin, G. B., "Field Evaluation of Low NOX Coal Burners on
Industrial and Utility Boilers," in Proceedings of the Third
Stationary Source Combustion Symposium, Volume I,
EPA-600/7-79-050a, February 1979.
4-51. Gershman, R., Heap, et a!., "Design and Scale-Up of Low Emission
Burners for Industrial and Utility Boilers," in Proceedings of the
Second Stationary Source Combustion Symposium, Volume II,
EPA-600/7-77-073B, NTIS-PB 271 756/9BE, July 1977.
4-52. Johnson, S. A., et al., "The Primary Combustion Furnace System —
An Advanced Low-N0x Concept for Pulverized Coal Combustion," in
Proceedings: Second N0y Control Technology Seminar, Electric
Power Research Institute, Report No. FP-1109-SR, Palo Alto, CA,
July 1979.
4-53. Wendt, J. 0. L., et al., "Reduction of Sulfur Trioxide and Nitrogen
Oxides by Secondary Fuel Injection," 14th Symposium (International)
on Combustion, The Combustion Institute, 1973.
4-54. Lyon, R. K., "Method for the Reduction of Concentrations of NO in
Combustion Effluents using Ammonia," U.S. Patent No. 3,900,554,
August 1975.
4-55. Bartok, W., "Non Catalytic Reduction of NOX with NH ," in
Proceedings of the Second Stationary Source Combustion Symposium:
Volume II, EPA-600/7-77-073b. NTIS-PB 271 756/9BE, July 1977.
4-56. Muzio, L. J., et al., "Homogeneous Gas Phase Decomposition of
Oxides of Nitrogen," EPRI Report FP-253, NTIS-PB 257 555, August
1976.
4-57. Lyon, R. R. and Longwell, J. P., "Selective, Non-Catalytic
Reduction of NOX by NH ," in Proceedings of the N0y Control
Technology Seminar, EPRI SR-39, NTIS-PB 253 661, February 1976.
4-58. Teixeira, D. P., "Status of Utility Application of Homogeneous
NOX Reduction," in Proceedings of the NOX Control Technology
Seminar. EPRI SR-39, NTIS-PB 253 661, February 1976.
4-59. "Exxon Says Stationary NOX Emissions Significantly Reduced at
Plant," Air/Water Pollution Report, p. 76, February 20, 1978.
4-60. Muzio, L. J., et al., "Noncatalytic NO Removal with Ammonia," EPRI
Report FP-735, Research Project 835-1, April 1978.
4-61. Varga, G. M., et al., "Applicability of the Thermal DeNOx Process
to Coal-Fired Utility Boilers," EPA-600/7-79-079, March 1979.
4-62. Castaldini, C., et al., "Technical Assessment of Thermal DeNOx
Process," EPA-600/7-79-117, May 1979.
4-54
-------
4-63. Faucett, H. L., et al., "Technical Assessment of NOX Removal
Process for Utility Application," EPA 600/7-77-127 (also EPRI
AF-568), NTIS-PB 276 637/AS, November 1977.
4-64. Mobley, J. D., "Flue Gas Treatment Technology for NOX Control,"
in Proceedings of the Third Stationary Source Combustion Sympos
Volume II. EPA-600/7-79-050b. February. 1979.
4-65. Cato, G. A., et al., "Field Testing: Application of Combustion
Modification to Control Pollutant Emissions from Industrial Boilers
- Phase II," EPA 600/2-76-086a, NTIS-PB 253 500/AS, April 1976.
4-66. Armento, W. G. and W. L. Sage, "The effect of Design Operation
Variables on NOX Formation in Coal-Fired Furnaces," Alliance
Research Center/B&W Pulverized Coal Combustion Seminar, June 19-20,
1973.
4-67. Blakeslee, C. E. and Burbach, H. E., "Controlling NOX Emissions
from Steam Generators," presented at the 65th Annual Meeting of Air
Pollution Control Association, June 1972.
4-68. Blakeslee, C. E. and Burbach, H. E., "NOX Emissions from
Tangentially-Fired Utility Boilers, Part II, Practice," AIChE
Symposium Series No. 148, Vol. 71, 1975.
4-69. Bartok, W., et al., "Systematic Field Study of NOX Emission
Control Methods for Utility Boilers," ESSO Research and Engineering
Co., Report No. GRU.4GNOS.71, December 1971.
4-70. Breen, B. P., "Combustion in Large Boilers: Design and Operating
Effects on Efficiency and Emissions," presented at the 16th
Symposium (International) on Combustion, Cambridge, Massachusetts,
August 1976.
4-71. Hunter, S. C., et al., "Application of Combustion Technology for
NOX Emissions Reduction on Petroleum Process Heaters," presented
at the 83rd National AIChE Meeting, Houston, Texas, March 1977.
4-72. "Standards Support and Environmental Impact Statement Volume 1:
Proposed Standards of Performance for Stationary Gas Turbines,"
EPA-450/2-77/017a, NTIS-PB 272 422/7BE, September 1977.
4-73. Breen, B. P., "Control of the Nitric Oxide Emissions from Fossil
Fueled Boilers," The Fourth Westinghouse International School for
Environmental Management, July 1973.
4-74. Bell, A. W., et al., "Nitric Oxide Reduction by Controlled
Combustion Processes," presented at the Western States Section, the
Combustion Institute, April 1970.
4-55
-------
4-75. Jain, L. K., et al., "State of the Art for Controlling NOX
Emissions, Part I, Utility Boilers," EPA-R2-72-072a, NTIS-PB 213
297, September 1972.
4-76. Mobley, J. D. and Stern, R. D., "Status of Flue Gas Treatment
Technology for Control of NOX and Simultaneous Control of NOX
and SOX," in Proceedings of the Second Stationary Source
Combustion Symposium: Volume III, EPA 600/7-77-073c, NTIS-PB 271
757/7AS, July 1977.
4-77. Goodwin, D. R., "Electric Utility Steam Generating Units.
Background Information for Proposed NOX Emissions Standards,"
EPA-450/2-78-005a, NTIS-PB 286 155, July 1978.
4-56
-------
SECTION 5
N0 CONTROL CHARACTERIZATION: EMISSION CORRELATION
Based on the general NO formation concepts discussed in
A
Section 4, it is evident that the basis for combustion modification NO
A
controls can be eventually traced back to chemical kinetics, turbulent
mixing, and heat transfer. However, incomplete understanding of the
combustion phenomena as well as insufficient data have prevented
researchers from fully characterizing NO formation and control in
X
utility boilers. Therefore, from a practical control point of view, it
would be valuable to relate NO emissions to gross overall furnace
A
parameters, such as surface heat release rates, for which there are
sufficient data from full scale utility boiler tests. These overall
furnace parameters can, in turn, be explained in terms of such local or
fundamental parameters as temperature and combustion regimes. This
information can be used to evaluate the basis and effectiveness of various
NO control techniques. Furthermore, it can aid in the cost-effective
A
design of new controls as well as in providing direction for future
research efforts since key boiler/burner design and operating variables
would have been identified.
However, it should be noted that the NO emission correlations
A
presented in this study are not intended to be definitive nor predictive.
Rather they are meant only to present general trends and highlight
important burner and boiler design and operating parameters. There is
insufficient data available to warrant in-depth interpretation.
This section begins with a review of previous NO modeling
A
efforts Section 5.1 then proceeds to a development of an interboiler NO
/\
correlation model in Section 5.2. The available data base of control
tests, spanning several test programs on many individual boilers, is also
discussed. Section 5.3 presents the results of applying this correlation
5-1
-------
model to some of the major boiler/fuel classifications, as data
permitted. The key boiler/burner design and operating variables and fuel
characteristics affecting NO formation are identified, and the results
A
are interpreted in the light of fundamental combustion theory and
knowledge of boiler operating practice. Finally, Section 5.4 summarizes
the findings.
5.1 PREVIOUS N0¥ MODELING EFFORTS
A
In recent years, numerous investigators have attempted to predict
NO formation in utility boilers, with varying degress of success. In
A
general, the studies have fallen into one of two broad categories: purely
empirical approaches to NO prediction based on available field test
A
data, and more fundamental approaches relying more heavily on heat and
mass transfer, flowfield, and combustion fundamentals. Some of the more
significant efforts are highlighted below.
Perhaps the most comprehensive of the fundamental models is that
developed by Combustion Engineering (CE) and reported by Bueters, et al.
(Reference 5-1) and Habelt and Selker (Reference 5-2). Combustion
Engineering spent several years developing a performance code to predict
NO emissions from tangentially fired utility boilers. Basically, their
A
model calculates the axial temperature/time history of combustion
products, then introduces NO generation via the Zeldovich mechanisms.
A
The calculation proceeds by dividing the furnace into "slices" and solving
the conservation equations for mass (including the combustion reactions),
momentum (flowfields), and energy in each slice.
Although this model is quite detailed and was designed to be
fundamental, it still relies heavily on actual data to determine the
numerous adjustable parameters incorporated. For example, although the
model accounts for vertical recirculation of the gaseous products, it
cannot determine the length of the recirculation region nor the length and
position of the heat release zone. In addition, the model needs as input
a "gas emissivity operator." Because these quantities depend upon the
operational mode of the boiler, e.g., load, fuel nozzle tilt, and excess
air level, data on the length and position of the heat release zone and
gas emissivity still must be correlated empirically against boiler
operational variables.
5-2
-------
In addition, since the CE model incorporates only Zeldovich
kinetics, it has inadequate provisions for predicting fuel NO
/\
generation. Thus, predictions for units burning high nitrogen content
fuels, such as high nitrogen oil and coal, are uncertain. The code
predicts NO to within 10 percent for gas-fired units, but only to
A
within 15 percent for oil-fired units, and is inadequate for coal-fired
units. Perhaps even the good agreement for gas-fired boilers is
fortuitous since the Zeldovich mechanism employed is still inaccurate in
that it does not account for superequilibrium N and 0 concentrations.
Nonetheless, the model has proved quite useful as a predictive and design
tool.
In another effort, Quan, et al., (Reference 5-3) attempted to
derive NO scaling relationships for industrial combustors by developing
A
dimensionless groups from the basic conservation equations for momentum,
energy, and species concentration. However, several simplifying
assumptions had to be made for their model. First, Quan, et al., employed
a single characteristic length to represent the composite of burner and
firebox dimensions important in NO formation. Second, the radiation
A
heat loss from the flame in the NO formation region was greatly
A
simplified. Also, the kinetic scaling neglected the effects of turbulent
mixing and hydrocarbon/NO coupling in NO formation. The results
J\ J\
predict NOX emissions as proportional to combustor characteristic
length, which is unrealistic. Thus, such simple scaling exercises are
limited in their abilities to model highly coupled phenomena such as NO
A
formation.
Quan, et al., (Reference 5-4) also attempted to predict NOY
X
formation by solving the governing partial differential equations for a
2-D combustor. 2-D elliptic flow code was used with approximate models
for turbulence, radiation heat transfer, and kinetics of hydrocarbon
oxidation and NO formation. Unfortunately, the results were
A
disappointing even for the idealized situation studied by the author.
N0¥ predictions were highly sensitive both to the finite difference grid
A
selected and to the form of the physical/chemical models used. Also, the
grid distribution required to compute the near burner events without large
truncation errors was impractical. Efforts by McDonald, et al., also on a
2-D combustor, encountered similar difficulties (Reference 5-5).
5-3
-------
Subsequent work by Quan and other investigators has been limited as
far as application to practical combustors. The phenomena to be modeled
are quite complex. First, it is thought that turbulent mixing can
dominate the combustion and NO formation rates in the primary flame
A
zone. Second, the hydrocarbon oxidation kinetics are strongly coupled to
NO formation in the near burner region, and simple Zeldovich kinetics
y\
are inadequate. Also, for oil or coal combustion, the droplet/particle
combustion rate is strongly coupled to luminous radiation heat transfer.
Although predicting these phenomena is possible in idealized lab-scale
situations, it is beyond current capability for practical combustors where
the mixing, chemistry, and heat transfer are strongly coupled. Thus,
predicting NO formation in practical combustors from solutions of basic
A
conservation equations is currently not possible.
In a more recent modeling effort, Dykema (References 5-6 and 5-7)
adopted a semi empirical approach employing a furnace model and a simple
Zeldovich mechanism for thermally generated NO , coupled with a largely
A
empirical model for the conversion of fuel bound nitrogen. Unfortunately,
the result was not very useful as a predictive tool. In fact, the model
often predicts wrong phenomena. For example, parametric variation of
excess air in Dykema's correlation showed a slight increase in NO with
A
decreased excess air. Also parametric variation of the location of
burners out of service rows indicated that middle rows out of service
would give the lowest NO emissions in gas-fired boilers, while bottom
A
rows out of service would be preferable in oil-fired units. Generally,
these conclusions do not concur with field experience.
In addition to the more fundamental models attempted above, several
empirical approaches to test data correlation have been pursued in recent
years. Bartok, et a!., (Reference 5-8) and Crawford, et al., (Reference
5-9) used a second order multiple regression analysis to correlate flue
gas NO concentrations with a limited number of boiler operating
variables. Because boiler design properties were not considered as
independent variables, Bartok1s and Crawford's analyses were restricted to
the individual boilers and loads studied.
Hoi linden, et al., (References 5-10 and 5-11) correlated NO
A
emissions with the same boiler operating variables used by Bartok. Again,
the empirical analysis was performed on an individual boiler basis, thus
5-4
-------
limiting the applicability of the correlation found. In addition, in the
earlier study (Reference 5-10), the author incorporated combustion staging
only as either "on" or "off" and did not allow continuous variation of the
staging parameters. In their later study (Reference 5-11), they included
only uncontrolled NO emissions in their analysis, thus limiting the
A
regression equations to predicting NO levels only under normal firing
A
conditions.
Cato, et al., (Reference 5-12) have also performed an empirical
study correlating uncontrolled NO emissions from industrial size
A
boilers with boiler design and operating variables. Although their data
base was quite diverse, covering more than a single boiler, their analysis
was limited to uncontrolled NOX emissions. In addition, industrial, not
utility, boilers were studied.
The above review was intended to highlight some of the difficulties
in attempting to model NO emissions on a fundamental basis as well as
A
to indicate a few of the practical limitations of reported empirical
models. The efforts of the investigators cited above, though admirable,
encountered monumental difficulties when attempting to relate real world
utility boiler combustion to fundamental parameters. The following
section presents the NO Emission Correlation Model, a multiple
A ,-
regression analysis developed in this study. The model is a crude attempt
to correlate emissions from a host of actual operating utility boilers to
some common burner and boiler design and operating variables. The intent
is by no means fundamental, but rather to highlight general trends.
5.2 N0¥ EMISSION CORRELATION MODEL
A
The formation of NOX in utility boilers is a complex and, at the
present time, imperfectly understood phenomena. Thus, although some
fundamental NO formation models are available, as discussed above,
A
especially for thermal NO , these models usually only relate NO
A A
generation to such fundamental combustion variables as stoichiometry flame
temperature and residence time of gases in the flame zone. Because the
flow in the furnace burner zone of an actual combustion source like a
utility boiler is extremely complicated, it is quite difficult to
determine the quantitative changes in these fundamental parameters
resulting from given operational or design changes. Thus, it is quite
5-5
-------
difficult to apply these fundamental models to N0x production in actual
combustion systems.
Consequently, a more empirical approach to NO correlation was
A
chosen in this study. From the point of view of the boiler operator or
designer, it is desirable only to establish which operational and design
variables are important in controlling NO emissions, and to obtain
A
estimates on how much NO levels will change with given changes in those
A
variables. Thus, a model based on multiple regression of existing data on
boilers would correlate NO emissions to specific boiler variables and
A
serve these desired ends. Such a model based on data from boilers under
actual operating conditions would be expected to reproduce the average
response of field boilers, in general, if the sample chosen for study is
representative of the field population. The following sections outline
the development of the correlation algorithm employed and the data base
used for analysis. It is noted that since the data base is limited, the
model should only be examined for general trends. Indeed it should not be
considered predictive, but interpolative.
5.2.1 Procedures
A large number of operational and design variables may be
postulated to affect NOV formation. A regression analysis can help in
A
screening these variables to determine which ones are most significant in
controlling NO . Moreover, as many of the variables are highly
A
intercorrelated, the analysis should incorporate a selection mechanism
whereby only the variables most strongly correlated with NO emissions
A
enter the regression and the other intercorrelated variables are
excluded. The model would, therefore, identify the important independent
variables and quantify the change in NO emissions due to specified
A
changes in the magnitude of the variables.
A second order regression model was used to fit the NO emission
A
data compiled as discussed below. A second order model was required
because NO formation mechanisms are usually nonlinear with the
A
fundamental parameters, therefore, a first order model does not correlate
the data very accurately. Also, as there is expected to be some degree of
interaction between the effects of the variables, cross product terms need
to be included in the model. A second order regression model includes
5-6
-------
quadratic and cross product terms. The predictive equation in such a
model takes the following form:
2 2
y = aQ + ajXj + ... + a^ + a22x2 + ... + &l2*i*2 + '"
where y is the response variable, in this case the NO emission level
A
x-j, X2 ... are the independent boiler variables and a , a,, ...
are the coefficients to be determined. The coefficients are determined by
obtaining the best fit to the data as defined by minimizing the sum of the
squares of the distance between the predicted values and the data (least
squares fit).
By choosing a second order model, it was assumed that terms of
order higher than quadratic are not important in the analysis, and that
exclusion of cross product terms with more than two variables does not
significantly affect the accuracy of the prediction. A third order model
could have been constructed in a fashion similar to the second order
model, but the number of terms required would make the analysis too long
and unwieldy.
Some of the assumptions underlying the regression analysis are that
the distribution of each variable is normal, and their joint distributions
are also normal. Also, all variables must be homoscedastic; that is, the
variance of each variable must be uniform over the sample space. Other
assumptions inherent in the analysis are that: (1) no important boiler
variables were overlooked by the investigators, (2) the data selected for
use in the regression were representative of typical boilers and operating
conditions, (3) the data reported were accurately measured and reported,
and (4) the data from different boilers and different tests all had
comparable errors of measurement.
In one case, a second order model did not yield the desired degree
of precision in predicting data. In that instance, a logarithmic model
was used. The logarithm of NO emissions are correlated linearly to the
logarithm of the dependent variables, and the predictive equation then
takes the form:
1 og y = aQ + a^ 1 og x, + a^ 1 og X2 + a~ 1 og x., +
5-7
-------
which is equivalent to
al a2 a3
jr — C X* Art 4 • • •
The assumptions underlying a logarithmic model are similar.
Specifically, the logarithms of the variables are assumed normally
distributed and the variances based on the logarithms of the variables are
homoscedastic. It is very unlikely that these conditions will ever be met
in practice. But, the procedure can be justified as useful in screening
variables and providing guidelines on the expected magnitude of changes in
NO emissions due to changes in boiler variables.
A
The multiple regression analysis was performed using the stepwise
regression procedure (References 5-13 and 5-14). In this procedure
variables are introduced into the correlation one at a time in order of
most significant correlation. In the specific procedure employed, a first
order linear multiple regression analysis, using the stepwise procedure,
was first carried out to identify the seven most important variables.
Second order analysis was then performed.
As each new variable was introduced into the regression, the
multiple correlation coefficient of the regression up to that point was
calculated. The stepwise procedure was terminated when the increase in
the multiple correlation coefficient with addition of new variables became
sufficiently small. As the multiple correlation coefficient is directly
related to the square of the variation explained by regression, this
criterion tends to inhibit variables which do not contribute substantially
to decreasing the standard error of the estimate from entering the
regression.
Multiple regression equations obtained in this manner were examined
to check whether NO emissions predictions were within desired degrees
/\
of accuracy. If sufficient precision could not be obtained with a second
order model, a logarithmic model was employed. Thus, in one case the
logarithmic model yielded better predictive correlations than the second
order model.
5.2.2 Data Base
For emissions correlations through the procedure described above,
uncontrolled and controlled N0y combustion data were obtained from a
A
5-8
-------
total of 61 boiler firing type/fuel combinations. Table 5-1 breaks out
the test data combinations employed. Data gathering was limited to
tangential, horizontally opposed and single wall firing types as these
were the most extensively tested. Fortunately, a representative
population was treated because these firing types represent approximately
87 percent of the current installed utility steam generating capacity.
NO emissions from turbo furnace, cyclone, vertical, and stoker furnaces
A
were not analyzed in depth because published emissions data from these
boilers were very limited.
Emissions data were assembled from emissions field test programs
sponsored by EPA and by several private utility companies. In many cases,
previously unreported test data were incorporated. References 5-6, 5-8
through 5-11, and 5-15 through 5-19) supplied the test data assembled.
Several specific units were tested in more than one program. In
these instances each individual program was considered as a separate
unit. Therefore, the totals in Table 5-1 include boilers that were tested
more than once during different programs. The additional sets of
emissions data from these units were considered valuable in the present
analysis because different levels of NO control were usually achieved
A
from test program to program. In addition, baseline and controlled boiler
operating conditions varied slightly between separate test programs on the
same boiler, thus providing more representative average operating
conditions and emission levels.
Table 5-2 itemizes the actual test points incorporated into the
emissions data base. The total of 563 tests represents approximately 25
percent of the total number of tests reported in the various test
programs. The table shows that the largest number of selected tests,
comprising 54 percent of the data, were on coal-fired utility boilers.
Gas-fired and oil-fired boilers were studied in 24 and 22 percent of the
test points, respectively. The single most studied category was
tangential coal-fired steam generators with 147 individual tests.
Six single control techniques were considered in the analysis:
• Low excess air (LEA)
• Overfire air (OFA)
• Biased burner firing (BBF)
• Burners out of service (BOOS)
5-9
-------
TABLE 5-1. FIELD TEST PROGRAM DATA COMPILED
Fuel
Coal
Oil
Natural Gas
Total
Firing Type
Tangential
13
2
1
16
Opposed Wall
6
7
8
21
Single Wall
10a
7
7b
24
Total
29
16
16
61
alncludes two wet bottom furnaces
blncludes one unit originally designed for coal firing with a
wet bottom furnace
5-10
-------
TABLE 5-2. INDIVIDUAL TEST POINTS CORRELATED
Firing Type
Tangential
Opposed
Wall
Single Wall
Tangential
Opposed
Mall
Single Wall
Tangential
Opposed
Wall
Single Wall
All Boilers
Fuel
Coal
Coal
Coal
Oil
Oil
011
Nat gas
Nat gas
Nat gas
All fuels
Baselineb
21
8
18
1
6
4
1
7
5
71
Single Controls
LEAC
29
11
23
--
5
6
1
9
4
88
oscd
46
11
29
1
11
f,
--
18
9
130
FGRe
--
7
--
—
2
4
2
--
2
17
Low
Loadf
24
7
19
1
7
8
2
13
7
88
Combined Controls3
Low load
+ OSC
27
S
19
1
7
6
1
13
7
86
Low Load
+ FGR
-
1
--
1
5
10
5
3
3
28
OSC +
FGR
—
2
--
--
2
10
1
3
4
22
Low Load +
OSC + FGR
—
—
--
1
11
8
-
8
5
33
Total
147
52
108
6
56
61
13
74
46
563
Low excess air also generally employed
Baseline = no controls applied; boiler load near or at maximum rating; excess air at
normal or above normal settings
LEA = low excess air setting
OSC = off stoicMometrlc combustion (Includes: biased burner firing, burners out of
service, overflre air)
FGR « flue gas recirculatlon; generally includes low excess air setting
Load less than 80 percent of maximum continuous rating (MCR)
5-11
-------
• Flue gas recirculation (FGR)
• Load reduction
Data on applying other NO reduction techniques, such as water
/\
injection, mill fineness setting, and reduced air preheat, were
occasionally reported. However, these were not included in correlations
because data were limited, and these techniques are considered of lesser
priority for study in the present analysis. At the time this correlation
analysis was performed, little boiler data from coal-fired units
retrofitted with low NO burners were available (References 5-15, 5-20,
X
and 5-21). These data were too limited for statistical treatment. In
addition, the correlation algorithm derived included no single variable
able to distinguish between conventional circular and dual register burner
designs.
Virtually all tests reported in which single controls or
combinations of these controls were applied in various degrees were
included in the data base. However, test points were excluded from the
analysis if they failed two general selection criteria:
t Were NO reductions representative of the unit tested?
A
• Were other boiler operating parameters, e.g., register settings
held nominally constant within normal ranges?
For example, test points were rejected as failing the first
criterion if the test crew reported inconsistencies between these points
and the remainder of the test program. Test data describing lowest NO
A
levels achieved on a given unit were in general included. However, if the
test report noted that operation at these levels was deemed unsafe by
plant personnel, the data were rejected.
Similarly, test data were excluded if other boiler operating
parameters, not explicitly treated in the correlation algorithm, were not
held to nominally constant values. For example, several test series
investigated the effects of burner register settings on NO emissions.
A
Changing register setting causes variations not only in burner swirl but
also in airflow through the burner. However, since register setting was
not treated explicitly in the emissions correlation model, these test
series were excluded from the data base. Only tests with "normal" and
nominally constant register settings were included. Similarly, tests on
5-12
-------
tangential units which varied burner or overfire air port tilt were
excluded.
Finally, for BOOS tests, only data taken with burners in the top
rows removed from service were included in the correlation data base. All
other BOOS patterns were disregarded.
Based on the above, only about 25 percent of the total reported
test data were suitable for inclusion in the assembled correlation data
base. Of the data excluded, much of it was due to insufficient
information, e.g., the boiler's heat release rate could not be obtained.
As Table 5-2 shows, a total of 71 baseline tests were chosen to represent
normal, uncontrolled boiler operating conditions. For comparisons of the
effectiveness of the individual NOY controls studied, 88 LEA firing
^
tests, 130 off stoichiometric combustion (OSC) tests, 17 F6R tests, and 88
load reduction (load at less the 80 percent of unit MCR) tests were
included. Off stoichiometric combustion in its various applications (BBF,
BOOS, and OFA) was by far the most extensively tested combustion
modification technique. In contrast, test data for FGR as a single N0x
control were insufficient for a good statistical analysis of the effects
of this technique.
The purposes of performing the NO correlation analyses were
J\
twofold. Of course, good statistical evaluations of the effectiveness of
commonly applied combustion controls singly and in combination, were
desired. But regression relationships between NO emissions and more
rt
fundamental combustion, boiler design, and operating parameters were also
sought through the model, to highlight general trends.
Thus, specific data on a set of design and operating variables
associated with each test point in the data base were needed to allow
correlation relationships to be obtained. These correlation variables
used in the analyses fell into three categories:
• Boiler operating variables
• Boiler design variables
• Fuel properties
Boiler Operating Variables
Correlation parameters in this category are the macroscopic
combustion variables describing boiler operation, which are altered when a
5-13
-------
combustion control is applied. The specific variables used in the present
analysis included:
• Overall furnace fuel/air stoichiometry
t Stoichiometry at active burners
• Percent flue gas recirculated
0 Firing rate (as percent MCR)
t Percent burners firing
• Heat input per active burner
For example, an LEA application can be trivially considered as a
change in overall furnace stoichiometry. Similarly, applying OFA alters
burner stoichiometry and perhaps also overall stoichiometry if higher
overall excess air levels are required. Burners out of service firing
elicits similar changes while also altering percent burners firing. Flue
gas recirculation is applied by changing the percent gas recirculation
variable. Load reduction is obviously accompanied by changes in firing
rate and oftentimes heat input per active burner and overall stoichiometry.
Boiler Design Variables
The boiler design variables considered in the regression analysis
included:
• Nameplate maximum continuous rating (MCR)
• Volumetric heat release rate
• Surface heat release rate
• Heat input per active burner
• Number of burners
• Number of furnaces
• Number of division walls
Table 5-3 lists the ranges and average values encountered for each of
these variables. Boilers of the same fuel and firing type were grouped
together in the correlation analysis.
Data on burner zone surface heat release rate were generally
unavailable for the test reports cited. This is unfortunate since NO
A
emission levels are expected to be stronger functions of this variable
than of the more global overall heat release rate (Reference 5-22). In
fact, adjusting this variable alone allows a significant degree of NO
control for gas- and oil-fired boilers. Still, use of the overall
5-14
-------
TABLE 5-3. BOILER DESIGN VARIABLES CONSIDERED
Equipment
Typ«
Tan-
gential
Opposed
Mill
Single
Hill
All
Boilers
Fuel
CM!
011
Natural
6«
CM!
0<1
Natural
6n
Coil
on
Nituril
Git
Coil
011
Sis
All
FutU
Ma>UH«
Cont 1 nuous
Rating, m*
Range
125-800
(6-320
920
218-820
220-410
220-600
100-340
80-250
80-315
100-820
66-480
80-600
66-820
Avenge
430
193
120
580
320
350
200
190
200
430
234
290
310
Volunetrtc Heit
Release Rite. kU/«3
HO* Btu/ft3-hr)»
Range
116-159
(11-15)
289-310
(28-30)
289
(26)
134-178
(13-17)
255-297
(25-29)
152-287
(15-28)
138-242
(13-23)
198-299
(19-29)
181-282
(18-27)
116-242
(11-23)
198-310
(19-30)
152-289
(15-28)
116-310
(11-30)
Avenge
139
(13)
300
(29)
289
(28)
157
(15)
270
(26)
254
(25)
196
(19)
250
(24)
232
(22)
164
273
(26)
258
(25)
232
(22)
Surfice Heit
Release Rite, ui/m>
(103 Btu/ft'hr)'
Range
78-466
(25-148)
349-541
(111-172)
541
(172)
228-312
172-99)
204-625
(65-199)
204-604
(65-192)
110-455
(35-145)
248-778
(79-247)
248-324
(79-103)
78-466
(25-148)
204-778
(65-247)
204-604
(65-192)
78-778
(25-247)
Average
228
(72)
443
(141)
Stl
(172)
259
(82)
422
(134)
395
(126)
236
(75)
343
(109)
265
(84)
241
(76)
402
(128)
400
(127)
348
(110)
Heat Input per
Active Burner,
W (10b Btu/hr)t>
Range
13-75
(45-261)
11-38
(38-133)
38
(133)
33-86
(115-299)
26-81
(90-282)
2S-63
(87-219)
21-55
(73-191)
22-43
(77-150)
21-43
(73-150)
13-86
(45-299)
11-81
(38-282)
21-63
(73-218)
11-86
(38-299)
Average
40
(139)
24
(85)
38
(133)
49
(170)
44
(153)
42
(146)
30
(105)
35
(122)
33
(115)
40
(138)
34
(118)
38
(132)
37
(129)
Total No.
of Burners
Range
16-64
8-24
24
20-54
12-24
12-36
16-24
12-24
12-24
16-64
8-24
12-36
8-64
fyptcil
Umber
32
16
24
34
20
22
18
14
16
28
28
20
22
Range In
NMber of
Furnaces
1-2
1-2
2
1-2
1
1
1-2
1-2
1-2
1-2
1-2
1-2
1-2
Range In
Nu»ber of
)1«ls1on
Halls
0
0
0
0-1
0-1
0-1
0-1
0-1
0-1
0-1
0-1
0-1
0-1
•Electrical output
"At lutaHX continuous rating
5-15
-------
parameter, for which data were generally available, did allow reasonably
good NO correlation.
X
Heat input per active burner could also be considered a boiler
operating parameter but was grouped here with design variables for
convenience. As Table 5-3 shows, heat input per active burner varied in
the data base from 11 to 86 MW. Horizontally opposed boilers in general
recorded the highest values, probably because of the generally greater
unit size, hence burner size, of these units. The value of this variable
can be changed when applying biased burner firing, burners out of service,
and load reduction. It should be noted, though, that load reduction can
be accomplished by totally removing burners from service, in which case
heat input per active burner could remain unchanged.
The number of furnace division walls was introduced as a design
variable as part of a crude attempt to account for gross changes in
furnace mixing patterns and burner zone surface heat release rate. The
use of dividing water walls allows boiler designs with smaller surface
heat release, at a relatively constant volumetric heat release rate.
Ideally, effects of division walls on NO emissions would be picked up
A
in the regression through these two variables directly. However, since
these water walls generally separate the furnace only part of the way up
to the convective passes, they also affect burner zone heat release rate
at constant overall heat release rate. In addition, gas mixing patterns
are altered from units of similar size, but of divided design, with
corresponding effects on NOV emissions.
A
For similar reasons, the number of furnaces was included as a
correlation variable. Twin furnace design is most prevalent in larger
tangentially fired units, though it is occasionally found with other
firing types. It should be noted here, though, that two potentially
important boiler design variables known to affect NO emission levels
A
were not included in the correlation analysis because the data were not
available. These are burner spacing and distance between the top burner
level and the OFA ports. Both of these variables can have significant
effects on NO production in a given unit by affecting flame
A
interactions, gas mixing, and heat absorption in the burner zone.
Furthermore, in BOOS and OFA applications, these variables will affect
first and second stage residence time and separation. Unfortunately,
5-16
-------
these data were unavailable for most units tested in the test reports used
in compiling the emissions data. Thus, they were not included in the
analysis.
Fuel Properties
The fuel variables considered in the correlation analysis were
nitrogen content, moisture content (coal only), and heating value.
Unfortunately, the information on these fuel properties was not always
available for each test point. In cases where fuel analyses were sparse
at the individual run level, they were assumed constant throughout a
series of tests on a specific boiler.
Table 5-4 lists the fuel properties considered in the present
analysis and their average values. Fuel nitrogen for all coals tested
varied by a factor of 3 from 0.62 to 1.84 percent by weight. Even though
this represents a significant range in fuel nitrogen content, NO
A
emissions were found not to be significantly affected by fuel nitrogen
content.
Moisture content of coals also varied significantly from 1.14 to
36.4 percent. This is to be expected since coal types used in various
tests varied from the low moisture content Eastern bituminous to the high
moisture content Western sub-bituminous and lignite coals. Coal moisture
content was also not found to affect NO emissions.
5.3 NOY EMISSION CORRELATION RESULTS
rt
In this section, the results of applying the correlation model to
the data base of test results are discussed. Key boiler design and
operating variables, burner characteristics, and fuel properties which
affect NO formation are identified. The basis and effectiveness of the
A
various NO control techniques are reviewed. These results are further
A
discussed in the light of fundamental combustion principles and boiler
operating practice.
The major boiler firing types, tangential, single wall, and opposed
wall fired, with the principle fuels, coal, oil, and gas were treated.
However, tangential oil- and gas-fired boilers were not considered in the
correlation study as the data were insufficient for a statistical analysis.
5.3.1 Tangential Coal-Fired Boilers
A multiple regression analysis was carried out on tangential
coal-fired boilers. Data were analyzed for 147 tests carried out on a
5-17
-------
TABLE 5-4. PROPERTIES OF FUELS FIRED
tn
i
CD
Equipment Type Fuel
Coal
Tangential
Oil
Coal
Opposed Wall Q11
Coal
Single Wall
Oil
Coal
All Boilers
Oil
Fuel Nitrogen, Percent by Weight
Range Average
0.6-1.6 1.2
0.3-0.6 0.5
1.0-1.8 1.3
0.2-0.4 0.3
0.8-1.5 1.3
0.2-0.3 0.3
0.6-1.8 1.3
0.2-0.6 0.3
Fuel Moisture, Percent by Weight
Range Average
3.4-31.9 12.5
1.1-36.4 7.2
4.5-28.9 8.9
1.1-36.4 9.5
Heating Value3, MJ/kg (103 Btu/lb)
Range Average
19.0-32.3 27.2
(8.19-13.9) (11.7)
NA NA
24.4-31.6 28.3
(10.09-13.57) (12.2)
43.7-45.8 44.2
(18.8-19.7) (19.0)
23.0-32.8 28.8
(9.9-14.1) (12.4)
43.7-45.6 44.7
(18.8-19.6) (19.2)
19.0-32.8 28.1
(8.19-14.1) (12.1)
43.7-45.8 44.4
(18.8-19.7) (19.1)
Dry basis
-------
total of 13 boilers. The data included 21 tests performed under baseline
conditions with the rest conducted under low NO conditions. Low NO
^ A
techniques tested included LEA, OSC, low load and a combination of low
load and OSC.
For tangential coal-fired boilers, the following equation
correlates the data with a correlation coefficient of 0.87, i.e.,
75 percent of the variance is explained by the regression:
y = 184 + 1.09 x 10"7(X1)(x2) - 1.67 x 10"5(/1) +
2.49 x 10"6(x3)(x4) + 6.54 x lO"14^)2
where
y = NO emissions (ppm dry at 3 percent 09)
A £
x, = Heat input per active burner (W)
x« = Stoichiometry to active burners (percent stoichiometric air)
2
*3 = surface heat release rate (W/m )
x. = Furnace Stoichiometry (percent stoichiometric air)
From the regression equation it is seen that burner Stoichiometry
and heat release rate are the most important parameters governing NO
/\
emissions in these boilers. This is in agreement with fundamental
combustion principles as Stoichiometry affects both thermal and fuel NO
y\
while heat release should mainly affect thermal NO . The equation
/\
indicates that, in general, NO emissions will be reduced by decreasing
^
both Stoichiometry and heat release, which is also consistent with theory.
A graphical representation of how NO emissions vary with surface
A
heat release rate and burner Stoichiometry is shown, in Figure 5-1. The
parametric lines in the figure are generated from the regression equation
by allowing surface heat release to vary while fixing the burner
Stoichiometry at the values shown beside the curves. All other
variableswere held constant at their mean values, except for furnace
Stoichiometry which was taken equal to burner Stoichiometry for non-OSC
operation, and fixed at some lower limit (typically 120 percent) for OSC
operation. The parametric lines are seen to match reasonably closely with
the data points. The reduction in NOX emissions with reduced burner
Stoichiometry and surface heat release rate is clearly evident from the
figure.
5-19
-------
en
i
ro
o
7001
600J
CM
c
»*
CO
§.
a.
tn
c
o
100
Surface heat release rate (kW/m2)
25 50 75 100 125 150 ~175"
Surface heat release rate (106 Btu/hr-ft2)
140
640
200
Stoichiometry to active
burners (percent)
O 140
CD 120
X 100
Z 80
Figure 5-1. Effect of surface heat release rate and burner Stoichiometry
on NOX from tangential coal-fired boilers.
-------
Figure 5-2 shows another graphical representation of NO
emissions variation with burner stoichiometry and heat input per active
burner. The parametric curves were generated from the regression equation
in a manner similar to that explained above for Figure 5-1. Again, the
effect of decreasing NO with decreasing burner stoichiometry is clearly
rt
seen. Decreasing heat input per active burner, however, seems to have a
mixed effect on NO. For burner stoichiometry above 120 percent, NO
X X
emissions decrease with reduced heat input, consistent with earlier
discussion. Note that burner stoichiometries above 120 percent generally
preclude OSC operation. For burner stoichiometries about 100 percent, the
NO emissions sometimes actually decrease with increasing heat input per
/\
burner. This can be explained by noting that the data points for burner
stoichiometries at about 100 percent or lower include tests with BOOS
operation. In such cases, increasing heat input per active burner is
tantamount to increasing the degree of off stoichiometry, as fuel flow to
active burners must be increased under BOOS operation to maintain load.
Under these circumstances the NO emissions should decrease with
A
increasing heat input per burner, and that is precisely what is observed.
The above example points out the need to be very careful in
interpreting the results of the regression analysis. The equations are
valid only within the range of conditions of the original data base,
sothat any generalizations should be made with caution. It should also be
noted that the independent variables are often related to each other
within certain ranges of operation. For example, LEA operation, without
OSC, will influence both burner as well as furnace stoichiometry. Also,
low load, without BOOS, will affect both the surface heat release rate and
heat input per active burner. And, as pointed out earlier, BOOS operation
will affect burner stoichiometry and heat input per active burner.
It is seen from the regression equation and from Figures 5-1 and
5-2 that the most effective operational technique for NO control on
tangential coal-fired boilers seems to be reduction of burner
stoichiometry. This can be accomplished to a certain extent with LEA and
to a greater extent with OSC. Lower surface heat release rates also
result in lower NOX emissions. In addition, lower-heat release per
burner also tends to reduce NOX emissions, at least when not operating
under BOOS firing.
5-21
-------
700
PO
ro
o
S*
£ 500
-
§.
Q.
j; 400
c
o
x 300
o
200
100
10 20 30 40 50 60 70
Heat input per active burner (MW)
, 120
80 90
Stoichiometry to active
burners (percent)
(D 140
m 120
X 100
Z 80
40
80
120
160
200
240
Heat input per active burner (10 Btu/hr)
280
Figure 5-2. Effect of heat input and burner stoichiometry on NOX from
tangential coal-fired boilers.
-------
5.3.2 Horizontally Opposed Coal-Fired Boilers
The multiple regression analysis was also applied to horizontally
opposed coal-fired boilers. Fifty-two tests on six boilers were selected
for the analysis. The data included tests performed under baseline
conditions as well as LEA, OSC, low load, and a combination of low load
and OSC. In addition, data from a boiler tested with FGR were included.
Some test data on a combination of low load and FGR, and OSC and FGR were
also included.
The regression analysis yielded the following equation, which has a
correlation coefficient of 0.91, i.e., 83 percent of the variance is
explained by the expression:
y = -471 + 5.38(x1) + 4.24 x 10"6(x2) + 7.41(x3)
-5.84(x4) - 6.64 x lO1^) + 2.46 x lO^Xg)
where
y = NOX emissions (ppm dry at 3 percent 02)
x-, = Stoichiometry to active burners (percent stoichiometric air)
x2 = Heat input per active burner (W)
Xo = Number of burners firing
x^ = Flue gas recirculation (percent)
Xj- = Number of division walls
Xg = Excess oxygen (percent)
The regression equation indicates that NO increases with
n
increasing Stoichiometry to burners, heat input to the burners, the number
of burners firing, and overall excess oxygen, whereas NO decreases with
n
increasing flue gas recirculation and number of division walls. These
results are, in general, in agreement with past experience and theoretical
considerations. Burner Stoichiometry and overall excess air are known to
have a large positive correlation with NO formation. Increased heat
^
input to burners would also be expected to increase NOX emissions. The
positive correlation of a number of burners firing with NO probably
A
stems from many factors. Larger boilers produce significant NO and
usually have more burners. At partial loads NOX generation is reduced
and so are the number of active burners. Finally, with BOOS, the number
5-23
-------
of active burners decreases and so does NOX< The number of division
walls is negatively correlated with NOX. This is most likely due to the
increased surface area available for heat transfer with consequent
lowering of flame temperatures.
The correlation of NOY emissions with FGR is interesting and is
A
shown in Figure 5-3 with burner stoichiometry as a parameter. Although
the data are relatively sparse, the statistical correlation do point to a
negative trend of NO emissions with increasing FGR. It should be noted
A
also that the decrease in NOX due to 20 percent FGR is approximately the
same as the decrease in NO with a 10 percent reduction in excess air at
A
the burners. FGR is known to inhibit thermal NO , whereas OSC controls
A
both thermal and fuel N0y. OSC is therefore expected to be a more
effective NO control technique than FGR, in agreement with experience.
A
The effect of heat input per active burner on NO emissions is
A
shown in Figure 5-4, again with burner stoichiometry as a parameter. The
data scatter is rather large and very few data points are available for
the substoichiometric region. Nevertheless, it is seen that, in general,
increasing heat input per active burner increases NO . It also
A
indicates the'influence of burner stoichiometry on NO emissions. It
/\
should be reiterated here that the data base is limited, and that data
from different manufacturers were incorporated together. Hence design
differences between burners are masked in the correlations. Thus a large
burner (high input) does not necessarily produce high NO . Indeed the
A
new burners coming onstream today have designs that limit air/fuel mixing
in the burner zone and hence limit NO production. In other words,
A
burner design can overcome the tendency of higher NO with increasing
A
heat input (Reference 5-24).
From the regression analysis, it can be seen that for horizontally
opposed coal-fired boilers, reducing burner stoichiometry is a very
effective means for controlling NOX emissions. LEA reduces burner
stoichiometry, but OSC must be employed if large reductions up to or below
the stoichiometric level are desired. FGR also reduces NO but to a
A
lesser extent than reduced burner stoichiometry. The implications for
boiler design from this study are that increased cooling surface and
decreased heat input per burner tend to decrease NO emissions. But as
A
noted above, burner design, though not included in the correlation, is
5-24
-------
en
i
ro
en
400
8 12 16 20 24
Fuel gas recirculatlon (percent)
28
Stoichiometry to active
burners (percent)
QJ 140
CD 120
A 100
X 80
Figure 5-3. Effect of FGR and burner Stoichiometry on NOX from
horizontally opposed coal-fired boilers.
-------
1000
PO
CTl
20
30 40 50 60 70 80
Heat input per active burner (MW)
80 120 160 200 240 280
Heat input per active burner (10 Btu/hr)
320
100
360
Stoichiometry to active
burners (percent)
E 140
CD 120
A, 100
X 80
Figure 5-4. Effect of heat input and burner Stoichiometry on NOX from
horizontally opposed coal-fired boilers.
-------
equally if not more important than burner heat input. The regression
model can be used to illustrate general trends in the change of NOX
emissions with design or operational changes. One should be very careful,
however, not to extrapolate the equation beyond the range of data on which
it was correlated.
5.3.3 Single Wall Coal-Fired Boilers
The multiple regression analysis was applied to single wall
coal-fired boilers with data from 86 tests performed on eight boilers.
The data included tests under baseline and low NO conditions. Low
NO techniques included LEA, OSC, low load, and a combination of low
^
load and OSC.
The regression analysis correlated the data with a correlation
coefficient of 0.896, i.e., 80 percent of the variance was explained by
the regression. The equation which best correlated the data was:
y = -140 + 1.98 x 10"1(x3)(x2) + 6.95 x 10"5(x1)(x5) + 4.5 x 10"6(x1)(x2)
+ 7.57 x 10"8(x4)(x2) - 1.02 x 10"11(x1)(x4)
where
y = NO emissions (ppm at 3 percent 0~)
x ?
x, = Surface heat release rate (W/m )
x2 = Stoichiometry to active burners (percent stoichiometric air)
x-j = Number of burners firing
x4 = Heat input per active burner (W)
Xj- = Furnace excess oxygen (percent)
This regression equation is complex with many variables appearing
several times in conjunction with other variables. The Stoichiometry
tothe active burners has a marked large positive correlation with N0x
emissions. The number of firing burners, the heat input per active
burner, and the furnace excess oxygen are all also positively correlated
with NO emissions. These positive correlations are consistent with
theoretical considerations and past experience. Stoichiometry to active
burners and overall excess oxygen have been shown to have a marked effect
on thermal and fuel NOY generation. The heat input per burner which is
^
related to the flame intensity and, hence, peak temperatures should affect
5-27
-------
thermal NO emissions. The number of firing burners increases with boiler
A
size, high load, and absence of BOOS firing, all of which tend to increase
NO emissions. The effect of surface heat release rate is not
A
straightforward and is discussed further below.
Figure 5-5 is a plot of NO emissions versus surface heat release
A
rate with burner stoichiometry as a parameter. Here the trends are
consistent with expectations based on previous correlations. NO
A
emissions tend to increase with increasing surface heat release rate and
burner stoichiometry. Figure 5-6 shows the variation of NO emissions
A
with heat input to active burners and burner stoichiometry. Again the
trends are consistent with expectations. NO emissions tend to increase
A
with increasing heat release per burner and increasing stoichiometry. The
data are sparse for higher heat release rates, so that predictions at those
values may not be very accurate. Nevertheless, the trends should be
correctly predicted.
From the regression analysis, it is seen that burner stoichiometry
again has the greatest effect on NO emissions from single wall coal-fired
A
boilers. LEA can be employed to decrease burner stoichiometry to a certain
extent. OSC should be employed if further reduction is desired.
Implications for boiler design are that decreasing heat input per burner can
reduce NO emissions. But, as discussed in Section 5.3.2, burner design,
A
a variable not incorporated here because of data limitations, can
predominate over heat input. Finally decreasing heat release rate, all
other factors equal, generally does reduce NO . The regression equation
A
can be used to estimate trends in NO emissions due to design or
A
operational changes. However, as most of the data on which the correlation
is based are confined to a small range, care should be exercised when making
numerical predictions.
5.3.4 Horizontally Opposed Oil-Fired Boilers
The data base for the multiple regression analysis on horizontally
opposed, residual oil-fired units was relatively good. The total of 56
test points from the seven boilers tested gave more than 1 test point for
each control and combination of control methods considered. The tests
included baseline, low excess air, off stoichiometric combustion, flue gas
recirculation, load reduction, and combinations of these control methods.
5-28
-------
1000-
20
40
60 80 100 120 140 160
Surface heat release rate (106 Btu/hr-ft2)
Stoichiometry to active
burners (percent)
0) 140
Qj 120
X 100
Z 80
Figure 5-5. Effect of surface heat release rate and burner Stoichiometry
on NOX from single wall coal-fired boilers.
-------
140
en
i
co
O
30
60
90
120
150
180
210
Heat input per active burner (10 Btu/hr)
MW
240
Stoichiometry to active
burners (percent)
a 140
O 120
A 100
X 80
Figure 5-6. Effect of heat input per active burner and burner stiochiometry
on N0x from single wall coal-fired boilers.
-------
The correlation explained variations in NO emissions to within
rt
80 percent. The significant parameters were found to be firing rate, number
of burners firing, stoichiometry to active burners, number of division
walls,and furnace stoichiometry. The second order multiple regression
equation which best correlated the data, with a correlation coefficient of
0.90 was:
y = -228 + 1.05 x 10'1 (xi) (x2) + 7.23 x 10-3(x3)2
- 1.30 (xj) (x4) + 2.392(x5)
where
y = NO emissions (ppm dry at 3 percent 0~)
/\ w
x-i = Firing rate (percent)
Xp = Number of burners firing
X3 = Stoichiometry to active burners (percent stoichiometric air)
x, = Number of division walls plus one
x5 = Furnace stoichiometry (percent stoichiometric air)
The variable dependencies were not unexpected. Reducing the firing
rate lowers the volumetric heat release rate. Thus, reduced heat release
rate will lower the bulk gas temperature in the furnace resulting in reduced
thermal NO formation. An increase in the number of active burners for a
given heat release rate should also increase NO emissions, as more active
^
burners firing will result in more thorough mixing of fuel and air in the
peak flame temperature regions. The stoichiometry to active burners and the
overall furnace stoichiometry influence NO formation in that higher
oxygen concentrations in the peak flame temperature regions will increase
NO formation. As shown in Figure 5-7, increasing either boiler load or
A
burner stoichiometry increases NO emissions. Also, as shown in
Figure 5-8, increasing either the number of burners firing or the burner
stoichiometry increases NO emissions. Finally, furnace division walls
add heat transfer surface to the furnace. The increased surface will result
in greater heat transfer from the furnace gases thus lowering the furnace
bulk gas temperature. This will reduce thermal NO formation, and thereby
A
lower NO emissions.
/\
5-31
-------
en
i
CO
r\>
o>
430
380
CM
O
^ 330
-------
450 _
140
en
i
CO
co
CM
O
co
rO
2?
I/I
c
o
•r~
in
i/i
150
12 16 20
Number of burners firing
24
120
100
80
Stoichiometry to active
burners (percent)
D 140
0 120
& 100
X 80
32
Figure 5-8. Effect of burner variables on NO from horizontally opposed oil-fired boilers.
/\
-------
Expected effects of flue gas recirculation were not picked up in the
correlation. One reason may be the limited data on the effect of FGR as a
single NO control. Second, although more data were included on FGR in
/\
combination with other techniques, the effectiveness of FGR may be
diminished when used in conjunction with other control methods.
5.3.5 Single Wall Oil-Fired Boilers
The data base for the multiple regression analysis for NO
A
reduction in single wall oil-fired boilers consisted of 61 test points from
seven boilers tested. This gave a minimum of four tests for each control
and combination of control methods considered. The tests included baseline,
low excess air, off stoichiometric combustion, flue gas recirculation, load
reduction, and combinations of these control methods.
The correlation explained variations in NO emissions to within
A
68 percent. The significant parameters were found to be volumetric heat
release rate, stoichiometry to active burners, the difference between
furnace stoichiometry and burner stoichiometry, heat input to active
burners, and number of burners out of service. The second order multiple
regression equation best explaining the data, with a correlation coefficient
of 0.83, was:
y = 173 + 2.28 x 10"5 (XjMXg) - 1.91 x 10"3 (x^ + 6.18 x 10"8(x3)(x4)
- 9.41 x 10'7 (x4)(x5) + 3.60 x 10"14 (x4)2
where
y = NO emissions (ppm dry at 3 percent 0,)
3
Xj = Volumetric heat release rate (W/m )
*2 ~ Stoichiometry to active burners (percent stoichiometric air)
x^ = Furnace stoichiometry minus burner stoichiometry (percent)
x^ = Heat input per active burner (W)
x5 = Number of burners out of service
The importance of these parameters was expected. A reduction in
the volumetric heat release rate lowers bulk gas temperature in the
furnace. This results in reduced thermal NOV formation. Figure 5-9
J\
shows that for burner stoichiometries of 100 percent and greater an
increase in NO is expected as the volumetric heat release rate
y\
5-34
-------
560
GO
en
80
8 12 16 20 24 28 32
Volumetric heat release rate (106 Btu/hr/ft3)
Stoichiometry to active
burners (percent)
O 120
A 100
X 80
Figure 5-9. Effect of volumetric heat release rate and burner Stoichiometry
on NOX from single wall oil-fired boilers.
-------
increases. For substoichiometric firing, the formation of NO is less
A
sensitive to changes in volumetric heat release rates. In fact, for the 80
percent burner stoichiometry curve, a slight decrease in NO formation is
3\
observed when the volumetric heat release rate is increased, though this is
expected to be only an artifact of the data. Stoichiometry to active
burners is again important in this instance, as it has been for other firing
type fuel combinations. Namely, a decrease in burner stoichiometry results
in a decrease in the oxygen concentration in the peak flame temperature
regions thus reducing NO formation.
A
The importance of the variable describing the difference between
furnace and burner stoichiometry is a little misleading. A large difference
could indicate a radical staging pattern which should greatly reduce NO
A
formation. However, since there are practical restraints limiting the
reduction of burner stoichiometry, a large difference between furnace and
burner stoichiometry would more likely indicate a higher overall excess air
level. Thus, an increase in this factor would lead to an increased level of
NO formation.
A
The heat input per active burner factor appears in several terms.
The net effect is that an increase in heat input per active burner results
in an increase in NO formation. The heat input per active burner effect
A
is tempered by a lower overall excess air level and by the number of burners
out of service or the degree of staging. Obviously, for a given load, if a
burner is taken out of service, the remaining active burners must increase
their heat input. As shown in Figure 5-10, NO formation increases as the
A
heat input per active burner increases but for most burner stoichiometries,
this dependence is weaker than the volumetric heat release rate dependence.
Flue gas recirculation did not appear as a strong contributing factor
in the NO predictions largely because of the scarcity of data on the
/\
effect of FGR as a single NO control. The effectiveness of FGR in NO
A A
control is diminished when used in conjunction with other control methods.
Stoichiometry to the active burners is probably the most significant NO
X
factor because the operator can regulate airflow to a certain extent without
affecting unit operation.
5.3.6 Horizontally Opposed Gas-Fired Boilers
The data base for the regression analysis for NO reduction on
horizontally opposed gas-fired boilers consisted of 74 tests on eight
5-36
-------
560 n
^480
CVJ
o
8
* 400
°-320
M
t/>
§
•I"
*/»
Ul
e 240
X
o
160
80
CD
ffl O
Stoichiometry to active
burners (percent)
CD 120
A 100
X 80
80
30 60 90 120 150 180
Heat input per active burner (106 Btu/hr)
210
Figure 5-10.
Effect of heat input and burner Stoichiometry on NOX
from single wall oil-fired boilers.
5-37
-------
boilers. This gave a minimum of three tests for each control and
combination of control methods considered except flue gas recirculation
alone. The tests included baseline, low excess air, off stoichiometric
combustion, load reduction, and combinations of these methods plus flue gas
recirculation.
A logarithmic equation correlated the data more effectively than
first or second order multiple regression schemes. The correlation equation
explained variations in NOV emissions to within 76 percent. The
/\
significant parameters were found to be firing rate, burner stoichiometry,
furnace stoichiometry, number of division walls, and flue gas recirculation.
The correlation equation best explaining the data, with a ""correlation
coefficient of 0.87, was:
y = 4.42
where
y = NO emissions (ppm dry at 3 percent 0?)
x, = Firing rate (percent)
x2 = Stoichiometry to active burners (percent stoichiometric air)
x., = Furnace stoichiometry (percent stoichiometric air)
x. = Number of division walls plus one
Xr = Flue gas recirculation rate plus 1 (percent)
Due to the logarithmic correlation equation, small changes in these
factors result in large changes in predicted NO emissions.
/\
The variation of NO emissions with firing rate was expected.
A
Since thermal NO formation dominates exclusively in natural gas firing,
A
any reduction in firing rate should reduce NO formation by reducing the
A
bulk furnace gas temperature. Figure 5-11 shows that both burner firing
rate and burner stoichiometry affect NO emissions significantly.
A
However, since furnace stoichiometry should not vary greatly, the key
factor will be burner stoichiometry. The effect of burner stoichiometry
is clearly shown in Figure 5-11. This is expected since staged combustion
is very effective for NO control with natural gas firing.
Flue gas recirculation entered the present correlation only weakly.
However, as with the oil firing correlations discussed above, FSR data
5-38
-------
s-s
CO
0.
Q.
1000
800
600
to
§ 400
200
Stoichiometry to active
burners (percent)
Q 120
G 100
A 30
X 60
20 30 40 50 60 70 80 90
Firing rate (percent)
100 110 120
Figure 5-11. Effect of firing rate and burner Stoichiometry on NO from horizontally
opposed gas-fired boilers.
-------
were only available in combination with other controls. Thus, less credit
was given FGR for NO reduction than if more data on FGR acting alone
A
were available. Figure 5-12 shows that gas recirculation has much less
effect on NO reduction than does burner stoichiometry. Aside from
A
using reduced firing rates, the data show that the operator can control
NO emissions most effectively by reducing burner stoichiometry.
A
5.3.7 Single Wall Gas-Fired Boilers
Forty-one tests from seven single wall gas-fired boilers were
selected for use in the multiple regression analysis. The NO control
A
techniques implemented in these tests were LEA, BOOS, FGR, load reduction,
and combinations of these methods.
For single wall gas-fired boilers, the following equation
correlates the data, with a correlation coefficient of 0.949; i.e.,
90.2 percent of the variance in the data is explained by the regression:
y = -37.2 + 1.45 x 10"5 (x1)(x2) - 1.85 x 10"4 (x1)(x3)
+ 2.09 x 101 (x3) - 6.46 x 10~3 (x2)(x4)
where
y = NO emissions (ppm dry at 3 percent 09)
?
x-j = Surface heat release rate (W/m )
x2 = Stoichiometry to active burners (percent stoichiometric air)
x., = Numbers of burners out of service
x^ = Flue gas recirculation (percent)
As was found in the other boiler/fuel classifications treated,
surface heat release rate and burner stoichiometry were the key parameters
affecting NO formation in single wall gas-fired boilers. In gas-fired
A
boilers, only thermal NO is formed and, as expected from basic combustion
principles, lowering surface heat release rate and burner stoichiometry
reduces this NO formation. This behavior is indicated in the regression
rt
equation and exhibited in Figure 5-13.
The second term of the regression equation, which has the product of
the two key parameters, surface heat release and stoichiometry, is the
dominant one. Number of burners out of service appears in the third term of
the regression, and it is seen that implementing BOOS decreases NO as
A
5-40
-------
en
i
CM
O
to
i
o.
o
«/)
I/)
1000
800 -
400
200
10
20 30 40 50
Flue gas recirculation (percent)
Stoichiometry to active
burners (percent)
0 120
A 100
X 80
Z 60
60
60
Figure 5-12. Effect of flue gas recirculation and burner Stoichiometry on NO from horizontally
opposed gas-fired boilers.
-------
tn
i
520 -i
40
10
120
Stoichiometry to active
burners (percent)
O 120
A 100
X 80
30
40
50
60
70
80
Surface heat release rate (106 Btu/hr-ft2)
90
100
Figure 5-13.
Effect of surface heat release rate and burner Stoichiometry
on NOX from single wall gas-fired boilers.
-------
expected. Surface heat release rate also appears in that third term. It
should not be interpreted as implying that an increase in heat release
decreases NO , because heat release appears in conjunction with BOOS. In
J\
other words, the regression suggests that BOOS produces a larger absolute
magnitude drop in NO for a boiler with a higher heat release rate (with
A
its expected higher baseline NO level). This points again to the dangers
A
of examining the individual terms of the regression without considering the
overall contribution of each variable to the entire correlation.
Finally, the last term indicates that flue gas recirculation does
lower NOV from single wall gas-fired boilers, as also shown in Figure 5-14.
X
It is seen that combining FGR with BOOS (lower burner stoichiometry) is an
effective NO control scheme.
A
In summary, it is seen that the two major variables affecting N0x
are surface heat release rate and burner stoichiometry. For an existing
boiler, the former can be decreased by reducing load while the latter can be
decreased by lowering excess air and implementing OSC (BOOS). This will
result in lower NO . Obviously, it would have been best to have
A
originally designed the boiler to operate with a lower surface heat release
rate and fire off stoichiometrically, for example, via wider burner spacing
and new burner design. For an existing boiler, this is not easily done, so
to reduce NO load reduction, LEA, BOOS, and FGR can be implemented.
5.4 SUMMARY
A multiple regression model was used to correlate N0x emissions
with boiler/burner design and operating variables and fuel properties. The
model explains the variation in NO on the average to within 20 percent
A
for each boiler design/fuel classification. The key variables affecting
NO formation were identified as:
/\
• Heat input per active burner
• Stoichiometry to active burners
• Firing rate
• Number of burners firing (or degree of BOOS)
• Surface heat release rate
• Furnace stoichiometry
• Percent flue gas recirculation
• Number of furnace division walls
5-43
-------
560-
480-
CVJ
O
400-
o
Stoichiometry to active
burners (percent)
O 120
A 100
X 80
10
20 30 40 50 60
Flue gas recirculation (percent)
Figure 5-14. Effect of flue gas recirculation and burner stoichiometry
on NOX from single wall gas-fired boilers.
-------
The only fuel property statistically adequate for use was the fuel type:
coal, oil, or natural gas.
Thus, the correlation model served a very useful purpose in
identifying key variables that affect NOX formation and highlighting
general trends. As an interpolative model, the correlation can be
considered good considering the high correlation coefficients achieved with
a large data base -- multiple tests on many different boilers under a
diversity of test programs and procedures.
5-45
-------
REFERENCES FOR SECTION 5
5-1. Beuters, K. A., et al., "NOX Emissions from Tangentially Fired
Utility Boilers — A Two Part Paper," presented at the 66th Annual
AIChE Meeting, Philadelphia, November 1973.
5-2. Habelt, W. W. and Selker, A. P., "Operating Procedures and Prediction
for NOX Control in Steam Power Plants," presented at the Central
States Section of the Combustion Institute, Madison, Wisconsin, March
1974.
5-3. Quan, V., et al., "Analytical Scaling of Flowfield and Nitric Oxide
in Combustors," in Proceedings: Coal Combustion Seminar.
EPA-650/2-73-021, June 1973.
5-4. Quan, V., et al., "Nitric Oxide Formation in Recirculating Flows,"
Combustion Science and Technology, Volume 7, No. 2, pp. 65 to 75,
1973.
5-5. McDonald, H., et al., "Two-Dimensional or Axially Symmetric Modeling
of Combusting Flow," in Proceedings of the Second Stationary Source
Combustion Symposium, Volume IV. EPA-600/7-77-073d, NTIS-PB 271
758/AS, July 1977.
5-6. Dykema, 0. W., "Analysis of Test Data for NOX Control in Gas and
Oil-Fired Utility Boilers," EPA-650/2-75-012, NTIS-PB 241 918/AS,
January 1975.
5-7. Dykema, 0. W. and Hall, R. E., "Analysis of Gas-, Oil-, and
Coal-Fired Utility Boiler Test Data," in Proceedings of the
Stationary Source Combustion Symposium, Volume III,
EPA-600/2-76-152C, NTIS-PB 257 146/AS, June 1976.
5-8. Bartok, W., et al., "Systematic Field Study of NOX Emission Control
Methods for Utility Boilers," Exxon Report 6RU-4GNOS-71,
NTIS-PB 210 739, EPA Contract CPA 70-90, Exxon Research and
Engineering Company, Linden, NJ, December 1971.
5-9. Crawford, A. R., et al., "Field Testing: Application of Combustion
Modifications to Control NOX Emissions from Utility Boilers,"
EPA-650/2-74-066, NTIS-PB 237 344/AS, June 1974.
5-10. Hollinden, G. H., et al., "NOX Control at TVA Coal Fired Steam
Plants," ASME Air Pollution Control Division, in Proceedings of the
Third National Symposium, April 1973.
5-11. Hollinden, G. H., et al., "Control of NOX Formation in Wall
Coal-Fired Boilers," in Proceedings of the Stationary Source
Combustion Symposium, Volume II, EPA-600/2-76-152b, NTIS-PB 256
321/AS, June 1976.
5-46
-------
5-12. Cato, G. A., et al., "Field Testing: Application of Combustion
Modifications to Control Pollution Emissions from Industrial Boilers
-- Phase 1," EPA-650/2-74-078a, NTIS-PB 238 920/AS, October 1974.
5-13. Draper, N. and Smith, H., Applied Regression Analysis, Wiley, New
York, 1966.
5-14. Sterling, T. D. and Pollack, S. V., Introduction to Statistical Data
Processing, Prentice-Hall, New Jersey, 1968.
5-15. Crawford, A. R., et al., "The Effect of Combustion Modification on
Pollutants and Equipment Performance of Power Generation Equipment,"
in Proceedings of the Stationary Source Combustion Symposium, Volume
III, EPA-600/2-76-152C, NTIS-PB 257 146/AS, June 1976.
5-16. Crawford, A. R., et al., "Field Testing: Application of Combustion
Modification to Power Generating Combustion Sources," in Proceedings
of the Second Stationary Source Combustion Symposium, Volume II,
EPA-600/7-77-073b, NTIS-PB 27l 756/AS,'July 1977.
5-17. Thompson, R. E., et al., "Effectiveness of Gas Recirculation and
Staged Combustion in Reducing NOX on a 560 MW Coal-Fired Boiler,"
EPRI Report No. FP-257, NTIS-PB 260 582, September 1976.
5-18. Blakeslee, C. E. and Selker, A. P., "Program for Reduction of NOX
from Tangential Coal-Fired Boilers," EPA-650/2-73-005, 5a and 5b,
NTIS-PB 226 547/AS, PB 245 162/AS, PB 246 889/AS, August 1973, June
1975, and August 1975.
5-19. Burrington, R. L., et al., "Overfire Air Technology for Tangentially
Fired Utility Boilers Burning Western U.S. Coal," EPA-600/7-77-117,
NTIS-PB 277 012/AS, October 1977.
5-20. Campobenedetto, E. J., "The Dual Register Pulverized Coal Burner --
Field Test Results," presented at the Engineering Foundation
Conference on Clean Combustion of Coal, Rindge, New Hampshire, July
to August 1977.
5-21. Vatsky, J., "Attaining Low NOX Emissions by Combining Low Emission
Burners and Off-Stoichiometric Firing," Paper No. 51d, 70th Annual
AIChE Meeting, New York, November 1977.
5-22. Durrant, 0. W., "Design, Operation, Control and Modeling of
Pulverized Coal-Fired Boilers," presented at the Boiler Turbines
Modeling and Control Seminar, University of New South Wales, Sydney,
Australia, February 1977.
5-23. Vatsky, J., Foster Wheeler Energy Corporation, Livingston, NJ,
Personal Communication, January 1980.
5-47
-------
SECTION 6
NO CONTROL CHARACTERIZATION: PROCESS ANALYSIS
A
To provide a meaningful evaluation of combustion modification NOX
controls, not only must their NO reduction capabilities be determined but
/\
also their impacts on boiler operation and maintenance, operating costs, and
effluent emissions other than NO . Therefore, consistent process analysis
A
procedures were developed, and applied to field test data, both published
and unreported, from full-scale applications of controls. The approach
adopted was to compare process variables that characterize the boiler system
under baseline or normal operating conditions to those under controlled or
low NO modes. Significant changes in the process variables were noted,
A
and these were highlighted as real or potential problems and concerns.
To lay the foundation for the detailed analysis of controls,
Section 6.1 summarizes the process analysis procedures and data sources
employed. Sections 6.2 through 6.13 then analyze NOX controls applied to
major boiler design/fuel classifications as available process data on
specific boiler tests permitted. A summary of the impact of NOY controls
^
on boiler operation and maintenance is then given in Section 6.14.
6.1 PROCESS ANALYSIS PROCEDURES
Process data collected during numerous utility boiler test programs
were assembled for boilers operated under baseline and low NOX
conditions. A list of process variables investigated is given in
Table 6-1. These data were then used to analyze changes in process
variables due to low NO operation and thereby estimate the potential
n
impact of such modes of firing on boiler operation and maintenance.
Potential adverse effects were identified and evaluated. The more
established combustion modification techniques for NO reduction were
A
studied extensively along with some newer and/or less common NO control
measures.
6-1
-------
TABLE 6-1. PROCESS VARIABLES INVESTIGATED
Process Variables
Process Variables
Boiler Load
Furnace Excess Air
Excess Air at Firing Zone
Percent Oxygen in Flue Gas
Percent Oxygen in Windbox
Furnace Cleanliness Condition
Percent Overfire Air
Percent Flue Gas Recirculation
Burners Out of Service
Damper Positions
Burner Tilt
Flowrates:
Superheater Steam
Reheater Steam
Superheater Attemperator Spray
Reheater Attemperator Spray
Airflow
Fuel Flow
Pressures:
Steam Drum
Superheater Steam Outlet
Reheater Steam Outlet
Furnace
Windbox
Fan Inlet
Fan Discharge
Temperatures:
Superheater Steam
Reheater Steam
Air Heater Air In/Out
Air Heater Gas In/Out
Furnace Gas Outlet
Stack Gas Inlet
Heat Absorption:
Furnace
Superheater
Reheater
Economizer
Fan Power Consumption
Gas Emissions:
NOX
SOX
Carbon Monoxide
Hydrocarbons
Polycyclic Organic Matter
Particulate Loading
Particulate Size Distribution
Ringleman Smoke Density
Carbon/Unburned Fuel Loss
Additional Factors Considered:
Corrosion Rates
Slagging and Fouling
Flame Instability
Furnace Vibration
Fan and Duct Vibrations
6-2
-------
6.1.1 Assumptions
Boilers differ widely according to type of furnace and fuels fired.
Accordingly, each boiler design/fuel classification was treated separately.
Within each classification, however, there may still be large variations in
variables which affect NOV emissions. As noted earlier, in Section 5,
/\
design variables such as furnace volumetric and surface heat release rates
can significantly affect baseline NO levels. The degree of NO control
rt «
achievable and hence the needed changes in process variables to effect these
emissions changes, therefore, may be substantially different between two
boilers of the same type and firing similar fuels. For the purpose of the
present study, however, it was assumed that these variations are small in
comparison to the variations associated with furnace and fuel types.
Moreover, it was assumed that the boilers for which data were available and
analyzed in this study are representative of that type.
Data from a few well designed tests were available in which one or
more operational variables were systematically varied to test their effect
on NO emissions. Of course for precise treatment, secondary variables
^
such as furnace conditions or fuel composition, must be maintained constant
in order to isolate the effect of the variables being studied. This is
often impossible when testing boilers under field conditions. In such
cases, it was assumed that the effect of these secondary variables on NO
A
emissions was small. Data were also available from some compliance tests.
Such tests are usually much less systematically conducted, and the
assumption that the secondary variables are maintained constant is much more
tenuous. Still some insight can be gained from analyzing these cases.
6.1.2 Procedures
Process variable data were compiled for baseline and low NO modes
of operation. The data were then analyzed and compared. Wherever possible,
comparisons of baseline and controlled operation were made on tests which
were similar in the general operating characteristics tested. Steam flow
and load conditions, overall excess air levels, furnace conditions, etc.,
were matched as closely as possible. In addition, for tangential boilers,
burner tilt and overfire air nozzle tilt were also matched for the baseline
and controlled tests selected for comparison.
In certain tests, where the process data were sufficiently detailed,
overall mass and energy balances were conducted. The mass balances were
6-3
-------
used to determine the amount of gaseous pollutants and particulate and solid
matter emitted by the boiler under baseline and low NOX conditions.
Overall energy balances were used to check boiler efficiencies. Energy
balances on individual boiler components established the distribution of
heat absorption in the boiler. Attemperator spray flowrates were checked by
heat and mass balances on superheater and reheater sections. Air and gas
volume flowrates were calculated to determine the effect of changed
operating conditions on fan draft and power requirements.
For coal-fired tests, data were collected on carbon loss in flyash,
furnace slagging, and furnace wall tube corrosion. Corrosion may be a
problem with coal-fired boilers due to the presence of sulfur and iron in
the coal. When firing under reducing conditions, increased slagging
combined with penetration of iron sulfide into the metal surfaces may
increase tube wall corrosion rates. Most of the corrosion data were from
tests conducted with corrosion coupons inserted in the furnace.
Unfortunately, although tests of this type are quite useful in determining
relative corrosion rates, they do not allow evaluating absolute wastage
rates. Data were also obtained from some tests on coal- and oil-fired
boilers on particle loading and size distribution. Some data were also
available, mainly for oil and gas fuels, on flame instability, furnace
vibrations, superheater tube temperatures and flame carryover to the
convective section. Comparison of the process data were made for baseline
and low NO modes of operation. Significant changes in the process
^
variables were noted and evaluated for their impact on emissions and boiler
operation and maintenance.
6.1.3 Data Sources
The boiler types investigated in this study were tangential,
horizonally opposed, single wall and turbo furnaces. These four types
encompass most of the fossil fuel fired utility boilers in service in the
United States. The major NO control techniques analyzed in detail were
/\
off stoichiometric combustion (OSC) and flue gas recirculation (F6R). Off
stoichiometric combustion includes firing with burners out of service
(BOOS), biased burner firing (BBF) and overfire air (OFA) injection above
the burner array. Off stoichiometric combustion was studied as applied to
coal-, oil- and gas-fired boilers, whereas FGR to the windbox was treated in
detail only in oil and gas fuels applications. In addition to these
6-4
-------
techniques various other methods on which sufficient process data were
available were included in the study. Low excess air (LEA) firing was
treated both as a NO control in this study and as a standard operating
/\
procedure. Low NO burners have been tested and are being installed in
some boilers. Some data are also available on water injection (WI) and
reduced air preheat (RAP), although they are not widely used as low NOX
techniques due to associated losses in boiler efficiency.
Table 6-2 gives a list of the boilers for which process data were
available under low NO operation and which were used in this study. The
^
sources of data are also listed in the table under the column marked
References. All available published NO control test reports were
A
reviewed for process data of sufficient detail for this investigation. In
addition, several major boiler manufacturers and utility companies
graciously supplied new or previously unpublished process data from their
own test programs.
It should be noted that the omission of a NO control technique in
rt
Table 6-2 for a given boiler/fuel classification does not necessarily
signify that that technique is not effective in controlling NO
emissions. Some NO control measures had to be left out due to lack of
^
adequate process data on those techniques for certain boiler/fuel
classifications. The summary given in Section 6-14 attempts to fill in
these gaps by giving a general survey of expected operational and
maintenance impacts for all important NO control measures.
^
In the following sections, the major boiler/fuel classifications and
applied controls, as discussed above, are analyzed as available data
permitted.
6.2 TANGENTIAL COAL-FIRED BOILERS
Tangential coal-fired boilers have been perhaps the most studied
boiler/fuel classification for potential NO control. Consequently, a
A
substantial quantity of process data have been collected on these units
operated under baseline and low NOY conditions. The major low NO
^ X
techniques tested have been LEA and OSC. Under OSC, both BOOS and OFA
firing have been investigated. Very few adverse effects attributable to low
NOX operation have been reported in the numerous tests conducted. The
major problem encountered was that of boiler derating associated with BOOS
operation. Particulate loading also seemed to increase substantially in
6-5
-------
TABLE 6-2. SUMMARY OF PROCESS DATA SOURCES
Furnace
Type
Tangential
Opposed Hall
Single Wall
Tangential
Opposed Wall
Fuel
Coal
Coal
•
Coal
011
Oil
Boiler
Barry No. 2
Barry No. 4
Hunting ton Canyon No. 2
Columbia No. 1
Navajo No. 2
Cooianche No. 1
Kingston No. 6
Harllee Branch No. 3
Four Corners No. 4
Hatfleld No. 3
E.C. Gaston No. 1
"BtW Units Nos. 1 & 2"»
•FH Unit No. A"
Widows Creek No. 5
Widows Creek No. 6
Crist Station No. 6
Mercer No. 1
"FH Unit No. B"»
•FH Unit No. C"»
South Bay No. 4«
Plttsburg No. 7
—
Moss Landing Nos. 6 & 7*
Oraond Beach Nos. 1(2
__
Sewaren Station No. 5
•FH Unit No. 0"
Manufacturer
CE
CE
CE
CE
CE
CE
CE
BM
BM
B&U
B&U
B&U
FU
BM
BM
FU
FU
FU
FH
CE
CE
CE
BM
FU
B&U
BM
FU
Utility Ccnpany
Alabama Power
Alabama Power
Utah Power and Light
Wisconsin Power & Light
Salt River Project
Public Service of Colorado
Tennessee Valley Authority
Georgia Power
Arizona Public Service
Allegheny Power Service
Southern Electric Generating
—
--
Tennessee Valley Authority
Tennessee Valley Authority
Gulf Power
Public Service Electric t Gas
—
—
San Diego Gas & Electric
Pacific Gas t Electric
Southern California Edison
Pacific Gas t Electric
Southern California Edison
Southern California Edison
Public Service Electric ( GAs
—
NOX Control
Technique
BOOS. OFA
LEA. BOOS
OFA
OFA
LEA. BOOS. OFA
OFA
LEA. BBF. BOOS
LEA. BOOS
BOOS. HI
BOOS. FOR
LNB, LEA. BOOS
LNB
LEA. BOOS. LR
LEA, BOOS
LEA. BOOS
LEA. BOOS
LEA. BBF
LEA. BOOS, LR
LEA. OFA, LR
LEA. BOOS. RAP
OFA. FGR
FGR, BOOS
OFA. FGR
FGR. OFA. BOOS. HI
FGR. OFA. BOOS
LEA. BOOS
LEA. OFA. BOOS. FGR
New or
Retrofit
Retrofit
Retrofit
New. NSPS
New, NSPS
New. NSPS
New
Retrofit
Retrofit
Retrofit
New
Retrofit
New, NSPS
Retrofit
Retrofit
Retrofit
Retrofit
Retrofit
New. NSPS
Retrofit
Retrofit
UAW
new
Retrofit
OFA New
FGR Retrofit
OFA New
FGR Retrofit
Retrofit
OFA New
BOOS. FGR
Retrofit
Reference
6-1
6-2
6-3
6-3
6-4
6-4
This report. Sec. 8.1
6-2
6-2
6-5
6-4
This report, App. A
6-6. 6-7
6-2. 6-6
6-2
6-7
This report. App. B
This report, App. B
6-8
6-9
6-10. 6-12
6-9. 6-11
6-10, 6-12
6-10
6-7
This report, App. B
'Denotes new results or previously unreported data.
-------
TABLE 6-2. Concluded
CTI
I
Furnace
Type
Single Hall
Turbo Furnace
Tangentl al
Opposed Hall
Single Hall
Turbo Furnace
Fuel
Oil
011
Gas
Gas
Gas
6as
Boiler
Enclna Nos. 1, 2 t 3«
South Bay No. 3*
Potrero No. 3-1
South Bay No. '4*
Plttsburg No. 7
Moss Landing Nos. 617*
Plttsburg Nos. 5(6
Contra Costa Nos. 9 I 10
Enclna Nos. 1, 2 t 3»
South Bay No. 3*
Potrero No. 3-1
Manufacturer
B&W
RS
RS
CE
CE
UH
BM
BM
BM
RS
RS
Utility Company
San Diego Gas t Electric
San Diego Gas t Electric
Pacific Gas I Electric
San Diego Gas t Electric
Pacific Gas t Electric
Pacific Gas t Electric
Pacific Gas t Electric
Pacific Gas t Electric
San Diego Gas I Electric
San Diego Gas t Electric
Pacific Gas t Electric
NOX Control
Technique
LEA, BOOS
Air adjustment
HI, RAP
OFA. FOR
LEA, BOOS
OFA. FOR
OFA, FGR
OFA. FGfl
OFA, FGR
BOOS
Air adjustment
HI, RAP
OFA. FGR
New or
Retrofit
Retrofit
Retrofit
Retrofit
Retrofit
Retrofit
Retrofit
Retrofit
Retrofit
Retrofit
Retrofit
Retrofit
Reference
6-13
6-8
6-9
6-8
6-9
6-9. 6-11. 6-14
6-9
6-9
6-13
6-8
6-9
•Denotes new results or previously unreported data.
-------
certain units under OSC. However, other important process variables such as
efficiency, corrosion, carbon losses, particulate size distribution, and
heat absorption profiles remained either unaffected or changed only by small
amounts under low NO operation. A detailed discussion on the results of
A
various tests is given below. The results are summarized at the end of the
subsection.
Details of extensive tests carried out on three tangential coal-fired
units by Combustion Engineering, Inc., are given in References 6-1 and 6-3.
The three boilers tested were: Barry No. 2, a 125 MW unit operated by
Alabama Power Company; Columbia No. 1, a 525 MW unit operated by Wisconsin
Power and Light Company; and Huntington Canyon No. 2, a 430 MW unit operated
by Utah Power and Light Company. The Barry boiler is an older unit which
was retrofitted with OFA ports during the course of testing. The other two
are new NSPS units with factory-equipped OFA ports. The types of coal fired
were Eastern bituminous, Western sub-bituminous and Western bituminous for
Barry, Columbia and Huntington, respectively. A comparison of process
variables under baseline is shown in Tables 6-3, 6-4, and 6-5 for the three
boilers.
From the tables, it is seen that OSC was quite effective in
controlling NO emissions from all three boilers. Of course, the range in
A
NO reduction varied from boiler-to-boiler and from test-to-test due to
A
variations in baseline excess air levels, amount of reduction in burner
stoichiometry, and differences in boilers and fuels fired. However, on the
average, for all three boilers, it was found that NO levels decreased by
fi
40 to 55 ng/J (0.09 to 0.13 lb/10 Btu) for a 10 percent decrease in air
to the burners.
The major impacts of NOV controls occurred with BOOS firing on the
/\
Barry boiler where the unit was derated by approximately 20 percent as is
shown in Table 6-3. In this boiler, derating occurred due to lack of spare
coal pulverizer capacity. In general, boiler derating will occur in all
coal-fired boilers without extra pulverizer capacity when operated under
BOOS. From Table 6-3 it is also seen that in the Barry unit, gas
temperature at the furnace outlet increased with OFA firing, as measured by
special thermocouples installed to make these measurements. This was
expected though, since OSC operation tends to lengthen the combustion zone
so that, for the same load and burner tilt, completion of combustion occurs
6-8
-------
TABLE 6-3. COMPARISON OF FLOW VARIABLES FOR A 125 MW TANGENTIAL EASTERN BITUMINOUS COAL-FIRED BOILER
OPERATED UNDER SIMILAR CONDITIONS AT BASELINE AND LOW N0¥ CONDITIONS (Reference 6-1)
I
UD
Process Variables
Test Condition
Furnace Condition
Load
Main Steam Flow
Furnace Excess Air
Th. Air to Fuel Fir. Zone
Burner Tilt
OFA Tilt
Boiler Efficiency
NO'
COS
C loss in Flyash*
Dust Loading'
SH Temp
RH Temp
Steam Pressure
SH Attemp. Spray Flow
RH Attemp. Spray Flow
Furnace Outlet Temp
Heat Absorption Profile
Economizer
Furnace
Primary Superheater
Secondary Superheater
Reheater
Total Heat Absorbed
Losses
kg/s (103 Ib/h)
Percent
Percent
Degrees
Degrees
Percent
PP« (OX 02)
ppm (OX 02)
Percent
g/m3 (10"3 Ib/scf)
;g;
HPa (psi)
kg/s (103 Ib/h)
kg/s (103 Ib/h)
K ( F )
Percent of total
Heat release
Heat release
Heat release
Heat release
Heat release
Heat release
Base 1 i ne
Full load
Clean
124
112.2 (890.7)
22.7
117.9
+3
-
89.0
494
31.2
0.48
4.2 (0.262)
812 (1002)
787 (957)
12.8 (1859)
1.1 (9.10)
0.25 (2.0)
1499 (2239)
4.0
47.9
17.8
8.1
11.1
89.0
11.0
BOOS
Operation
Maximum
possible load
Clean
102
87.2 (692)
24.2
94.7
-5
-
88.8
285
26.6
0.25
8.6 (0.540)
811 (0.011)
788 (959)
12.7 (1845)
2.0 (15.6)
0.11 (0.9)
1468 (2183)
4.0
47.5
16.0
9.8
11.4
88.8
11.2
Significant
Difference
from Baseline
for BOOS firing
-18 X
-22 X
-23.2
-42X
-48X
+1061
»71X
-55X
-31K (-56°F)
-10*
»21t
OFA
Operation
Full load
Clean
125
115.6 (917)
21.6
90.7
-4
0
89.0
339
26.1
0.61
8.58 (0.539)
811 (1000)
809 (997)
12.9 (1873)
4.94 (39.2)
0
1560 (2350)
2.6
46.0
17.7
9.8
12.8
89.0
11.0
Significant
Difference
from Baseline
for OFA firing
-27.2
-31X
+27X
+1061
22K (40°F)
+33X
-100X
61K (+111°F)
-35X
*?1*
+ 15X
aAt economizer outlet
-------
TABLE 6-4. COMPARISONS OF PROCESS VARIABLES FOR A 525 MW TANGENTIAL WESTERN
SUB-BITUMINOUS COAL-FIRED BOILER OPERATED UNDER SIMILAR
CONDITIONS AT BASELINE AND LOW NOX MODES (Reference 6-3)
Process Variables
Test Conditions
Furnace Conditions
Load
Main Steam Flow
Furnace Excess Air
Th. Air to Fuel Fir. Zone
Burner Tilt
OFA Tilt
Boiler Efficiency
N0xa
C0a
C loss in Flyash8
SH Temp
RH Temp
SH Attemp. Spray Flow
RH Attemp. Spray Flow
Steam Pressure
FD Fan
ID Fan
Heat Absorption
Economizer
Furnace
Primary Superheater
Secondary Superheater
Reheater
Total Heat Absorbed
Losses
MW
kg/s (106 Ib/hr)
Percent
Percent
Degrees
Degrees
Percent
ppm (OX 02)
ppm (OX 02)
Percent
K (°F)
K ( F)
kg/s (103 Ib/hr)
kg/s (103 Ib/hr)
MPa (psi)
Amps
Amps
Percent of Total
Heat Release
Heat Release
Heat Release
Heat Release
Heat Release
Heat Release
Heat Release
Baseline
Full Load
Clean
524
442 (3.51)
21.8
118.9
+1
0
87.5
520
16
0.03
813 (1004)
815 (1008)
11.0 (87.3)
12.0 (95.2)
16.88 (2448)
401
920
14.4
27.3
17.2
11.2
17.2
87.5
12.5
OFA
Operation
Full Load
Clean
523
444 (3.52)
26.9
106.0
-5
0
87.3
389
10
0.02
817 (1011)
819 (1015)
19.0 (150.8)
8.0 (63.5)
16.95 (2458)
434
1000
15.7
25.2
17.0
13.2
16.3
87.3
12.7
Significant
Difference
Significant
-12.9
-25X
+73X
-33X
+8X
+9X
aAt economizer outlet
6-10
-------
TABLE 6-5. COMPARISON OF PROCESS VARIABLES FOR A 430 MW TANGENTIAL WESTERN
BITUMINOUS COAL-FIRED BOILER OPERATED UNDER SIMILAR CONDITIONS
AND LOW NOX CONDITIONS (Reference 6-3)
Process Variables
Test Conditions
Furnace Condition
Load
Main Steam Flow
Furnace Excess Air
Th. Air to Fuel F1r. Zone
Burner T1U
OFA T1lt
Boiler Efficiency
«;
CO*
C loss In Flyash*
SH Teap
RH Tenp
Stean Pressure
SH Attewp. Spray Flow
RH Attest. Spray Flow
FD Fan
10 Fan
Heat Absorption Profile
Economizer
Furnace
Primary Superheater
Secondary Superheater
R Chester
Total Heat Absorbed
Losses
MW
kg/s (106 Ib/hr)
Percent
Percent
Degrees
Degrees
Percent
PPM (OS 02)
PPM (OX 02)
Percent
K (°F)
K (°F)
KPa (psl)
kg/s (103 Ib/hr)
kg/s (103 Ib/hr)
Amps
AMPS
Percent of Total
Heat Release
Heat Release
Heat Release
Heat Release
Heat Release
Heat Release
Heat Release
Base 1 1 ne
Max Load
Nod Dirty Furnace
433
375 (2.98)
20.2
118.1
+8
0
90.34
514
20
0.50
809 (997)
811 (1000)
17.24 (2500)
0
6.0 (47.6)
434
772
16.6
28.5
12.9
16.4
15.8
90.3
9.7
OFA
Operation
Hax Load
Mod Dirty Furnace
426
370 (2.94)
19.2
96.6
+10
0
90.46
446
162
0.24
804 (998)
816 (1009)
16.99 (2464)
0
1.0 (7.9)
433
774
14.8
31.2
13.8
15.4
15.4
90.5
9.5
Significant
Difference
-12.5
-13X
+138 ppM
-52X
-83!
-Ill
+9»
*At econo»1zer outlet
6-11
-------
higher in the furnace. The furnace outlet gas temperature, therefore, rises
and the heat transfer to the convective section is correspondingly
affected. In Table 6-3, it is seen that the superheater attemperator spray
flowrate is approximately quadrupled on OFA firing at Barry. This is,
however, atypical. Table 6-4 shows that at Columbia the increase in spray
flow was 73 percent under OFA, and Table 6-5 indicates no attemperation was
necessary under OFA at Huntington. Even at Barry and Columbia, where
increased attemperation was required, the spray flow never exceeded
5 percent of the main steam flow, which is well within acceptable design
limits. There was, therefore, no danger of attemperator capacities being
exceeded which would have caused serious problems and resulted in boiler
derating. Moreover, in all three boilers, the reheater spray attemperator
flowrates did not increase with OSC operation; thus, there was no adverse
effect on cycle efficiencies.
The changed gas temperature profile due to OSC may be expected to
change the heat absorption profile in the boiler. In extreme cases, this
could necessitate hardware changes such as removal of superheater and
reheater surface and perhaps addition of economizer surface to make up the
difference. The heat absorbed in the various components of the boiler is
given in Tables 6-3, 6-4, and 6-5 and is depicted graphically in
Figures 6-1, 6-2, and 6-3 for the Barry, Columbia and Huntington units,
respectively. It is seen that changes in the heat absorption profile were
only minor.
To show the effect of OSC operation on other emissions, an overall
mass balance is given in Figure 6-4 for the Barry No. 2 unit. The sulfur
dioxide emissions varied only slightly, mostly due to the variation in
sulfur content of the coals fired. Carbon monoxide emissions were also
mostly unaffected. Carbon monoxide generation usually increased sharply
once the burner stoichiometry or overall excess air level dropped below a
certain limit. Boiler operating conditions under OSC should therefore be
set so as to always operate above this limit. The particulate carryover at
the economizer outlet also increased substantially both under BOOS and OFA
operation. Results of tests on other boilers (as discussed in subsequent
sections) show mixed results of the effects of low NO operation on
}\
particulate emissions. Nevertheless, the possibility of increased
particulate emissions remains a source of concern.
6-12
-------
50
CTl
i—•
OJ
OJ
I/)
u 40
s-
4-J
OJ
O)
CL
Q.
O
I 10 _
-------
30 ^
O)
1/1
T3
O)
OJ
2Q -
O
01
u
Q
a.
15 -
- 10 _
a.
o
00
-Q
QJ
Baseline
OFA
Operation
Economizer Furnace Superheater Reheater
Losses
Figure 6-2. Heat absorption profile for Columbia Unit No. 1
(Reference 6-3).
6-14
-------
35 -n
30 _
QJ
HJ
«! 25
QJ
O)
ro
4->
O
QJ
O
QJ
Q.
Q.
S-
O
l/l
03
QJ
20 -
15 -
10 -
5 -
Baseline
OFA
Operation
Economizer Furnace Superheater Reheater Losses
Figure 6-3. Heat absorption profile for Huntington Canyon Unit
No. 2 (data from Reference 6-3).
6-15
-------
Coal
Pulverizers
To Stack
Stream
Location
Material
Baseline
BOOS
OFA
Input kg/s (Ib/hr)
Pulverizer
Coal
13.0
(103 x 10J)
10.6 ,
(84.1 x 103)
13.6 ,
(108 x 103)
FD Fan
Air
142 ,
(1.13 x 106)
116 fi
(0.922 x 10°)
147 fi
(1.17 x 10°)
Furnace Bottom
Ash
0.835
(6.63 x 10J)
0.393 ,
(3.12 x 10J)
0.741
(5.88 x 10J)
Output kg/s (Ib/hr)
Economizer Outlet
Total Gas
154 A
(1.22 x 10b)
(0.994 x 106)
159 ,
(1.26 x 106)
"°x
87.6 x 10"3
(695)
40.8 x 10'3
(324)
64.4 x 10'3
(511)
so2
0.404 ,
(3.21 x KT)
0.454 ,
(3.60 x 103)
0.444 ,
(3.52 x 103)
CO
3.36 x NT3
(267)
2.33 x 10'3
(18.5)
3.02 x 10'3
(24.0)
Particulates
0.528 ..
(4.19 x 103)
?78at x 103)
1.12
(8.92 x 103)
Figure 6-4. Overall mass balance for Barry Unit No. 2 boiler (Reference 6-1).
-------
The impacts of NOX controls on other indicators of boiler
operations, such as carbon loss and boiler efficiency, were also
investigated in tests on these three boilers. The carbon loss in flyash,
averaged over all the tests, increased by about 0.25 percent for every
10 percent decrease in burner air. There was, however, a large scatter in
the results as evidenced by the entries in the tables. The tables also show
that boiler efficiencies were largely unaffected by OSC operation.
Corrosion rates were also measured on the tests by the use of
corrosion coupons inserted in the furnace for 30-day periods. It was found
that the average weight loss per unit area of the coupons increased by about
75 percent under OSC operation for Barry, while it remained essentially
unchanged for Columbia and actually decreased by 25 percent for Huntington.
However, the weight losses for Barry and Huntington were within the range of
losses that would be expected for the oxidation of carbon steel for a 30-day
period. So, the result of the corrosion coupon tests must be regarded as
inconclusive. Still, furnace wall corrosion does not appear to be a major
problem with OSC operation on tangentially fired boilers.
Emissions and other data for the three boilers were also taken under
various conditions of furnace slagging for baseline as well as BOOS and OFA
operation. It was expected that NO emissions would increase with
rt
increased slagging due to lower heat absorption in the furnace and resulting
higher temperatures. For example, at Barry where the furnace outlet
temperatures were measured, the temperatures under baseline full-load
conditions rose by an average of 52K (125°F) when the furnace was
extremely dirty as compared to when it was clean. Surprisingly, however,
furnace conditions had a wide but inconsistent effect on N0¥ emissions in
A
all three boilers. Carbon monoxide emissions and boiler efficiencies also
did not show any trends nor did heat absorption profiles change
significantly. However, carbon loss decreased slightly with increasing
furnace water wall deposits, presumably because the higher furnace
temperatures promoted complete burnout.
Additional data on tangential coal-fired boilers are available from
two other studies. A 350 MW boiler, Alabama Power Company's Barry No. 4 was
tested under low excess air and BOOS firing (Reference 6-2). Particulate
emissions increased from an average of 1.54 ug/J (3.57 lb/106 Btu) under
baseline conditions to 2.38 yg/J (5.53 lb/106 Btu) under low NO
^
6-17
-------
operation. Carbon loss in flyash actually decreased from an average value
of about 25 percent baseline to 17 percent for low NO . No discernible
A
differences were detected in corrosion rates or boiler efficiency. The
boiler operates with five levels of burners. Boiler derate of up to
20 percent occurred with one tier on air only and close to 50 percent with
two tiers on air only.
In two other test programs, the 800 MW, Salt River Project, Navajo
No. 2 boiler was tested under low excess air, BOOS, and OFA firing; the
350 MW, Public Service Company of Colorado, Comanche No. 1 boiler was tested
under OFA firing (Reference 6-4). The Navajo No. 2 boiler did not require
derating with the top tier of burners on air only out of a total of seven
burner levels. But the particulate loading increased from an average of
1.58 yg/J (3.68 lb/106 Btu) to 2.29 yg/J (5.33 lb/106 Btu) when going
from baseline to low NOV operation. Carbon loss did not vary much, nor
A
did corrosion rates or boiler efficiency. Particulate size distribution
also did not change very much on a percentage basis with low NO firing.
A
For the Comanche No. 1 boiler, the particulate loading actually decreased
from an average of 1.35 yg/J (3.15 lb/106 Btu) at baseline to 1.07 yg/J
(2.49 lb/10 Btu) at low NO operation. Carbon loss in flyash also
A
decreased from 0.60 percent at baseline to 0.43 percent at low NO
A
corrosion, efficiency and particulate size distribution data were not
obtained for this boiler.
In summary, OSC has been shown to be an effective NO control
A
technique for tangential coal-fired boilers. Of the two common methods for
implementing OSC, namely OFA and BOOS, the former is to be preferred in
cases where a lack of spare pulverizer capacity would result in derating
with BOOS. OFA ports are included in all new post-NSPS tangential utility
boiler designs. Older boilers can be retrofitted with OFA ports, if
necessary, to prevent boiler derating with BOOS, though retrofit OFA is
generally less effective in reducing NO than BOOS. OSC does not result
in any other major adverse effect, except for potential increases in dust
loading, as observed in some boilers. This may in some cases necessitate
installation of larger or more efficient dust collection devices. However,
no change in particle size distribution was reported. No significant
changes in heat absorption profiles were noted. Superheater spray
attemperation increased substantially in some cases but were still well
6-18
-------
within normal design limits. Reheater attemperation did not increase with
OSC. The efficiency of the boilers remained, by-and- large, unaffected.
Finally, corrosion rates did not increase significantly with OSC operation,
based on tests with corrosion probes.
6.3 HORIZONTALLY OPPOSED COAL-FIRED BOILERS
A number of studies have been conducted to evaluate the effects of
NO control techniques on horizontally opposed coal-fired boilers (e.g.,
n
References 6-2, 6-4, 6-5, 6-7, and 6-15). Some data have been reported on
potential adverse affects resulting from NOX control measures such as
excessive slagging and corrosion, loss in efficiency, boiler derating,
increased dust loading, etc. Test results are also available on
horizontally opposed coal-firing units equipped with low NO burners. In
general, it has been found that low NO operation of horizontally opposed
A
boilers does not result in serious side effects with the exception of boiler
derating associated with burner out of service (BOOS) firing. Also,
although short-term tests with corrosion coupons do not indicate increased
furnace wall corrosion rates with low NO operation, long-term tests are
underway to resolve several uncertainties associated with the short-term
tests.
In an Exxon study (Reference 6-2), the 480 MW, B&W, Georgia Power,
Harllee Branch No. 3 boiler, and the 800 MW, B&W, Arizona Public Service,
Four Corners No. 4 boiler were tested for particulate emissions, corrosion,
efficiency, and carbon loss under several NO control modes of operation.
^
The Harllee Branch boiler had a baseline NO emission of 711 ppm. Low
excess air reduced this by 10 percent. Staging with four to six burners of
the top burner row on air only reduced NO emissions by one third without
^
any reduction in load. With all 10 burners on the top row on air only, the
NOX emissions decreased by half, but also resulted in a load reduction of
17 percent from 480 to 400 MW. Reducing load alone by 17 percent without
OSC or LEA decreased N0y by only about 20 percent. Particulate emissions
from this boiler did not increase significantly with low N0¥ operation; an
6
average of 1.44 yg/0 (3.36 lb/10 Btu) was measured at baseline compared
to an average of 1.60 yg/J (3.72 lb/106 Btu) at low NO conditions.
A
Carbon loss in flyash also increased from 3.8 percent on average at baseline
to 9.0 percent on average at low NOX conditions. Changes in boiler
6-19
-------
efficiency were negligible, and corrosion rates as measured on corrosion
coupons indicated wide scatter in the results and no evidence of higher
rates associated with low NOY firing.
/\
The Four Corners boiler had a baseline NOV emission of 935 ppm.
/\
BOOS firing, by having from 8 to 12 burners on air only, reduced NO
/\
emissions by approximately 50 percent without a reduction in load. Some
firing patterns, however, did result in a boiler load reduction from 800 MW
down to 600 MW. During the course of some BOOS tests, about 0.2 pound of
water per pound of coal fired was injected into the furnace by the operator
to help improve precipitator efficiency. This resulted in up to 80 ppm
additional reduction in NO emissions. The particulate emissions and
A
carbon losses actually decreased with low operation. An average of
3.56 pg/J (8.28 lb/106 Btu) of particulates and 0.61 percent of carbon in
flyash under baseline firing reduced to 3.00 yg/J (6.99 lb/10 Btu) and
0.32 percent, respectively, under low NO firing. Corrosion and
/\
efficiency measurements exhibited no significant changes.
Due to the uncertainties involved in extrapolating data from
corrosion coupons to furnace wall wastage rates, long-term data on corrosion
of actual furnace tubes are needed. Thus, Exxon has installed furnace tube
panel test specimens on the 500 MW, Foster Wheeler, Gulf Power Company,
Crist Station No. 7, horizontally opposed coal-fired boiler to evaluate the
long-term effects of low NO operation on corrosion (Reference 6-7). The
A
boiler was operated under low NO conditions, including low excess air and
A
staging, for a period of about 1 year. Testing should be complete at
present, and results should be available in the near future.
In another study, the 560 MW, B&W, West Penn Power, Hatfield Unit
No. 3 was tested by KVB, Inc. (Reference 6-5). This horizontally opposed
coal-fired boiler was tested for NO emissions and possible adverse
A
effects under BOOS firing and operation with flue gas recirculation (FGR).
Baseline NO emissions of about 900 ppm were reduced by 35 percent by
/\
putting 10 out of 40 burners on BOOS, and reductions of up to 17 percent
were achieved with 15 percent FGR. Combination of BOOS and FGR resulted in
about a 10 percent further reduction in NO from levels achieved using
A
BOOS alone. BOOS operation resulted in approximately 50 MW derate of the
boiler, and a decrease in efficiency of up to 0.3 percent. Operation with
FGR and BOOS resulted in decreases in efficiency up to 1 percent. No
6-20
-------
corrosion or erosion tests were performed so the effects of F6R and/or BOOS
on corrosion and erosion are not known for this particular boiler. No other
operational difficulties or adverse effects were encountered. Stable flames
and uniform combustion were obtained throughout the test program.
Particulate loading, flyash resistivity and carbon carryover were
essentially unchanged during low NO firing. No significant slagging or
/\
fouling of the tube surfaces was observed. Average tube metal temperatures
remained essentially unchanged, and steam temperatures were maintained near
normal levels with automatic control. High gas recirculation rates did not
require use of additional reheat attemperation.
The Hatfield unit was also tested for polycyclic organic matter (POM)
emissions by KVB, Inc. under baseline and low NO conditions. The low
A
NO conditions tested were: BOOS (8 burners out of service on the rear
A
wall, out of a total of 40 burners), 15 percent F6R, and 8 BOOS +15 percent
FGR. The baseline and low NO tests were all carried out under similar
A
load and excess air conditions. The load was maintained between 445 to
455 MW during the tests and the excess oxygen varied from 4.9 to 5.5 percent.
The POM emissions from these tests as measured upstream of the precipitator
are summarized in Table 6-6. Total POM emissions increased by about
30 percent due to BOOS operation, decreased slightly with FGR operation and
increased by about 40 percent when BOOS and FGR operations were carried out
simultaneously. The individual constituents of the total POM emission
showed varied and somewhat inconsistent trends with low NO operation.
For example, the anthracene/phenanthrene levels, which constitute about half
of the total POM emissions, did not change significantly from baseline with
BOOS firing, decreased by 18 percent from baseline with FGR operation, but
increased by 29 percent from baseline with combined BOOS + FGR operation.
As sampling and laboratory analysis methods for POMs are changing rapidly,
these results should be treated with due caution. At present the only
conclusion that can be drawn is that POMs are likely to increase slightly
with OSC operation.
Exxon has tested a horizontally opposed coal-fired boiler retrofitted
with the B&W dual register low NOX burners. The 270 MW, B&W, Southern
Electric Generating Company, E.G. Gaston Boiler No. 1 was tested and
compared with a sister unit, Gaston Boiler No. 2, not equipped with the dual
register burners (Reference 6-4). The boiler with regular burners had a
6-21
-------
TABLE 6-6. SUMMARY OF POM EMISSIONS FROM HATFIELD UNIT NO. 3 MEASURED UPSTREAM OF ESP (Reference 6-5)
Substance
Athracene/Phenanthrene
Methyl Anthracenes
Fluoranthene
Pyrene
Chrysene/Benz( a)Anthracene
Total POM
Baseline
yg/MJ
54.3
16.3
15.6
4.55
0.09
90.9
BOOS Operation
yg/MJ
54.6
14.8
33.6
15.8
—
18.8
Percent
Difference
from
Baseline
+0.5
-9.3
+114.5
+247.9
—
+30.7
FGR Operation
yg/MJ
44.3
27.2
7.49
7.11
—
86.1
Percent
Difference
from
Baseline
-18.5
+66.9
-52.1
+56.3
—
-5.3
BOOS + FGR Operation
yg/MJ
70.1
30.0
13.3
14.3
—
127.7
Percent
Difference
from
Baseline
+29.1
+83.7
-15.2
+214.6
—
+40.5
i
ro
ro
-------
baseline NOX emission of 595 ppm compared with a baseline of 387 ppm on
the boiler with the new burners. Gaston No. 1 was also tested under LEA and
BOOS firing. LEA reduced NOX further by 29 percent. BOOS with one top
row of burners on one wall on air only reduced NO to 240 ppm accompanied
by a reduction in load to 250 MW. With the top rows of burners on both
walls on air only the NOX levels could be reduced to as low as 182 ppm at
190 MW.
No significant differences were observed in boiler efficiency and
corrosion rates between the two units. The carbon loss in flyash for the
Gaston No. 2 boiler under baseline conditions averaged 1.87 percent and the
particulate loading averaged 2.31 yg/J (5.34 lb/105 Btu). The Gaston No.
1 boiler, with the retrofitted low NO burners, when operated under
baseline conditions averaged 4.37 percent on carbon loss in flyash and
2.67yg/J (6.21 lb/106 Btu) on particulate loading. The particle size
distribution seemed to shift towards smaller particle sizes with LNB. For
Gaston No. 2 over 90 percent by weight of particles were above 2.5ym and
about 2 percent less than 0.5ym. For Gaston No. 1, with LNB, about
60 percent by weight of particles were larger than 2.5ym and 10 percent
smaller than 0.5ym. The particle distribution in Gaston No. 1 did not
change significantly when the boiler was operated under BOOS conditions. It
should be mentioned that comparisons between two different boilers, even if
similar in design, is subject to uncertainties as slight differences in
operating conditions in the boilers can lead to significant differences in
results. B&W claims that its new burners when operating under normal
conditions do not result in adverse effects such as increased carbon loss or
particle loading (Reference 6-15).
Predicted performance specifications on two similar B&W boilers, one
with the standard cell burners and the other factory equipped with the new
low NOX burners, are summarized in Table 6-7 (Reference 6-16). The two
units are identical except for burner design. Table 6-7 shows that only
minor changes are expected in the process variables due to installation of
the new burners. Note that Unit 2 with low NOX burners operates with a
slightly lower unit efficiency, and this is due to the higher excess air
level employed with Unit 2. However, the incremental excess air is used to
cool the installed OFA ports (not in use) that came with Unit 2 and is not a
requirement of the low NOV burners.
6-23
-------
TABLE 6-7. COMPARISON OF PERFORMANCE SPECIFICATIONS ON TWO SIMILAR
HORIZONTALLY OPPOSED COAL-FIRED BOILERS (Reference 6-16)
Process Variables
Load Condition, MU
Number of Burners
Furnace Volume m3 (ft3)
Furnace surface a£ (ft^)
Quantity kg/s (103 Ib/hr)
Steam
Fuel
A1r
Temperature K (°F)
Steam at SH outlet
Steam at RH outlet
Flue gas at economizer outlet
Pressure MPa (pslg)
. Steam at SH outlet
Steam at RH Inlet
Excess air at economizer outlet, %
Heat loss due to unburned combustion, X
Unit Efficiency
Unit 2
(Low NOX Burners)
550
40
12,000 (425,000)
6,595.5 (70,993)
478.8 (3,800)
57.6 (457)
589.0 (4,675)
813.7 (1,005)
813.7 (1,005)
64.3 (698)
18.17 (2,620)
4.15 (587)
22
0.3
88.24
Unit 1
(Cell Burners)
550
40a
12,000 (425,000)
6,595.5 (70,993)
478.8 (3,800)
57.6 (457)
589.0 (4,526)
813.7 (1,005)
813.7 (1.005)
643.2 (698)
18.17 (2,620)
4.15 (587)
20
0.3
88.35
*20 cell burners, 2 burners each
6-24
-------
Process data on a pre-NSPS Foster Wheeler unit have recently been
released (Appendix B). The boiler, designated Unit A, has a capacity of
456 kg/s (3.62 x 105 Ib/hr) of superheated steam and 403 kg/s (3.2 x 10
Ib/hr) of reheated steam. It has 24 high turbulence burners arranged in a 4
wide by 3 high array on two opposite walls. Tests were performed on the
boiler, and the effect of excess air, load and staging on the boiler process
variables were determined. The process data are shown in Table 6-8. The
first column gives the baseline case at full load, 20 percent excess air and
no burners out of service. No low excess air test data are available, but
the effect of increasing the excess air level to 35 percent is shown in the
second column. The NO emissions are seen to increase by about 5 percent
^
over the baseline case.
The last two columns give the process data for the boiler operated at
75 percent MCR. The effect of a reduction in load alone is shown in column
3, where the NO emissions are 17 percent below the baseline level. In
rt
column 4, the combined effect of a load reduction and staging is shown.
Since this unit is not equipped with OFA ports, staging was accomplished by
taking the top eight burners (4 from each wall) out of service. OSC
operation in this unit is therefore accompanied by a loss in capacity. The
decrease in NO emissions due to staging is substantial: an approximately
A
50 percent reduction compared to the baseline level, and a 40 percent
reduction compared to the low-load, no-staging case.
Unit efficiency is not affected by staging. As seen from Table 6-8,
efficiency seems to depend mainly on the overall excess air level which
controls the dry gas loss. Staging does tend to increase the unburned
combustible losses, but since they form a small part of the total loss,
their effect on unit efficiency is negligible. As mentioned earlier, one
major detrimental effect associated with staging on this unit was a loss in
capacity. Another problem that can occur with staging is that furnace
conditions become unacceptable. In the tests given in Table 6-8 the furnace
conditions, that is, the furnace wall and flame conditions, were monitored.
During the baseline case (column 1) the flames were bright and clear, and
the furnace walls were clean with slight accumulation of dry and sponge
ash. As the burner stoichiometry was decreased the flames became hazier and
started to fill the furnace. Also, slag accumulation increased and it
became more plastic and started to run in certain spots on the furnace.
6-25
-------
TABLE 6-8. COMPARISON OF PROCESS VARIABLES FOR A HORIZONTALLY OPPOSED COAL-FIRED BOILER
AT BASELINE AND LOW NOX CONDITIONS (APPENDIX B): UNIT A
•
Process Variables
NCR
Main steam flow
Reheat steam flow
Furnace excess air
Burners out of service
Boiler drum pressure
Superheat steam pressure
Reheat steam pressure
Superheat steam temperature
Reheat steam temperature
A1r flow leaving AH
Gas flow entering AH
A1r entrance temperature
Air leaving AH temperature
Gas leaving AH temperature
Gas leaving economizer temperature
Furnace draft
Fuel burned rate
Volumetric heat release rate
Surface heat release rate
Heat losses
Dry gas
Hydrogen and moisture in fuel
Moisture in air
Unburned combustible
Radiation
Unaccounted for
Total losses
Efficiency
X
Mg/hr (1Q3 lb/hr)
Mg/hr (103 lb/hr)
X
Number out
MPa (ps1)
MPa (ps1)
MPa (psi)
K (OF)
K (0F)
Mg/hr (103 lb/hr)
Mg/hr (103 lb/hr)
K (0F)
K (0F)
K (0F)
K (°F)
Pa (in. HgO)
Mg/hr (103 lb/hr 1
kW/m2 (Btu/hr-ft3)
kW/m2 (Btu/hr-ft2)
%
%
%
I
Baseline
100
420 (3333)
379 (3010)
20.0
0
17.24 (2500)
16.41 (2380)
3.61 (524)
806.5 (992)
795.9 (973)
492 (3905)
533 (4234)
340.9 (154)
519.3 (475)
417.6 (292)
665.4 (738)
N.A.
45.99 (365)
154.26 (14915)
218.01 (69155)
3.921
4.398
0.094
0.221
0.190
0.500
9.324
90.676
II
High Excess Air
100
419 (3327)
370 (2940)
35.0
0
17.29 (2507)
16.53 (2398)
3.56 (517
807.0 (993
805.9 (991
568 (4505)
610 (4843)
337.0 (147)
530.9 (496)
417.6 (292)
669.3 (745)
N.A.
47.75 (379)
153.75 (14865)
217.51 (68997)
4.644
4.623
0.112
0.230
0.190
0.500
10.299
89.701
III
Load Reduction (LR)
76
321 (2550)
290 (2300)
19.5
0
17.56 (2547)
16.53 (2397)
2.83 (410)
811.5 (1001)
805.4 (990)
382 (3034)
415 (3294)
347.6 (166)
528.7 (492)
414.3 (286)
641.5 (695)
3359 (13.5)
36.16 (287)
121.36 (11734)
171.50 (54402)
3.403
4.204
0.083
0.155
0.250
0.500
8.595
91.405
IV
BOOS (w/LR)
78
328 (2600)
296 (2350)
20.5
8
17.15 (2488
15.69 (2275
2.83 (410
799.8 (980)
785.9 (955)
400 (3175
435 (3451
348.2 (167
526.5 (438
414.3 (286
643.2 (698
2986 (12)
38.18 (303)
122.08 (11803)
172.52 (54726)
i
3.539
4.304
0.086
0.414
0.250
0.500
9.093
90.907
I
ro
-------
TABLE 6-8. Concluded
Process Variables
Coal ultimate analysis
Ash
H2
C
H20
N£
02
Heating value
Flue gas analysis
CO?
.5?
Gas emission data
H02
S02
CO
X by wt.
X by wt.
X by wt.
X by wt.
X by wt.
X by wt.
X by wt.
kJ/kg (Btu/lb)
X by vol .
ng/J (lb/106 Btu)
ppm
ppn
I
Baseline
9.68
2.59
4.52
65.56
7.10
1.42
9.13
27630 (11879)
13.942
8.701
0.207
471 (1.1)
2209
115
II
High Excess Air
10.72
3.18
4.58
64.09
6.90
1.31
9.22
26558 (11418)
12.355
8.105
0.230
496 (1.2)
1908
100
III
Load Reduction (LR)
9.33
2.86
4.28
65.42
7.88
1.26
8.97
27628 (11878)
14.101
8.591
0.231
392 (0.91)
2503
44
IV
BOOS (w/LR)
8.82
2.75
3.99
65.27
9.37
1.35
8.45
26393 (11347)
14.103
8.469
0.224
236 (0.55)
2554
128
ro
-------
Foster Wheeler judged the conditions in the furnace to be unacceptable at
burner stoichiometries of approximately 90 to 95 percent — the estimated
level during the test in column 4. Hence, although OSC operation of this
unit resulted in substantial decreases in NOX emissions, it caused a
25 percent derating of the unit and furnace conditions deemed unacceptable
for long term operation.
In summary, test results on horizontally opposed coal-fired boilers
indicate that low excess air, staging and low NOX burners are all
successful in reducing NO emissions without major adverse effects.
A
However, staging by taking burners out of service often does lead to boiler
derating. Problems with slagging may also arise under OSC operation.
Moreover, although corrosion does not seem to be a problem from results of
short-term tests on corrosion coupons, the question cannot be definitely
resolved until results of long-term corrosion tests on furnace tubes become
available.
Other methods, such as flue gas recirculation, have also been found
to reduce NOV in these design units. However, the reduction in NO
X A
emissions by these methods is generally much smaller, due to the effects of
fuel nitrogen, than that by staged combustion combined with low excess air
firing. Moreover, additional testing would be required to ascertain that no
side effects are associated with use of the other methods.
6.4 SINGLE WALL COAL-FIRED BOILERS
A number of coal-fired boilers with burners located on one wall have
been tested under low NO operation and compared with operation under
A
baseline conditions. The effect of low NO operation on furnace slagging,
corrosion, efficiency, carbon loss, particulate emissions, etc., have been
investigated on some units. In general, the effect of low NO operation
A
on single wall coal-fired boilers is not expected to be substantially
different from that discussed above for horizontally opposed coal-fired
boilers.
The Tennessee Valley Authority has conducted extensive tests on its
124 MW, B&W, rear wall fired, Widows Creek Unit No. 5 (Reference 6-6). A
sister unit, the Widows Creek Boiler No. 6, was used as a control for
long-term corrosion tests. Tests were carried out over the whole range of
boiler loads. Typical NO baseline emissions varied from about 560 ppm at
A
full load (125 MW) to 320 ppm at 50 MW. The combustion modifications
6-28
-------
employed to reduce NO emissions were taking burners out of service and
/>
lowering overall excess air levels. NO reductions from 30 to 50 percent
n
were achieved by applying these methods in combination. No boiler derating
was encountered if only 2 burners out of a total of 16 were taken out of
service at full load. At partial loads more burners could be operated on
air only. However, at lower loads BOOS operation was not as successful in
reducing NO and at 50 MW, where NO levels were already quite low, BOOS
/\ ^
operation resulted in an increase in NOX emissions.
Particulate emissions from the Widows Creek No. 5 boiler increased
under low NO conditions, but the increased amounts were not considered
significant. Carbon loss in flyash, however, increased by about 30 percent
at full loads. Efficiency was also adversely affected by low NO
operation, decreasing by about 1 percent at full load and by about 0.7
percent at 50 MW. The results from corrosion tests were inconclusive.
Corrosion was estimated both by the use of corrosion coupons and by actual
measurement of tube wall thicknesses in Unit No. 5, which was operated under
low NO conditions, and Unit No. 6, which was used as a control boiler.
A
The results from the corrosion coupon tests might be invalid due to possible
weight loss during acid cleaning. The wall thickness measurements are also
subject to uncertainty, due to suspected errors in instrument calibration at
the control boiler. Also, the tests were of short term duration
(approximately 6 weeks). The low NO boiler showed a corrosion rate of
^
about 40 mils/year on the side wall and about 12 mils/year on the division
wall, as deduced from wall thickness measurements. These rates are an order
of magnitude higher than the 1 to 3 mils/year corrosion rates experienced by
furnace walls under normal firing conditions.
The boiler used for control purposes in the above tests, Widows Creek
Unit No. 6, was also tested for low NO in an Exxon study
A
(Reference 6-2). Baseline NO emissions from this boiler averaged 634 ppm
o
at full load. A 10 percent reduction in excess air reduced NO emissions
by 25 percent at full as well as reduced load under normal firing
operation. The same percentage reduction in stoichiometric air to active
burners under staged conditions reduced NO emissions by an average of
A
24 percent at full load and 28 percent at reduced load. Low NOX operation
of the boiler did not result in a significant change in efficiency. The
percentage of carbon in flyash, however, increased from an average of 6.1 at
6-29
-------
baseline to 10.5 under low NOV firing. The particle loading actually
6
decreased from an average of 2.7 yg/J (6.3 lb/10 Btu) at baseline to
2.1 yg/J (4.8 lb/106 Btu) at low N0x operation. No corrosion tests were
carried out for this boiler. No other side effects were noted under
modified combustion.
In the same study (Reference 6-2) a 320 MW Foster Wheeler boiler,
Crist Station No. 6, operated by Gulf Power Company was tested for low
NO emissions under LEA and BOOS. Reducing excess air to the burners,
A
with or without staging, had a significant effect on NO emissions. The
/\
largest reduction in emissions occurred with the top row of 4 burners out of
service from a total of 16. Under those conditions, NOX emissions dropped
from a baseline of 845 ppm to a low of approximately 520 ppm. The load
capacity of the boiler, however, also decreased by about 25 percent. Some
data are available for this boiler on the effect of low NO operation on
other process variables. Particulate loading increased from 1.87 yg/J
(4.34 lb/106 Btu) at baseline to 2.77 yg/J (6.45 lb/106 Btu) under low
NO operation. The carbon loss in flyash increased from 5.08 percent at
/\
baseline to 8.15 percent at low NO conditions. The efficiency of the
/\
boiler changed from 88.5 percent at baseline to 88.1 percent at low NO
^
conditions.. No data were available on corrosion data for this boiler.
The Widows Creek Unit No. 5 tested by TVA was also tested by Exxon in
another study (Reference 6-7). Baseline NO emissions at full load for
A
this boiler were 567 ppm at full load (125 MW) and 506 ppm at partial load
(100 MW). These values are lower than the baseline emissions from the
Widows Creek's sister Unit No. 6. The differences are attributed to the
difference in coals fired and the fact that Unit No. 5 was much cleaner when
tested than Unit No. 6. Low NO testing of Unit No. 5 involved staging,
n
lowering overall excess air levels and changing burner register settings.
For both load levels, lowest NO emissions were obtained by taking burners
^
out of service from the top row, and setting the secondary air registers on
the active burners at 20 percent open. Setting the air register at
20 percent open on the active burners reduces the amount of air available in
the primary combustion zone, thus increasing the off stoichiometric effect
of the staging process. It was also found that the overall excess air
levels could be reduced to a greater extent when the active burner registers
were at 20 percent open compared to 60 percent open. The minimum overall
6-30
-------
excess oxygen levels attainable, subject to the constraint of maintaining
carbon monoxide emissions below 200 ppm, was about 3 to 3.5 percent under
staged firing. At 125 MW, two burners could be fired on air only without a
reduction in load, which resulted in NO levels as low as 468 ppm. At 100
MW, four burners could be placed on BOOS and caused NOX levels to drop to
317 ppm.
The Unit No. 5 was also tested for particulates, carbon loss,
corrosion and efficiency. The particulate emissions decreased from a
baseline average of 2.3 yg/J (5.3 lb/106 Btu) to 1.9 yg/J (4.4 lb/
10 Btu) under low NO conditions. The carbon loss on flyash also
A
decreased from an average of 11.1 percent at baseline to 7.1 percent under
low NOX firing. The corrosion rates as measured by corrosion coupons
showed a slight increase (less than 3 mils/year) in corrosion rates due to
low NO firing. Finally, efficiency increased by an average of 1 percent
n
when the boiler was operated under low N0¥ conditions. This was most
^
likely due to the reduced levels of overall excess air maintained during the
low NO tests. Note that these results are quite different from those
^
obtained by the TVA tests on the same boiler. The effects of low NOX
operation on particulate emissions, carbon loss, and efficiency are exactly
the opposite of those found in the TVA tests. In addition, the corrosion
rates, although increasing in both series of tests under low NOX
conditions, show a much smaller increase in the Exxon tests.
The Exxon study (Reference 6-7) also reported the results of tests on
the 270 MW, Foster Wheeler, single wall fired, Public Service Electric and
Gas Company (New Jersey), Mercer Station Boiler No. 1. This boiler is a wet
bottom unit and has limited operational flexibility. The baseline NOX
emissions were 1383 ppm, which is not uncommon for this type of unit. The
furnace floor is relatively close to the bottom row of burners so that high
gas temperatures are maintained in the lower part of the furnace which keeps
the slag in a molten state. Lowering excess air had the greatest effect on
NOX emissions. NOX levels were reduced by 24 percent by this method.
Biased firing, which was accomplished by firing top row burners fuel lean
and bottom and middle row burners fuel rich, reduced NOX by only
16 percent from baseline. No derating occurred due to biased firing. Low
NOX operation of the boiler increased particulate emissions slightly from
1.1 yg/J (2.6 lb/106 Btu, average) at baseline to 1.2 yg/J
6-31
-------
(2.9 lb/10^ Btu). The carbon loss also increased from an average of 1.9
percent at baseline to 3.5 percent under low NO conditions. The
A
participate size distribution was not affected significantly by low NO
A
operation. Corrosion rates as measured by corrosion coupons also showed no
significant difference between baseline and low NOX operations. The
efficiency of the unit also did not seem significantly affected by low NO
X
operation.
Data have recently become available on two front wall coal-fired
units manufactured by Foster Wheeler Energy Corporation (Appendix B). One
of the units is of a pre-NSPS design, Unit B, while the other unit is
designed to meet NSPS requirements, Unit C. Both units employ the old
standard FWEC Intervane Burner which produces a high turbulence, high
intensity flame. (Newer units are being installed with the FWEC low NO ,
A
dual register burners which produce a reduced turbulence flame). Although
direct comparisons between the two units are difficult due to the different
capacities and types of coal fired, the differences in NO emissions may
be largely attributed to changes in the design of the NSPS Unit. The most
significant of these changes are the larger furnace design to provide lower
burner zone liberation rates, and the inclusion of OFA ports to provide OSC
operation without derating.
The pre-NSPS design, Unit B, has an MCR of 292 kg/s (2.32 x
106 Ib/hr) of superheated steam and 256 kg/s (2.03 x 106 Ib/hr) of
reheated steam. Under those conditions the unit has a burner zone heat
liberation rate of approximately 1.17 MW/m2 (370 x 103 Btu/hr-ft2).
Emissions and some process data for the unit are shown in Table 6-9 for
different operating conditions. The major variables are amount of excess
air, load condition and degree of staging. Column 1 shows the near baseline
conditions with 93 percent of MCR, 26.8 percent excess air and all 16
burners in service (no staging). The effect of reducing load to 75 percent
MCR while maintaining other conditions invariant is given in column 2, where
the N0¥ emissions decrease by 10 percent. The effect of excess air at
A
this reduced load is shown in column 3 where the excess air is increased to
53.7 percent which brings the NO emissions back close to the baseline
J\
level.
The effect of OSC operation by taking the 4 burners on the top tier
out of service is given in columns 4 through 8. One of the major impacts of
6-32
-------
this operation is to derate the unit to 75 percent of its capacity. OSC
also results in substantial NO reductions, even at relatively low degrees
A
of staging. For example, the conditions of the test shown in column 4 are
similar in load and excess air level to the test in column 2. The column 4
test has four upper burners out of service, but the idle registers are
closed so that there is nominally no staging. However, the leakage of air
from the out of service burners* is sufficient to create a staging effect,
and this is demonstrated in the 22 percent drop in NOX emissions from
column 2 to column 4. Further reductions in NOX emissions occur as the
degree of staging is increased by opening the idle registers, first to
10 percent as in column 5 and then to 50 percent as in column 6. For these
tests the load and excess air levels are similar to those in columns 2 and
4. The maximum decrease in NO emissions due to OSC alone (column 2
versus column 6) is 59 percent and due to combined OSC and low load (column
1 versus column 6) is 63 percent.
The effect of overall excess air level under OSC operation is shown
by comparing columns 5 and 7, and comparing columns 6 and 8, for idle
register settings of 10 and 50 percent, respectively. As expected N0x
emissions increase with excess air. Also, the sensitivity of NOX
emissions to excess air increases with increasing burner heat liberation
rates. This is demonstrated in Table 6-10 where the change in NOX
emissions with excess air is shown for different burner liberation rates.
Since BOOS firing usually involves operating the remaining burners at high
heat release rates, the overall excess air levels during BOOS operation need
to be carefully controlled. The sharp rise in N0x emissions with
increasing excess air at high burner heat release rates is probably due to
increased flame turbulence as burner throat velocities exceed their design
values.
During the tests reported in Table 6-9, the boiler was monitored for
adverse effects on unit performance and operation. Carbon monoxide and
unburned combustibles did not increase significantly during the low N0x
tests. However, there was a problem with slagging at high degrees of
staging. During the normal firing tests (columns 1 through 3 in Table 6-9),
*Boiler designs usually allow for some leakage of air through out-of-service
registers to keep the burner components cool.
6-33
-------
TABLE 6-9. COMPARISON OF PROCESS VARIABLES FOR A PRE-NSPS FRONT WALL COAL-FIRED BOILER
AT BASELINE AND LOW NOV CONDITIONS: UNIT B
x
CT)
co
Test Variable
Load t NCR
Excess air t
Burners out of service
Idle registers % open
Steaa flow rate kg/s (103 Ib/hr)
Fuel flow rate kg/s (103 Ib/hr)
Air leaving AH kg/s (103 Ib/hr)
Gas entering AH kg/s (103 Ib/hr)
Coal ultlMte analysis:
Ash by Height
S by weight
H2 by wight
C by weight
H20 by Height
N2 by Height
02 by Height
Heating value kj/kg (Btu/lb)
Flue gas analysis:
COj, S by volme
H20 t by voluw
SO; S by voluae
Gas emission data:
NO; ng/J (lb/10* Btu)
S02 ng/J (lb/106 Btu)
1
Baseline
93
26.8
0
-
271
(2150)
37.7
(299)
380.5
(3020)
414.4
(3289)
10.06
0.66
4.29
60.54
9.2S
1.20
14.00
24137
(10377)
13.549
9.033
0.055
619
(1.44)
507
(1.18)
2
Load
Reduction
75
26.8
0
--
227
(1800)
32.3
(256)
325.7
(2585)
354.8
(2816)
10.06
0.66
4.29
60.54
9.25
1.20
14.00
24137
(10377)
13.549
9.033
O.OSS
555
(1.29)
456
(1.06)
3
High Excess
A1r Load
Reduction
75
53.7
0
-
220
(1745)
34.5
(274)
383.9
(3047)
413.5
(3282)
13.86
0.55
3.77
56.60
8.18
1.09
15.95
22176
(9534)
11.601
7.670
0.042
632
(1.47)
512
(1.19)
4
BOOS
75
27.5
4
0
227
(1800)
29.7
(236)
313.9
(2491)
340.2
(2700)
11.38
0.66
5.17
59.90
9.60
1.22
12.07
23653
(10169)
12,796
9.876
0.053
423
(1.00)
494
(1.15)
S
BOOS
75
28.2
4
10
227
(1800)
30.0
(238)
318.3
(2526)
344.9
(2737)
11.38
0.66
5.17
59.90
9.60
1.22
12.07
23653
(10169)
12.730
9.836
0.053
387
(0.90)
469
(1.09)
6
BOOS
75
24.5
4
50
227
(1800)
29.5
(234)
303.9
(2412)
330.1
(2620)
11.38
0.66
5.17
59.90
9.60
1.22
12.07
23653
(10169)
13.086
10.053
0.054
228
(0.53)
473
(1.10)
7
BOOS
75
53.3
4
10
227
(1800)
31.6
(251)
401.4
(3186)
429.4
(3408)
11.38
0.66
5.17
59.90
9.60
1.22
12.07
23653
(10169)
10.746
8.622
0.044
533
(1.24)
503
(1.17)
8
BOOS
75
56.7
4
50
227
(1800)
30.2
(240)
383.8
(3046)
410.6
(3259)
11.38
0.66
5.17
59.90
9.60
1.22
12.07
23653
(10169)
10.524
8.486
0.043
456
(1.06)
494
(1.15)
-------
TABLE 6-10. COMPARISON OF SENSITIVITY OF NOX EMISSIONS TO CHANGES IN EXCESS AIR LEVELS
WITH INCREASING BURNER HEAT LIBERATION RATES: UNIT B
O1
Test Numbers
Load
Burners In Service
NOX change per percent
excess air Increase
Idle Registers
Burner Heat Liberation MW
% HCR
ng/J/X (Ib/ltf Btu/X)
% Open
(106 Btu/hr)
5-7
75
16
2.87 (6.67E-3)
--
48.6 (166)
2-4
75
12
5.85 (13.6E-3)
10
58.6 (200)
1-3
75
12
7.09 (16.5E-3)
50
58.6 (200)
-------
the furnace side and rear walls had a covering of dry ash which is
considered normal for this unit. The flames were bright and stable. In the
low NO tests the slag and flame conditions varied with degree of
/\
staging. For example, in the test shown in column 8 of Table 6-9, the idle
registers were open 50 percent and the overall excess air level was about 60
percent. The stoichiometry at the active burners was approximately
120 percent. The furnace conditions in this test were about the same as
during normal firing. As excess air was lowered, increased slag formation
occurred in the lower furnace. At 25 percent excess air and 50 percent open
idle registers (column 6 of Table 6-9) slag was running rapidly on the rear
wall creating conditions unacceptable for continuous operation. The active
burner stoichiometry for this test was approximatey 90 to 95 percent. Flame
conditions also became hazy at lower burner stoichiometry, although they
were stable. Due to the problems associated with high degrees of staging,
FWEC recommends that burner stoichiometry be maintained above 95 percent in
their units (Appendix B).
The NSPS design, Unit C, has an MCR of 117 kg/s (9.29 x 105 Ib/hr)
5
of superheated steam and 93.2 kg/s (7.40 x 10 Ib/hr) of reheated steam.
o
The unit is designed with a low burner zone liberation rate of 678 kW/m
(215 x 103 Btu/hr-ft2) (Cf. the pre-NSPS unit B which has a value of
1170 kW/m2 (370 x 103 Btu/hr-ft2) at MCR). The unit is also equipped
with OFA ports. Hence, OSC Operation is possible without loss in capacity.
Test data for emissions and process variables are shown in Table 6-11. The
major variables tested are amount of excess air, load conditions and degree
of staging. The first four columns represent tests without OSC. The OFA
ports were kept closed during these tests. The test in column 1 is a
baseline test at full load and about 20 percent excess air. The effect of a
change in load, while maintaining excess air levels approximately constant,
is shown by columns 2 and 3 which are at 93 and 68 percent MCR,
respectively. NO emissions are seen to drop by approximately 30 percent
A
for a 30 percent decrease in load. The effect of excess air with no staging
is shown in column 4, which has the same load condition as the test in
column
2. The change in NOX with excess air is 6.23 ng/J (14.5 10"3 lb/106 Btu)
for each percent change in excess air at near maximum load conditions.
The data for Unit C operation under OSC are shown in columns 5
through 7. In all these tests the OFA ports were 100 percent open. Column
6-36
-------
TABLE 6-11.
COMPARISON OF PROCESS VARIABLES FOR AN NSPS FRONT WALL COAL-FIRED BOILER AT
BASELINE AND LOW NOX CONDITIONS: UNIT C
Process Variables
CR
lain stew flow
leheat steam flow
;urnace excess air
Jverfire airport
loller drtM press.
Superheat steam press.
eheat steam press.
uperheat stean temp.
eheat steam temp.
Mrflow leaving AH
as flow entering AN
Mr entrance temp.
Air leaving AH temp.
ias leaving AH tenp.
Sas leaving economizer
temp.
Furnace draft
Fuel burned rate
Volumetric heat
release rate
Surface heat
release rate
Heat losses
Dry gas
Hydrogen and
moisture In fuel
Moisture 1n air
Unburned combustible
Radiation
Unaccounted for
otal losses
Efficiency
X
Mg/hr (103 Ib/hr)
Ng/hr (103 Ib/hr)
%
t open
MPa (psl)
MPa (ps1)
MPa (ps1)
K (Of)
K (°F)
Mg/hr (103 Ib/hr)
Mg/hr (103 Ib/hr)
K Of)
K °F
K °F)
K OF)
Pa (1n. H20)
Mg/hr (103 Ib/hr)
kU/m2 (8tu/hr-ft3)
kM/«2 (Btu/hr-ft2)
%
X
X
I
Baseline
100
117 (930)
93 (742)
21.6
0
14.49 (2102)
13.03 (1890)
3.41 (500)
810.9 (1000)
810.9 (1000)
144.0 (1143
155.7 (1236
284 (51
569 (565
428 (311
648 (706)
-150 (-0.60)
46.22 (101.9)
185.34 (17920)
193.44 (61361)
5.896
5.23
0.142
0.35
0.25
0.50
12.368
87.63
II
Load Reduction
92
109 (865)
NA
19.9
0
14.78 (2144)
13.05 (1893)
3.46 (502)
800 (980)
798 (977)
126.5 (1004
137.1 (1088
320 (116
579 (582
451 (353
646 (704
-220 (-0.90)
41 (91)
165.03 (15956)
172.24 (54636)
5.26
5.11
0.13
0.35
0.25
0.50
12.26
87.74
III
Load Reduction
68
80 (637)
NA
20
0
13.81 (2003)
12.96 (1880)
2.50 (363)
800 (980)
800 (980)
94.6 751)
102.6 814)
321 119)
570 566)
443 338)
629 (673)
-170 (-0.70)
31 (68)
123.31 (11922)
128.71 (40827)
4.864
5.06
0.117
0.35
0.25
0.50
11.14
88.86
IV
High Excess
Air
92
108 (861)
NA
33.8
0
14.79 (2145)
13.00 (1885)
3.46 (502)
791 (965)
789 (960)
143.5 (1139)
154.2 (1224)
315 (107)
570 (566)
448 (346)
642 (696)
-200 (-0.80)
42 (92)
167.09 (16155)
174.39 (55318)
5.922
5.094
0.143
0.35
0.25
0.50
12.26
87.74
V
100
118 (935)
94 (746)
25.1
100
14.51 (2104)
13.03 (1890)
3.45 (500)
807 (993)
808 (995)
149.1 (1183)
160.8 (1276)
278 (40
559 (547)
421 (298)
643 (697)
-190 (-0.75)
46.3 (102)
185.45 (17930)
193.55 (61395)
6.03
5.25
0.145
0.35
0.25
0.50
12.51
87.49
VI
92
108 (860)
NA
20.1
100
14.49 (2102)
12.98 (1882)
3.46 (502)
804 (988)
803 (985)
130.2 (1033)
141.0 (1119
316 (110
575 (576
450 (350
647 (705)
-190 (-0.75)
42 (93)
168.90 (16330)
176.28 (55919)
5.356
. 5.088
0.129
0.35
0.25
0.50
11.67
88.33
VII
92
108 (855)
NA
32.3
100
14.77 (2142)
13.03 (1890)
3.45 (500)
812.6 (1003)
805 (990)
144.1 (1144)
155.1 (1231)
311 (100
579 (582
450 (351
651 (713
-250 (-1.0)
43 (94)
170.47 (16482)
177.92 (56437)
6.12
5.176
0.148
0.35
0.25
0.50
12.54
87.46
CO
-------
TABLE 6-11. Concluded
Process Variables
Coal ultlnate analysis
Ash
S
H2
C
H20
N2
°2
Heating value
:1ue gas analysis
C02
HpO
S02
Gas emission data
N02
CO
X by wt.
X by wt.
X by wt.
X by wt.
X by wt.
X by wt.
X by wt.
kJ/kg (Btu/lb)
X by vol.
ng/J (lb/106 Btu)
ppn
I
Baseline
8.10
0.50
5.40
67.78
5.86
1.04
11.32
28426 (12221)
13.621
9.195
0.037
387 (0.901)
NA
II
Load Reduction
7.38
0.24
5.36
67.34
6.92
0.82
11.94
28342 (12185)
13.841
9.451
0.018
358 (0.832)
39
III
Load Reduction
7.38
0.24
5.36
67.34
6.92
0.82
11.94
28342 (12185)
13.830
9.445
0.018
268 (0.623)
26
IV
High Excess
Air
7.38
0.24
5.36
67.34
6.92
0.82
11.94
28342 (12185)
12.606
8.790
0.017
435 (1.011)
35
V
8.10
0.50
5.40
67.78
5.86
1.04
11.32
28426 (12221)
13.259
9.005
0.037
247 (0.535)
NA
VI
7.74
0.37
5.38
67.56
6.39
0.93
11.63
28384 (12203)
13.801
9.360
0.028
184 (0.428)
44
VII
7.74
0.37
5.38
67.56
6.39
0.93
11.63
28384 (12203)
12.453
8.645
0.026
266 (0.619)
35
01
00
00
-------
5 can be compared directly with the baseline case (column 1) since both are
at full load and similar excess air conditions. Also, the test in column 6
can be compared with column 2 for 93 percent load and 20 percent excess air
levels. Similarly, the test in column 7 is also at 93 percent load but at
32 percent excess air, and can therefore be compared to the test in column 4
which has similar conditions except that the latter was without the use of
OFA Ports. From these sets of comparisons, it is seen that NO levels
fi
dropped by approximately 160 ng/J (0.37 lb/10 Btu) on opening the OFA
Ports with all other conditions maintained constant. For the baseline case
this represents a decrease of 35 percent in NO emissions. Also, under
^
OSC operation, the change in NO with excess air level is 6.0 ng/J
(14.0 x 10"3 lb/106 Btu) for each percent change in excess air at near
maximum load (column 6 and column 7), which is comparable to the sensitivity
under normal firing conditions.
The process variables in Table 6-11 do not indicate significant
difference between normal and OSC firing, if variables such as load and
furnace excess air levels remain unchanged. As mentioned earlier, there is
a significant decrease in NO emissions when the OFA ports are opened, but
rt
the carbon monoxide emissions rise only by small amounts (39 ppm in column 2
versus 44 ppm in column 6) for the high load, low excess air condition.
Unit efficiency also remains essentially unchanged. The effect of reduced
load on the process variables is to lower all flowrates, including the
volumetric and surface heat release rates which results in a reduction in
NO emissions. However, since the flowrates decrease in proportion to the
load reductions the unit efficiency is not significantly affected (compare
columns 2 and 3). One the other hand, a change in excess air levels changes
the air and gas flowrates through the system for a given fuel flowrate. The
dry gas heat losses increase with an increase in excess air causing the unit
efficiency to decrease (compare columns 2 and 4, and also 6 and 7). Reduced
excess air levels are therefore desirable both for high efficiency and low
NO emissions. However, very low levels can lead to unacceptably high
A
carbon monoxide emissions. Also, low excess air levels combined with OSC
operation can result in highly substoichiometric burner conditions which
could cause problems with slagging and corrosion. Nevertheless, during the
tests given in Table 6-11, slagging conditions were monitored and were found
to be minimal both under normal and OSC operation. The overall excess air
6-39
-------
levels were always maintained at about 20 percent or above in those tests.
Carbon monoxide emissions are also all within 50 ppm at those excess air
levels.
Comparing the tests on Unit B and Unit C, the newer NSPS boiler (Unit
C) is seen to have lower NO emissions. A direct comparison is not
y\
entirely justifiable due to the difference in the types of coal used in the
tests. However, the lower NO emissions in the newer boiler are at least
A
partly attributable to the larger furnace design and consequently lower
burner zone liberation rates. The newer unit is also equipped with OFA
Ports which avoids taking burners out of service during OSC operation. No
derating therefore occurs during low NO operation of the newer unit.
A
In summary, the single wall coal-fired boilers tested to date show
somewhat varied effects due to low NO operation. Boiler derating may
A
occur in some units where OSC operation is performed by taking burners out
of service. Efficiency losses due to low NO operation can be minimized
A
if it is possible to maintain low overall excess air levels without
excessive carbon monoxide generation. Increased carbon losses can probably
be reduced or eliminated by ensuring proper air distribution to each
burner. Particle loading may be expected to increase slightly in certain
cases, but there is no evidence of a shift to smaller particle size.
Slagging problems may also occur with OSC operation. On corrosion, the data
are inconclusive. Corrosion rates seem to increase under low NO
)\
operation, but the extent or severity of the increase cannot be estimated at
the present time. Long-term accurate tests on actual furnace water tubes
are required to resolve the discrepancies observed on tests with corrosion
coupons. Ongoing tests on a horizontally opposed coal-fired unit should
help resolve the matter. No other major adverse effects are expected from
low NOX operation on wall fired coal burning boilers.
6.5 TURBO FURNACE COAL-FIRED BOILERS
The unique configuration of turbo furnace boilers is designed to
produce lower NO emissions than uncontrolled wall fired boilers. Rawdon
A
and Johnson have presented general papers on the performance of turbo
furnaces (References 6-17 and 6-18). Published emission and process data of
sufficient completeness from turbo furnaces have been minimal to date.
Due to the special design of the turbo furnace, certain combustion
modification techniques can be tried on these boilers which would be
6-40
-------
difficult to implement on other designs. In particular the burners, which
are directed at an angle downwards from horizontal, are equipped with
velocity dampers and directional vanes by which the flow and direction of
the combustion air can be independently controlled above and below the
burner centerline. By changing the positions of the dampers and the vanes,
it is possible to simulate overfire air injection in certain cases.
The furnace design also includes a throat or waist section. The
flame basket is generally held in the lower part of the furnace below this
throat. By installing overfire air ports above the throat section, it is
possible to separate the combustion process into two distinct zones. Thus,
if the lower part of the furnace is maintained very rich, a precombustion
zone may be simulated. However, such conditions may also lead to an
increased tendency to smoke with oil fuels. BOOS is not very effective as a
NO control technique for these boilers due to the horizontal inline
/\
arrangement of the burners.
A series of tests were performed by Exxon on the Big Bend No. 2 turbo
furnace burning pulverized coal to establish the NO reduction capability
A
of combustion modification (Reference 6-2). Controls that were investigated
consisted of low excess air, staged combustion with burners out of service,
and directional changes of the combustion air vanes.
By far, the most effective technique was LEA resulting in about a
20 percent NO reduction. Excess air was reduced from 15 to 7 percent at
rt
near boiler steam generating capacity. No adverse boiler operating
condition was reported for the LEA test. On the contrary, a slight net
increase in thermal efficiency, although not reported, may be suspected from
the reduction in flue gas stack temperature. Staged combustion with BOOS
resulted only in slight NO reduction (10 percent), however, at a penalty
A
of derating the boiler. When 8 of the 24 available burners were set on air
only, the impact on boiler capacity was a reduction from 380 to 230 MW.
Directional changes of combustion air vanes on this boiler also resulted in
a slight NO reduction (11 percent) but, when combined with LEA, it proved
n
to be the optimum retrofit control system for this unit with a total NO
^
reduction of 27 percent at maximum operating load. The combination of LEA
and directional air vane changes had no reported adverse impact on boiler
operation. In fact, it may have caused a slight increase in efficiency as a
result of lower flue gas stack temperatures.
6-41
-------
6.6 TANGENTIAL OIL-FIRED BOILERS
Some process data are available on tangential oil-fired boilers from
utility companies in California. The California utilities have gained
considerable experience in NO reduction from oil- and gas-fired
A
equipment. Generally tangentially fired boilers have lower NO emissions
A
than wall fired boilers due to a lesser degree of flame interaction and
lower flame intensity. In some cases, therefore, simple modifications such
as low excess air operation have been sufficient to bring NO emissions
A
down to acceptable levels. In other cases the usual combustion
modifications used with oil and gas fuels, such as flue gas recirculation
and off stoichiometric firing, have, in general, reduced NO emission to
A
the desired values.
Table 6-12 shows a comparison of some process variables on South Bay
Boiler No. 4 operated under baseline and low NO conditions at partial
/\
load (Reference 6-8). These tests were conducted prior to recent burner
modifications implemented to increase the efficiency of the boiler while
still meeting Southern California NO emission standards). These burner
A
modifications are discussed later in this section following an analysis of
the data presented in Table 6-12. The South Bay Boiler No. 4 is a
Combustion Engineering tangentially fired cycling boiler, operated by San
Diego Gas and Electric Company. The boiler has three levels of burners and
can generate up to 198 kg/s (1.57 x 106 Ib/hr) of steam at 786K (955°F)
with a maximum drum pressure of 16 MPa (2300 psig). The furnace has a
straight-through configuration without a gas/air heat exchanger. The
combustion air is heated up to 394K (250°F) by means of steam coils.
Under normal operation of the boiler the excess air level was maintained at
levels higher than design values due to formation of local smoke pockets at
lower excess air levels. That was found to be due to maldistribution of air
at the burners. By closing the auxiliary air dampers down to around 90
percent while leaving the fuel air dampers 100 percent open, a more uniform
airflow distribution was obtained which allowed operation at lower excess
air levels. Baseline and low excess air operation are summarized in the
first two columns of Table 6-12. The adjustment in the auxiliary air damper
setting allowed excess oxygen to be decreased from 7.5 to 3.3 percent.
In addition to decreasing NOX by 17 percent, the boiler efficiency
increased, as evidenced by the decrease in stack gas temperature, and the
6-42
-------
TABLE 6-12. COMPARISON OF SOUTH BAY UNIT NO. 4 UNDER BASELINE AND LOW NO
CONDITIONS UNDER PARTIAL LOAD (Reference 6-8) x
I
.£»
CO
Process Variables
Load
Excess Oxygen
Burners Out of Service
Burner Tilt
Flowrates:
Steam
Fuel 011
Temperatures:
SH Steam
RH Steam
AH A1r Out
Stack Gas
011 Supply
Pressures:
Steam Drum
Oil at Burner
Furnace
Wlndbox/Furnace Differential M
E
MM
Percent
Degrees
kg/s (106 Ib/hr)
kg/s (103 Ib/hr)
K (°F)
K (°F)
K (°F)
K (°F)
K (0F)
MPa (ps1)
MPa (ps1)
kPa (1n. H20)
kPa (In. H20)
kPa (In. H20)
Baseline
173
7.5
None
+27
146 (1.16)
1.26 (10.0)
772 (930)
761 (910)
375 (216)
658 (724)
365 (19B)
12.1 (1760)
1.03 (135)
1.6 (6.4)
1.5 (6.2)
1.4 (5.7)
Low Excess
Air
183
3.3
None
+27
155 (1.23)
1.27 (10.1)
773 (931)
755 (899)
386 (236)
638 (688)
366 (200)
13.0 (1870)
1.05 (138)
1.0 (4.0)
2.5 ( 10)
2.2 (8.8)
Off Stoichlometrlc
Combustion
175
5.5
a from top tier
+30
145 (1.15)
1.22 (9.7)
780 (944)
760 (908)
381 (227)
645 (702)
366 (200)
13.1 (1890)
1.22 (162)
1.2 (4.8)
2.5 (9.9)
2.5 ( 10)
-------
TABLE 6-12. Concluded
CT>
I
Process Variables
FD Fans:
Discharge Pressure
Current
Fuel Air Damper
Auxiliary A1r Damper
Emissions:
NOX (at 3* 02)
CO
Ringleman Smoke Density
kPa (In. H20)
Amps
Percent open
Percent closed
ppm
ppm
ppm
Baseline
4.85 (19.5)
377
100
0
200
10
0
Low Excess
Air
4.28 (19.2)
325
100
87
166
7.5
0
Off Stolchiometric
Combustion
5.67 (22.8)
352
100
90
197
2.5
0.25
-------
increase in power output for approximately similar rates of fuel
consumption. Lower excess air also had the advantage of reducing fan power
consumption. The NO emissions for this boiler and the excess 0?
/\ ^
requirements over a range of loads is shown in Figure 6-5 for both baseline
and low excess air operation. Figure 6-6 shows the fuel consumption and the
stack gas temperature for these two modes of operation. For this boiler, a
reduction in excess air levels from about 6 to 3 percent at full load led to
a decrease in fuel consumption of approximately 5 percent.
Low NO techniques other than LEA were also tried on the South Bay
A
Boiler No. 4 and were found to be less effective. The results of OSC
operation by taking two burners out of service on opposite corners of the
top tier are shown in column 3 of Table 6-12. Although the excess air level
was lower than the baseline value due to more uniform air distribution
obtained by closing the auxiliary air dampers 90 percent, the excess air
level was higher than that at the low excess air mode of operation.
Moreover the NO level was only marginally lower than the baseline
n
emission. OSC operation was, therefore, not recommended as a NOX control
technique for this boiler. Reduced air preheat (RAP) was also tested as a
means to control NO on this boiler. The boiler has a unique arrangement
A
where steam coils are used to heat the combustion air. Reducing air
preheat, therefore, does not lead to a direct increase in stack gas
temperature and corresponding loss in boiler efficiency which would be the
case in more common arrangements where flue gas is used to heat combustion
air. However, for oil fuel no consistent trends in NO emissions were
A
obtained which could be attributed to RAP.
Another unique aspect of this cycling boiler was its capability to
operate over a large range of steam drum pressures. The drum pressure was
accordingly varied to test whether it would have an effect on NOX
emissions. It was found that NO emissions increased with increasing
A
steam pressure at high loads and decreased with increasing steam pressure at
low loads. It was not recommended that this method be used for NO
A
control. Finally, burner tilt was tested to test its effect on NO
n
emissions. The tests were carried out with the steam coil air heater out of
service due to a malfunction in its operation. The normal position of
burner tilt was +30 degrees on oil fuel for steam temperature control.
NO emissions were not affected by burner tilt in the range tested from
6-45
-------
cr>
i
cr>
CM
o
I
>
c?
a
O
CO -~>
w >,
-------
0)
+J
n
16
14
12
4/.10
•* 8
6
*
t
700 .
2 2 65°
fi 3
ro+J
01 £ 600
550
130
-C
110 -
90 -
! 70 -
50 -
s-
^
800 -
700 _
600 _
500 _
80
5.3% reduction
4.5% reduction
Normal operation (baseline) ___
Low excess air
100
120
140
160
180
200
220
240
Load (MW)
Figure 6-6. Comparison of oil consumption and stack gas temperature under baseline
and low excess air conditions for South Bay Unit No. 4 (Reference 6-8).
-------
+28 to +8 degrees. It was found, however, that under normal operating
conditions (i.e., high excess air operation due to nonuniform airflow), the
minimum excess oxygen level decreased with decreased burner tilt. As no
tests were carried out under low excess air condition (uniform airflow) and
with the air heaters in service, the effect of burner tilt under these
conditions is not known. In general, operating variables such as steam
pressure or burner tilt, which affects steam temperature, are not
recommended for NOV control due to the potential impact of these variables
A
on plant operation and efficiency.
Subsequent to these tests investigated for the South Bay Unit 4, San
Diego Gas and Electric Company recently installed new burners from an
English manufacturer, designed to fire efficiently with low excess air. The
objective of the burner retrofit was to improve the boiler efficiency while
still maintaining the NO emission standard of 225 ppm at 3 percent (L.
A £
A system-wide study by the utility had revealed that the efficiency of South
Bay Unit 4 was significantly less than optimum. Although capable of meeting
local NO regulations without combustion modifications, excess air
A
requirements were high (15 to 20 percent versus a design of 10 percent) and
steam temperatures were lower than normal (772K (930°F) versus a design of
783K (950°F) ).
The LEA burner retrofit program by the utility has resulted in a
2 percent increase in boiler efficiency due to low excess air operation and
improvement in steam temperature to design conditions. The effect of the
new burners on NO emissions and flue gas oxygen requirements are
A
illustrated in Figure 6-7. As shown, the LEA burners can meet the NO
A
standard without further control over a wide boiler load range up to 200 MW
or 87 percent of boiler capacity. Above this load, overfire air injection
is utilized to keep NO below the 225 ppm limit (Reference 6-19).
A
Pacific Gas and Electric Company has retrofitted its Pittsburg Boiler
No. 7 with overfire air ports capable of supplying 20 percent of the
combustion air, and a flue gas recirculation system designed to dilute the
oxygen in the windbox to 17 percent (Reference 6-9). This Combustion
Engineering boiler, with five levels of tangential burners, has a design
capacity of 677 kg/s (5.36 x 106 Ib/hr) of steam. The baseline NO
A
emissions are approximately 400 ppm at full load. With NOX control
techniques (FGR + OSC) the emissions decrease to about 280 ppm. PG&E,
6-48
-------
Original w/overfire
air
Original w/overfire air
GROSS LOAD - MW
250
' Preliminary Test Data
Figure 6-7.
NQX emissions and excess flue gas oxygen
requirements of new LEA burners retrofitted on
South Bay Unit 4. (Reference 6-19).
6-49
-------
however, has had a number of problems with low NO operation on its
A
boilers. The FGR fans on this boiler have caused vibration problems. It
was found that the fans were limited to maximum temperature changes of
56K/hr (100°F/hr). This necessitates slower unit startups from cold
conditions, and also limits load changes to 5 MW/min. This is in contrast
to the 75 MW/min attainable prior to the modification. There is also a
tendency for the unit to smoke under low NO operation which, in turn, has
A
limited the amount of overfire air to 50 percent of the design value. High
water wall tube temperatures were also encountered in this unit, requiring
some water wall inlet orifice changes.
Southern California Edison has modified six Combustion Engineering
tangentially fired units for low NO operation (Reference 6-10). The
A
units are rated at 320 to 335 MW, and have inverted furnaces so that the
gases flow downwards in the furnace. There are three levels of burners and
gas recirculation to the secondary air which were included in the original
design for steam temperature control purposes. At 20 percent gas
recirculation NO levels drop from a baseline value of 350 ppm at full
A
load to values ranging from 215 to 245 ppm. This amounts to a reduction in
NOV emissions of about 30 percent. Removing two burners from service from
A
opposite corners in the lower firing elevation in addition to FGR reduces
emissions by 42 percent from baseline values. No adverse effects were
reported with low NO operation (BOOS + FGR) of these boilers.
A
In summary, tangential oil-fired boilers can be modified to reduce
NOX emissions. For boilers with relatively low baseline emissions,
reduction of NO emissions to acceptable levels may be obtained simply by
A
low excess air operation. In many boilers this will require tuning and
adjustments to ensure uniform air distribution to the burners. LEA
operation has the advantage of increasing boiler efficiency and has,
therefore, been recommended as standard operating practice for most utility
boilers. Boilers which have higher baseline emissions will require flue gas
recirculation, off stoichiometric firing or a combination of the two to
reduce N0y emissions to desired levels. In some cases, retrofit
application of these modifications have led to problems such as vibrations,
high tube temperatures and impaired load pick up response.
6-50
-------
6.7 HORIZONTALLY OPPOSED OIL-FIRED BOILERS
A considerable body of data is available on horizontally opposed
oil-fired boilers retrofitted for FGR and OSC firing to control NO
A
emissions (e.g., References 6-7, 6-9 through 6-12). The reduction in NOX
emissions have, in some cases, been accompanied by a number of problems.
This may be due to the need to reduce NO emissions to very low levels
^
relatively quickly as required by local codes. Boilers designed with high
volumetric heat release furnaces tend to encounter problems with OSC
operation as the expanded combustion zone rapidly fills the small furnace.
Also addition of FGR to the windbox results in high burner throat velocities
which often result in flame instability and vibration problems.
Pacific Gas and Electric Company has reported on the modification of
six of its horizontally opposed units for NO reduction to meet local air
^
quality regulations (Reference 6-9). The Moss Landing Boiler No. 6, which
is a Babcock and Wilcox unit capable of generating 640 kg/s (5.1 x
106 Ib/hr) of steam, was modified to allow FGR to the windbox. The
existing FGR fans, used to control steam temperature by injecting flue gas
through the bottom hopper, were replaced with larger fans. New ducts,
dampers and an air foil mixing device were installed. The system was
designed to reduce the oxygen level in the windbox down to 17 percent. In
addition to the FGR system, the unit was also modified to operate with
BOOS. Out of a total of 48 burners, 8 on the top rows were operated on air
only. The remaining burners were enlarged to accommodate the increased fuel
flow. Up to 17 percent of the total combustion air could be injected
through the BOOS ports. In addition to the hardware modification for FGR
and OSC operation, the existing control and safety devices were modified to
control and monitor the new system. New flame scanners, windbox oxygen
analyzers and fully automated burner management systems were installed, and
the combustion controls were modified for minimum air and for automatic
proportioning of FGR to the windbox and hopper as a function of the load.
Table 6-13 gives a comparison of some process data on the Moss
Landing Boiler No. 6 operated at partial load under baseline, BOOS, FGR, and
BOOS + FGR modes (Reference 6-11). There seems to be very little difference
in the process variables from one operating mode to another except in the
NO emissions and the fan pressure and power requirements. The discharge
J\
pressure on the forced draft fan increased by about 15 percent when flue gas
6-51
-------
TABLE 6-13. COMPARISON OF MOSS LANDING UNIT NO. 6 UNDER BASELINE AND LOW NO
CONDITIONS UNDER PARTIAL LOAD (Reference 6-11) x
Process Variables
Load
Burner Firing Pattern
Gas Recirculatlon to Windbox
Overall Excess Q£
03 in Windbox
Fuel Oil Flow
SH Steam Flow
RH Steam Flow
SH Attemper ator Flow
RH Attemper ator Flow
SH Steam Pressure
RH Steam Pressure
SH Steam Temperature
RH Steam Temperature
Windbox Pressure
Furnace Pressure
Economizer Out Press
MM
Percent
Percent
Percent
kg/s (103 lb/hr)
kg/s (103 lb/hr)
kg/s (103 lb/hr)
kg/s (103 lb/hr)
kg/s (103 lb/hr)
MPa (psig)
MPa (psig)
K (°F)
K (°F)
kPa (inch H20)
kPa (inch H20)
kPa (inch H20)
Baseline
503.6
Normal
0.0
3.24
21.1
32.61 (258.3)
418.8 (3317)
354.3 (2806)
24.41 (217.1)
0
25.38 (3681)
2.74 (397)
816 (1009)
814 (1006)
3.48 (14.0)
3.11 (12.5)
0.57 (2.3)
BOOS
Operation
501.9
Upper row BOOS
0.0
4.48
21.0
32.34 (256.1)
422.1 (3343)
357.4 (2831)
27.17 (215.2)
0
25.22 (3658)
2.74 (398)
815 (1007)
809 (996)
3.36 (13.5)
2.59 (12.0)
0.67 (2.7)
FGR at
Windbox
501.0
Normal
19.5
4.20
18.4
32.52 (257.6)
421.8 (3341)
357.6 (2832)
27.17 (215.2)
0
25.38 (3681)
2.75 (399)
815 (1008)
811 (1000) '
3.73 (15.0)
3.11 (12.5)
0.59 (2.4)
» '
BOOS + FGR
Operation
500.3
Upper row BOOS
19.6
4.2
18.3
31.34 (248.2)
422.1 (3343)
357.7 (2833)
27.88 (220.8)
0
25.34 (3675)
2.74 (397)
816 (1010)
810 (999)
3.73 (15.0)
3.19 (12.8)
0.59 (2.4)
cr>
i
en
ro
-------
TABLE 6-13. Concluded
Process Variables
FO Fan No. 1 Discharge Press
FD Fan No. 2 Discharge Press
Flue Gas Recirculation Fans:
Current Consumption
Inlet Pressure
Discharge Pressure
Inlet Damper Position
A1r Foil Damper Position
Hopper Damper Position
NOX (3X 02 base)
S02
CO
kPa (Inch H20)
kPa (Inch H20)
Amps
kPa (Inch H20)
kPa (Inch H20)
Percent open
Percent open
Percent open
ppm
ppm
ppm
Baseline
4.35 (17.5)
4.11 (16.5)
224
-4.55 (-18.3)
3.23 (13.0)
30
0
100
273
2.5*
1000+
BOOS
Operation
4.35 (17.5)
4.11 (16.5)
220
-4.75 (-19.1)
2.91 (11.7)
30
0
100
221
--
1000+
FGR at
Wlndbox
4.98 (20.0)
4.73 (19.0)'
290
-3.66 (-14.7)
4.21 (16.9)
45
100
0
251
41.9*
1000+
• ' BOOS + FGR
Operation
4.98 (20.0)
4.73 (19.0)
290
-3.56 (-14.3)
4.28 (17.2)
42.5
100
0
169
46.4*
1000+
en
to
*Data appear lower than expected, but no explanation available.
-------
was recirculated through the windbox, while the power requirements of the
F6R fan increased by approximately 30 percent. Note that the carbon
monoxide levels were unacceptably high in all cases. Generally, the overall
excess air levels had to be increased to bring carbon monoxide emissions
down to reasonable levels (below 100 to 200 ppm). This increased excess air
can, in turn, affect superheater and reheater steam temperatures, and also
increase FD fan power requirements. This could lead to problems if the
steam attemperators or the fans are operating near their maximum capacity.
The Moss Landing boilers, and other horizontally opposed boilers
discussed in Reference 6-9, showed reduction in NOX emissions at full load
ranging from 33 to 50 percent from baseline when operated under OSC + FGR.
In some of those boilers, OFA ports were installed, designed to inject up to
20 percent of the combustion air above the burner zone. All boilers had FGR
systems installed capable of reducing excess oxygen in the windbox down to
17 percent. The combustion modifications resulted in a number of
operational problems. The most common problems were flue gas recirculation
duct and fan vibrations, furnace vibrations, and high furnace pressures.
Duct vibration problems usually necessitated installation of splitter vanes
and duct reinforcement. Fan vibration problems were resolved by reinforcing
the fan housing. Furnace vibrations and associated flame stability problems
were reduced by modifying the impeller air louvers to reduce the air
velocity at the lip of the impeller. Higher excess air requirements to
prevent smoking and excessive carbon monoxide generation were also
encountered in some units. No accurate data were available on the effect of
combustion modification on efficiency.
Southern California Edison Company (SCE) has also reported results of
low NO operation on two sets of its horizontally opposed oil-fired
A
boilers to satisfy emissions regulations (References 6-10 and 6-12). A set
of 480 MW, Babcock and Mil cox units with divided furnaces were retrofitted
with FGR to the windbox. The units have 32 burners each, divided into four
rows, and came factory-equipped with OFA ports. In general, OFA firing
reduced NOX emissions by 14 percent, from a baseline level of 330 ppm at
full load. In comparison, BOOS firing was capable of reducing NOX by
30 percent. The optimum BOOS pattern was obtained by operating the second
highest level of burners on air only. FGR alone decreased NO emissions
rt
by 9 percent. However, the combination of FGR + BOOS resulted in NOW
6-54
-------
reductions of 44 percent. FGR was also beneficial in that it reduced the
minimum oxygen level by 1/2 to 1 percent. It was found that combined OFA
and BOOS operation was not much more effective than BOOS alone and had the
disadvantage of increasing the required minimum excess air levels to prevent
smoke formation.
Another set of 750 MW, Foster Wheeler units was also retrofitted with
an FGR to windbox system by SCE. The units have 16 burners each, with four
levels of burners. OFA ports were included in the original design. These
units have baseline NO emissions at full load of 700 ppm. The high level
J\
of NO emissions are due to the small furnaces and high heat release rates
A
for these units. Operation with OFA ports reduced NO by 18 percent.
/\
With BOOS the reduction was 25 percent, again with the optimal BOOS pattern
obtained by operating the next to highest rows of burners on air only. BOOS
and OFA combined were not very effective and required an increase in overall
excess air levels. FGR alone, at 15 percent gas recirculation, decreased
NO emissions by 45 percent. A combination of BOOS and FGR resulted in a
/\
59 percent NOV reduction at 600 MW. Large reductions in NOV emissions
X A
were accompanied by a reduction in boiler capacity due to problems with
vibration and fan capacity. The maximum power generation, with NOX
emissions below the statutory limit of 225 ppm, was 680 MW. Some
experiments were performed with water injection on these units. Spraying
0.6 kg of water per kg of oil reduced emissions by 43 percent at 600 MW.
This is comparable to the reduction achieved by FGR. However, water
injection, in contrast to FGR, increased minimum oxygen requirements and
decreased boiler efficiency.
The problems encountered with low NO operation of the SCE units
A
involved flame detection, flame instability, boiler vibration, and limited
load capability. Flame detection problems arose due to changes in flame
characteristics with combustion modification, rendering some of the
conventional flame scanners inadequate. Addition of flicker (visible light)
scanners did not completely resolve the problem. Flame stability problems
were caused by the increased fuel flow in the active burners due to BOOS
operation and the increased burner throat velocities resulting from the
addition of FGR. Flame instabilities and pulsations also led to boiler
vibrations. Extensive testing and burner modifications were required to
resolve these problems. In the 750 MW boiler the modifications tested
6-55
-------
included adding diffusers to the oil guns, increasing burner throat
diameter, extending oil guns further into the furnace and changing burner
airflow distribution. Multiflame burner ("splitter") nuts, devices that
delay fuel/air mixing, were also installed as they had the effect of
increasing local fuel richness and reducing NO emissions by about 10
A
percent. Vibrations were finally reduced to an acceptable level by
modifications which provided up to 10 percent tertiary airflow around the
active oil guns. However, this resulted in an increase in minimum excess
oxygen requirements from a normal value of 3 percent to 5.5 to 6 percent.
The excess oxygen level was reduced to about 4 to 4.5 percent by subsequent
modification involving the swirl vanes in the active burners. The increased
excess air levels and additional head capacity requirements due to gas
recirculation have caused the forced draft fan to reach maximum capacity at
partial load. The maximum capacity of the boilers have been limited to 680
MW, which is much lower than 800 MW maximum rated capacity, and 750 MW
maximum continuous rating of the boilers.
Exxon has reported the results of a test on the 330 MW, B&W, Public
Service Electric and Gas Company (New Jersey), Sewaren Station Boiler No. 5
(Reference 6-7). The tests were limited to a maximum load of 285 MW as the
high pressure feed water heaters were out of service during the testing
period. At that load staged firing, taking 4 or 5 burners out of service
from a total of 24, reduced NO levels by 22 percent. Injecting flue gas
A
through the bottom of the furnace (not the windbox) reduced NO by only 7
A
percent. NO reductions of 16 to 24 ppm were obtained per 1 percent
A
reduction in excess oxygen level. Reducing load by 25 percent from 285 MW
decreased NO emissions by 19 percent. The boiler was also tested for
^
particulate emissions under baseline and low NO conditions. The
A
particulate loading under both firing conditions was the same 13 ng/J
(0.03 lb/10 Btu). Particulate size distribution also did not vary
significantly. Under baseline conditions 84.6 percent of the particles were
greater than 2.5^m and 8.2 percent were less than 0.5ym. Under low NO
,/ x
conditions the corresponding percentages are 80.0 and 10.8. No flame
stability or vibration problems were reported.
Foster Wheeler Energy Corporation has recently reported the results
of low N0¥ operation on a horizontally opposed oil fired unit (Appendix
/\
B). The unit was rated at 800 MW and was capable of producing 705 kg/s
6-56
-------
(5.6 x 106 Ib/hr) of main steam and 592 kg/s (4.7 x 106 Ib/hr) of reheat
steam. It was fitted with OFA Ports and was guaranteed at 500 ppm of NO
^
at 3 percent oxygen. During construction, however, the NO limits were
^
set at 250 ppm by the local authorities. A flue gas recirculation system
was therefore added. During start up of the unit it became clear that OSC
operation using OFA ports would not be capable of reducing NO emissions
A
to the desired levels even with the FGR system. OSC operation with burners
out of service was, therefore, tried and a program was initiated to
determine the optional BOOS pattern. The results of the study are shown in
Table 6-14 for various BOOS patterns. In all those tests, 8 out of a total
of 32 burners were out of service and 15 percent of the flue gas was
recirculated. The range of NO emissions shown corresponds to excess
A
oxygen levels ranging from the smoke threshold limit (minimum value) to 1
percent above the minimum. As seen from the table, the second row of
burners out of service gave the best results both for NO and minimum air
n
requirements. However, the increased air flow through the burners due to
addition of FGR caused severe flame instability and associated boiler
vibration problems which limited the load to 630 MW — approximately
80 percent of MCR.
As a consequence, an experimental burner modification program was
initiated. Various burner modifications were tried and rejected since they
did not improve flame stability. A burner modification which resulted in
stable flame characteristics was the inclusion of a tertiary air nozzle and
sleeve to provide a 10 percent air flow around the oil gun. The minimum
excess oxygen level, however, increased and the boiler reached its forced
draft fan capacity limit. Further modifications including reduction of the
tertiary air to 5 percent and installation of swirl vanes on the top row of
burners, have increased the boiler capacity to 680 MW (85 percent of MCR)
with NOX emissions at 225 ppm. Hence, operating within the constraints of
acceptable boiler vibration and NO compliance has resulted in a boiler
derate of 15 percent.
In summary, there is a potential for flame instability and furnace
vibration problems when horizontally opposed oil-fired boilers are modified
for low NOX operation on a retrofit basis. These problems can be quite
severe if NOX reductions of the order of about 50 percent or more are
required, and if the furnaces have high heat release rates. In such cases,
6-57
-------
TABLE 6-14. BURNER OUT OF SERVICE TEST PATTERNS FOR A
HORIZONTALLY OPPOSED OIL-FIRED BOILER
TEST
PATTERN
NO, PPM
(DRY AT
3% 02)
MIN Oz
TEST
PATTERN
NO, PPM
(DRY AT
3% 02)
MIN. 02
oooolocoo
195-225
3.1
oooqoooo
oooqoooo
oooqoooo
oto*p«o«
oocoooo
235-275
4.2
200-245
3.5
oooqoooo
•••*3Goo
oooqoooo
240-325
oooooooo
oooopooo
4.8
OOOOJOOOO
oooopooo
oooopooo
205-275
4.6
8
••••OOOO
oooooooo
245-300
OOOOfOOOO
4.5
••••oooo
220-265
4.0
oooopooo
ooooooo
oooopo
oo*«oot*
oooqoooo
255-300
4.7
OfOf'O«Ot
oooopooo
oooopooo
230-265
4.1
BURNER OUT-OF- SERVICE, AIR
REGISTER OPEN
6-58
-------
boiler derating by about 10 to 15 percent may occur. Usually the instability
and vibration problems require extensive testing and modifications to permit
acceptable operating conditions. Boiler operation may also become much more
complex especially during startup and periods of load fluctuation. Some
operations which are normally carried out by automatic control devices may
require manual control with low NO firing. Installation of new control
A
and safety equipment is often necessary. The particle loading and size
distribution are not significantly affected by low NO operation. More
^
data are needed on boiler efficiency. It is likely that some degradation in
performance will occur if the low NO operation results in increased
/\
excess oxygen requirements. On the other hand, flue gas recirculation to
the windbox has, in some cases, led to a reduction in minimum excess air
requirements. Also installation of an F6R system will improve performance
if the unit currently employs dampers and/or excess air to control steam
temperatures. No effect on superheater or reheater temperatures have been
noted, nor any on spray attemperators, due to low NOX firing. However,
such problems are site specific, and could possibly be encountered on other
oil-fired boilers.
6.8 SINGLE WALL OIL-FIRED BOILERS
Some process data are available on single wall oil-fired boilers
modified for N0¥ control. Specifically, three 100 MW, B&W, San Diego Gas
A
and Electric Company units (Encina Units No. 1, 2, and 3 (Reference 6-13))
were modified to achieve low NOX emissions. These modifications were
carried out in two steps: first to meet a December 31, 1971 local
regulation of 325 ppm N0¥, and second to meet a January 1, 1974 regulation
A
of 225 ppm NO . All three units exceeded these levels over much of their
A
operating range under normal baseline firing.
The 1971 regulations were met by BOOS operation with 2 burners out of
service out of a total of 10. A number of tests were carried out to
determine the best BOOS pattern, register settings, and overall excess 02
levels to achieve trouble free operation with low NO . The tests were
A
subject to the constraints of keeping CO concentrations below 100 ppm and
insuring no visible smoke plume formation (Ringleman No. less than 0.5). It
was found that burners No. 2 and No. 4 (see Figure 6-8) out of service with
6-59
-------
400
300 _
CO
GJ
Q.
Q.
200 _
c
o
•r~
I/)
>
•g 100
0)
X
o
Normal
operation
1971 regulation
Two-BOOS
operation
20
i
40
60
80
Load (MW)
100
120
Burner No. 12345
Register (% open) 70 70 70 70 70
O O O OO
O O OO O
Burner No. 67 8 9 10
Register (% open) 70 70 70 70 70
Normal operation
12345
100 100 70 100 100
O ® O ® O
O O O O O
6 7 8 9 10
100 70 70 70 100
BOOS operation,
fuel flow to No. 2 & No. 4
burners terminated
Figure 6-8. Comparison of NOx emissions with normal and two-BOOS
operation for Encina Unit No. T. (Reference 6-13).
6-60
-------
No. 2 and No. 4 registers full open resulted in the most satisfactory
combination when both NO reduction and operational suitability were
/v
considered. Tests also showed that the wing burners (Nos. 1, 5, 6, and 10)
received less air than the remaining burners so that they tended to smoke
load, under fuel-rich operation. This problem was resolved by opening the
wing registers 100 percent while keeping the registers of the remaining
burners in service throttled to 70 percent open to force more air to the
wings. The burner and register patterns before and after modification are
shown in Figure 6-8.
With the burner and register patterns fixed, excess Op levels were
varied to determine the minimum levels which would provide low NOX
operation without excessive CO or smoke emissions, and would not lead to
problems such as flame instability, etc. Figure 6-9 shows a comparison of
the recommended excess 0? levels as a function of load under normal
operation and the recommended values for operation with combustion
modification for the Encina Unit No. 1. Figure 6-8 shows the NOX
emissions associated with these excess 02 levels and modes of firing.
Table 6-15 provides a comparison of some process data under baseline and low
NO operation with two burners out of service.
J\
Operation with lower excess air and combustion modification in the
furnace sometimes led to a decrease in superheater temperatures in the
Encina units at full load. Due to the restriction on excess airflow, the
operator was obliged to rely on flue gas recirculation to increase
superheater temperatures.* As shown in Table 6-15, the Encina Unit No. 1
had difficulty reaching a normal superheater temperature of 811K (1000°F),
although the flue gas recirculation had been increased as evidenced by the
increase in RC fan amperage. In another unit, Encina No. 3, the problem
with superheater temperature occurred only at peak
*In these units, recirculated flue gas is introduced between water tubes
on the back wall of the furnace and not with the combustion air. Effect
of the flue gas on NOX emissions should therefore be small.
6-61
-------
6 _
5 -
*
S_
C
-------
TABLE 6-15. COMPARISON OF ENCINA UNIT NO. 1 OPERATED UNDER BASELINE
CONDITIONS WITH TWO BURNERS OUT OF SERVICE (Reference 6-13).
Process Variables
Load
Control Room 02
Burners Out of Service
Steam Temperature
Indicated Oil Flow
Indicated Airflow
Oil Pressure
Burner Supply
Burner Return
Oil Temperature
Furnace Draft Pressure
AH Gas Out Temp.
AH Gas In Temp.
RC Fan
FD Fan
ID Fan
Measured NOX
Measured CO
MW
Percent
K (op)
(Meter settings,
arbitrary units)
MPa (psi)
K (OF)
kPa (inch H20)
K (OF) N
S
K (OF) N
S
Amps
Amps N
S
Amps N
S
ppm N
S
ppm N
S
Baseline
Operation
98
3.0 to 3.2
None
811 (1000)
960/1100
59/70
5.0 (725)
3.0 (440)
374 (214)
-0.1 (-0.5)
455 (360)
444 (340)
669 (745)
666 (740)
39
78
80
118
120
335
340
30
35
Two- BOOS
Operation
98
2.5
Nos. 2 & 4
800 (980)
960/1100
59/70
5.1 (740)
3.1 (450)
374 (214)
-0.1 (-0.5)
455 (360)
444 (340)
678 (760)
669 (745)
50
74
75
120
122
200
235
30
30
6-63
-------
i.e., loads above 110 MW (the boilers are rated at an MCR of 110 MW). The
superheater temperature at 114 MW fell from a normal of 811K(1000°F) to
800K (980°F), and the reheater temperature fell from a normal of 800K
(980°F) to 788K (958°F).
Due to the increased oil flow to the burners in service, the oil tips
and return passages had to be enlarged. The flames at each burner were
observed to be satisfactory with no flame instability or blow off noted.
Off stoichiometric firing also caused longer flames. Flame filled the
furnace at the burner levels, and some intermittent flame carryover to the
superheater inlet occurred. This did not result in short-term problems such
as high tube temperatures, though. Off stoichiometric firing also resulted
in the flame zone becoming very hazy and obscure. No other problems were
encountered with two-BOOS operation. The boilers were operated for over
2 years in this mode with no signs of abnormal tube deposits nor chemical
attack or erosion.
To meet the 1974 regulations (225 ppm NO ), the boilers were
A
operated with three burners out of service. The optimal BOOS pattern was
obtained by terminating fuel flow to Nos. 2, 4, and 8 burners while the air
registers were left at 100 percent open to act as air injection ports. The
air registers on the remaining seven burners were set at 55 percent open.
The oil burner tips were enlarged again to accommodate the increased flow in
the active burners. The oil tip diameter, the tangential slot width, and
the return passage diameter were all widened to provide adequate flow and
desired flame structure. The operating excess 02 level had to be
increased generally above the levels recommended for two-BOOS or normal
operation to curtail smoke formation. This did not lead to any measurable
degradation in boiler performance. However as the boilers are mainly ID fan
limited, an increase in excess air levels resulted in a peak load
curtailment up to 5 MW in some cases.
Figure 6-10 shows the burner and register patterns used to achieve
the 1974 standards in the Encina units when firing oil. It also shows NO
J\
emission from the Encina Unit No. 1 when operating with the excess 02
levels shown in Figure 6-11. The range of excess 02 levels shown in
Figure 6-11 are the minimum required to operate the boiler without smoking
while permitting the operator a certain range of flexibility. A comparison
with Figure 6-9 indicates that the recommended levels of 0« with
6-64
-------
240
CM
O
CO
o
•(->
-o
u
O)
S-
o
o
>>
-a
200 -
180 _
160
Maximum excess 0,
Recommended excess 0,
Minimum excess 09
140 J
120 -
100 -
80 -
60 -
40 -
20 -
^s<
Burner No.
Register (% open)
Burner No.
Register (% open)
OO Indicates air only
^-^^
i i I I
' TV
•Q
1
55
O
o
6
55
i
20 30 40 50 60 70
m
2
100
\^*\
Vcj'
O
7
55
1
80
^
3
55
O
fxi
8
100
i
90 1
r
4
100
(j(\
O
9
55
I
00 1
5
55
O
o
10
55
I
10 12
Load (MW)
Figure 6-10. NOX emissions for oil fuel with seven-burner operation
for Encina Unit No. 1 (Reference 6-13).
6-65
-------
8.0 -
7.0 -i
6.0
i.
O
Burner No.
Register
open)
oo® oo
678 9 10
55 55 100 55 55
Maximum airflow
Minimum excess 0
Recommended
excess 00
Indicates air only
30
40 50 60
I
70
I
80
! I I I
90 100 110 120
Load (MW)
*As read in control room
Figure 6-11. Operating excess 02 curve for Encina Unit No. 1 for
oil fuel with seven-burner operation (Reference 6-13)
6-66
-------
three-BOOS operation are much higher than the levels of 02 with two-BOOS
or normal operation, especially at higher load levels. This is due to the
increased tendency for the boilers to smoke as the degree of off
stoichiometric firing is increased. The Encina Unit No. 1 is atypical,
however, 1n that all registers have a common swirl direction in the lower
row of burners. This causes the flame to turn upwards and cling to the
furnace wall instead of protruding into the furnace. This flame pattern
caused local smoking which cleared only when excess air was increased. The
other two Encina units have alternating swirl directions which apparently
cause the flames to protrude into the furnace. Those two units can
therefore operate with lower excess air levels (about 2 to 4 percent at
100 MW) than Encina No. 1. Nevertheless, they still require higher excess
Op levels than with normal or two-BOOS operation.
Peak load tests for Encina No. 1 showed that with an excess 02
level of 4.5 percent the maximum turbine load was 100 MW. Lowering the 02
level to 3.9 percent allowed the maximum turbine load to increase to
103 MW. Thus a 1 percent increase in Op level decreased the unit's
capacity by 5 MW. As mentioned earlier, this is mainly due to the boiler
being airflow limited at full load. An increase in excess air levels,
therefore, translates to a smaller fuel flow for the same airflow rate, thus
reducing boiler capacity. In general, it was found that for all the three
boilers, a derate of up to 5 MW could be expected with three-BOOS operation.
Table 6-16 shows a comparison of some process variables when Unit No.
1 was operated with two and three burners out of service. In contrast to
the data for two-BOOS operation shown in Table 6-15, the data in Table 6-16
for two-BOOS operation show that the superheater temperature reached the
normal level of 811K (1000°F). This was due to the increased excess 02
level for the two-BOOS operation in Table 6-16. It was found that for
normal and two-BOOS operation increasing excess air levels always increased
steam temperatures. Surprisingly, however, with three-BOOS operation
increasing excess air levels decreased steam temperatures. In Table 6-16
the SH steam temperature for three-BOOS operation is 800K (980°F) at 3.8
to 4.0 percent 02 (below the recommended 02 range). Increasing the
excess 02 level will further decrease the SH temperature. Lower steam
temperatures generally result in lower cycle efficiency. However, a
comparison of cycle efficiency has not been attempted here due to possible
6-67
-------
TABLE 6-16. COMPARISON OF ENCINA UNIT NO. 1 OPERATED WITH TWO AND
THREE BURNERS OUT OF SERVICE (Reference 6-13)
Process Variables
Load
Control Room Q£
Burners Out of Service
SH Steam Temp.
RH Steam Temp.
Steam Flow
Oil Flow
Indicated Airflow
Attemperator Temp. In
Attemper ator Temp. Out
Furnace Draft
AH Gas In
AH Gas Out
Ringleman Smoke
Chart No.
Measured NOX
Measured CO
MW
Percent
K (0F)
K (OF)
kg/s (103 Ib/hr)
kg/s (103 Ib/hr)
(Meter Setting)
K (OF) N
S
K (OF) N
S
kPa (inch H20)
K (OF) N
S
K (OF) N
S
ppm N
S
ppm N
S
Two- BOOS
Operation
99.8
3.3 to 3.5
Nos. 2 & 4
811 (1000)
808 (995)
81.9 (650)
7.5 (59.5)
54
689 (780)
675 (755)
680 (765)
675 (755)
-0.11 (-0.45)
633 (680)
633 (680)
433 (320)
425 (305)
0.41
206
174
0
0
Three-BOOS
Operation
99.8
3.8 to 4.0
Nos. 2, 4,
& 8
800 (980)
800 (980)
81.0 (643)
7.3 (58)
52
678 (760)
672 (750)
678 (760)
678 (760)
-0.11 (-0.45)
622 (660)
630 (675)
433 (320)
422 (300)
0.45
145
115
0
0
6-68
-------
variations in the fuel oils used in the tests listed in Table 6-16.
In all other respects there is no significant difference in the process
variables between two-BOOS and three-BOOS operation.
It should be noted that the various test conditions for Encina Unit
No. 1, as discussed in this section, do not necessarily reflect current
operating practice. The utility has been active in maintaining low NOX
emissions and improving boiler efficiency with different firing
configurations and burners (Reference 6-20).
In summary, front wall oil-fired boilers, at least of the type
studied here, show no significant deterioration with BOOS operation. With
increased off stoichiometric firing some derating may occur, but the extent
of derating is generally small. Higher excess 02 levels will probably be
associated with increased staging and some loss in efficiency may be
expected. From a boiler operator's point of view, the major change will be
in flame patterns and increased tendency to smoke. Careful inspection of
the furnace and convective tubes would be recommended at periodic intervals,
but again no major problems would be anticipated.
6.9 TURBO FURNACE OIL-FIRED BOILERS
A series of tests for NO reduction were carried out on South Bay
rt
Boiler No. 3 operated by San Diego Gas and Electric Company (Reference 6-8).
The boiler is a 12 burner turbo fired Riley Stoker unit with a maximum
continuous steam flow rating of 145 kg/s (11.5 x 105 Ib/hr). Before any
combustion modifications were undertaken, the flow of air between the two
windboxes was balanced. Pitot tubes were installed in the individual burner
compartments. The splitter vane which divides the airflow to the two
windboxes was then adjusted to ensure even air distribution as measured by
the pitot readings.
The velocity dampers and the directional vanes at the individual
burners were next optimized for low NO operation. It was found that
/\
velocity dampers had little effect on NO emissions. Still, the optimum
n
position was found to be the normal position of from 60 to 80 percent open
for both top and bottom dampers. NO levels were however, sensitive to
^
directional vanes positions. By raising both the upper and lower vanes up
by 30° with respect to the direction of the fuel guns, NO reductions by
A*
40 to 50 ppm were achieved. These NO reductions were apparently due to
n
6-69
-------
an overfire air effect caused by directing the airflow 30 degrees above the
direction of fuel flow.
The first two columns of Table 6-17 show a comparison of some process
data for the boiler operated at partial load under normal and modified
airflow conditions. Except for the reduction in NO , there is very little
A
change in the process variables.
Although the modified airflow conditions reduced NO emissions, the
A
reductions were not sufficient to meet statutory requirements especially at
higher loads. Water injection was then tried as a NO reduction
A
technique. Water was introduced into the heated combustion air as a fine
mist by means of a bank of spray nozzles. Reductions in NO emissions up
A
to 50 percent of baseline at maximum load were obtained. Water injection
tests with water to fuel loadings of up to 1.016 kg HLO/kg oil were
carried out without any flame instability problems encountered. Steam and
tube temperatures were only slightly affected. However, oil consumption
increased by as much as 6 percent at full loads. The last two columns in
Table 6-17 give some process data for the unit operated at partial load
under different water injection rates. NO emissions decreased
J\
substantially with water injection. Note that the unit load decreased with
water injection for approximately constant fuel flowrate. Figure 6-12 shows
the variation of NO emissions over the boiler load range under baseline,
rt
modified airflow, and water injection conditions. The water injection was
increased with load as shown in the figure to maintain NO levels within
A
the regulation limits under all load conditions without excessive
performance losses. Due to the increased fuel consumption associated with
water injection, this control technique was considered as an interim measure
by SDG&E until OFA ports could be installed and tested.
Another NO control technique tested on this unit was Reduced Air
A
Preheat (RAP). Combustion air temperature was lowered by bypassing the
preheater. NO reductions of 40 to 70 ppm per 56K (100°F) reduction in
A
air temperature at 75 and 100 percent of full load, respectively, were
obtained. It was found, however, that to achieve the same NO reduction,
A
water injection was more cost-effective than RAP due to lower boiler
efficiency losses. RAP \
technique for this unit.
6-70
efficiency losses. RAP was, therefore, not recommended as a NO control
A
-------
TABLE 6-17. COMPARISON OF SOUTH BAY UNIT NO. 3 AT PARTIAL UNDER BASELINE
AND LOW NOX OPERATION ON OIL FUEL UNDER PARTIAL LOAD
(Reference 6-8)
Process Variables
Load
!xcess Oxygen
Steam Flow
Fuel Oil Flow
Water Injection:
Flowrate
Water/Fuel Ratio
Velocity Dampers
Top
Bottom
Directional Vanes
Upper
Lower
Burner Air Dampers
Pressures:
Steam Drum
Burner Supply
Burner Return
Uindbox
Furnace
Temperatures:
SH Steam
RH Steam
Oil Supply
AH Air In
AH Air Out
AH Gas In
AH Gas Out
F.D. Fan Current
Emissions:
NOX (at 3X 02)
CO
Ringleman Smoke Density
HU
X
kg/s (105lb/hr)
kg/s (103lb/hr)
kg/s (103lb/hr)
kg/kg
X open
X open
degree
degree
X open
MPa (psi)
MPa (psi)
MPa (psi)
kPa (In H20)
kPa (In H20)
K (°F)
K (°F)
K (°F)
K (°F)
K ( F)
K (°F)
K ( F)
Amps
ppm
ppm
ppm
Baseline
138
5.4
119 (9.5)
7.9 (63)
0
0
75
75
0
0
100
14.4 (2090)
4.86 (705)
2.59 (375)
2.3 (9.2)
1.5 (6.1)
811 (1001)
811 (1001)
372 (210)
297 (76)
580 (585)
655 (720)
408 (275)
156
267
0
0
Airflow
Adjustment
138.5
4.8
121 (9.6)
8.0 (63.5)
0
0
75
75
up 30
up 30
100
14.4 (2090)
4.90 (7.10)
2.59 (375)
2.2 (9.0)
1.4 (5.5)
810 (999)
802 (985)
372 (210)
298 (77)
578 (582)
655 (720)
407 (273)
160
229
0
0
Water Injection
133
4.5
111 (8.8)
8.1 (64)
3.72 (29.5)
0.461
77
77
up 30
up 30
100
14.2 (2055)
4.76 (690)
3.31 (480)
2.3 (9.4)
1.9 (7.5)
806 (992)
796 (974)
368 (204)
298 (77)
422 (300)
646 (703)
399 (260)
165
162
0
0
131.5
4.8
110 (8.7)
8.1 (64)
6.05 (48.0)
0.750
77
77
up 30
up 30
100
14.1 (2050)
4.76 (690)
3.34 (485)
2.4 (9.6)
1.9 (7.8)
806 (992)
809 (997)
368 (203)
297 (76)
<422 (<300)
652 (715)
402 (265)
164
143
0
0
6-71
-------
500
I
~-J
rv>
400-
300.
o
s«
a.
a.
I 200-
in
•t-
V
100.,
Mode
Normal
Modified
Directional
Vanes
Horizontal
30° up
Velocity
Dampers
60 - 80% open
60 - 80% open
Normal
Modified
vane/damper
positions
Water injection flowrate
30
To"
50
60
10 Ib/hr
70
80
Modified
vane/damper
positions
and water
injection •
- - 100
80
60
40 £
m
o
20
0
90
12
10
8
6
4
2
0
o>
-------
Pacific Gas and Electric Company has also modified its Riley Stoker
Potrero Boiler 3-1 for NO reduction (Reference 6-9). This boiler is
c
capable of generating 189 kg/s (1.5 x 10 Ib/hr) of steam. The hardware
modification on the boiler included installation of OFA ports designed to
handle up to 25 percent of combustion air. An F6R system was also
retrofitted to the unit. Windbox oxygen content could be diluted to 17
percent with FGR. In addition some reheater surface was removed and a new
and larger fin tube economizer was installed as part of the modifications.
Baseline NOX emissions for the Potrero Unit 3-1 were approximately
350 ppm at full load. Low NO operation (OFA + FGR) of the boiler
A
reduced the emissions down to about 250 ppm. Some tests were carried out
with OFA alone, but it was found that stack smoking occurred with the OFA
ports opened only a small amount. The tendency to smoke required that the
boiler be operated with a minimum of 4 percent excess 0« under normal
low NOY operation (OFA + FGR). The high excess air requirements
n
combined with increased convective transfer due to FGR and altered flue
gas temperature profiles due to OFA caused tube metal and steam
temperature limits to be approached. During a period of time the boiler
experienced one superheater tube failure per month. The boiler was
curtailed to 95 percent of full load and removed from automatic dispatch
operation due to unacceptable temperature excursions at high loads. Due
to the addition of economizer surface to this boiler, the boiler
efficiency was expected to improve despite the higher excess air
requirements.
In summary, traditional NO control techniques such as FGR and
^
OSC have been successful in reducing NO emissions from oil-fired turbo
^
furnace boilers. Moderate amounts of NO reduction can be obtained by
^
experimenting with velocity dampers and directional vanes to achieve an
overfire air effect. In addition, substantial reduction in NOX
emissions can possibly be obtained by the use of OFA ports above the
throat to create a precombustion fuel-rich zone in the lower portion of
the furnace. However, this may be accompanied by an increased tendency
towards smoking. Higher oxygen levels required to eliminate smoke may in
turn lead to higher superheater and reheater tube and steam temperatures.
These temperatures are usually increased by FGR and OFA operation, so that
6-73
-------
an increase in excess air requirements may cause an exacerbation of the
problem. Finally, water injection can be used to control NO emissions
/\
from oil-fired boilers. There is, however, a penalty associated with this
type of low NO operation in reduced boiler efficiency and consequently
A
higher fuel costs per unit of electrical energy generated. Water injection
may be useful as a temporary measure to control NO emissions until major
A
hardware modifications such as FGR and OFA can be retrofitted.
6.10 TANGENTIAL GAS-FIRED BOILERS
Process data on tangential gas-fired boilers closely resemble that on
tangential oil-fired boilers as most such units are designed to accept both
oil and gas fuels depending upon availability. Many tangential units have
low baseline NO emissions due to the nature of combustion in tangential
A
furnaces. For these units simple modifications such as low excess air
operation are often sufficient to reduce NO emissions to meet statutory
A
requirements. In other cases, where baseline NO emissions are much
A
higher than the desired levels, the usual NO reduction techniques used
A
with gas and oil such as flue gas recirculation (FGR) and off stoichiometric
combustion (OSC) have been employed.
A comparison of process data on South Bay Boiler No. 4 under baseline
and low NO operation is shown in Table 6-18 (Reference 6-8). The unit is
A
a 230 MW Combustion Engineering tangentially fired cycling boiler, with a
straight-through furnace, capable of generating 198 kg/s (1.57 x 10
Ib/hr) of steam. The boiler is operated by San Diego Gas and Electric
Company. The unit had always required higher operating excess oxygen levels
than the design values due to a tendency for high carbon monoxide
generation. Stack traverse data showed large local carbon monoxide
concentrations in one portion of the stack. Due to the straight-through
design of the boiler, there is little mixing of the gases from the furnace
to the stack so that high local CO levels in one portion of the stack
reflect high CO generation in a corresponding section of the furnace. It
was found that the airflow to the burners was maldistributed. Uniform
distribution was achieved by closing the auxiliary air dampers, but instead
of closing the dampers fully, they were left open by 10 percent to cool and
purge the auxiliary air compartment. The fuel air dampers were left fully
open. Better distribution of air resulted in a lowering of minimum excess
air levels, which consequently led to a decrease in NO emissions. The
A
6-74
-------
TABLE 6-18.
COMPARISON OF SOUTH BAY UNIT NO. 4
OPERATED UNDER BASELINE AND LOW NO
CONDITIONS UNDER PARTIAL LOAD
(Reference 6-8)
Process Variables
Load
Excess Oxygen
Burners Out of Service
Burner Tilt
Flowrates:
Steam
Natural Gas
Temperatures:
SH Steam
RH Steam
AH Air Out
Stack Gas
Pressures:
Steam Drum
Natural Gas at Burner
Furnace
Windbox/Furnace Differential
FD Fans:
Discharge Pressure
Current
Fuel Air Damper
Auxiliary Air Damper
Emissions:
N0x (at 3X 02)
CO
Ringleman Smoke Density
MW
Percent
Degrees
kg/s (106 Ib/hr)
nm3/hr (106 scfh)
K (°F)
K (°F)
K (°F)
K (°F)
MPa (psi)
MPa (ps1)
kPa (inch H20)
kPa (inch H20)
kPa (inch H?0)
Amps
Percent open
Percent closed
ppm
ppm
asellne
176
3.8
None
-14
1«5 (1.15)
4.8 (169)
775 (935)
769 (925)
384 (231)
636 (685)
12.3 (1790)
0.110 (16.0)
1.0 (4.0)
1.0 (4.0)
3.31 (13.3)
302
100
0
119
7
0
Low Excess
Air
182.5
1.3
None
-18
153 (1.21)
5.0 (175)
784 (951)
785 (953)
380 (225)
628 (670)
11.2 (1630)
0.116 (16.8)
0.85 (3.4)
2.2 (8.8)
4.06 (16.3)
302
100
90
97
145
0
OSC Operation
178.5
3.3
2 from top tier
-18
151 (1.20)
5.0 (175)
783 (950)
784 (952)
383 (229)
630 (675)
11.1 (1610)
0.155 (22.5)
1.0 (4.0)
>2.5 (>10)
4.70 (18.9)
320
100
90
106
4
0
6-75
-------
lowered throughput in the furnace also reduced stack losses as indicated by
lower stack gas temperatures and lower fuel consumption. These effects are
shown for the range of boiler loads in Figures 6-13 and 6-14. The first two
columns in Table 6-18 also give some process data at partial loads for the
boiler operated under normal conditions (with nonuniform air distribution)
and low excess air conditions (with uniform air distribution).
The boiler was also tested with some burners on air only. The
furnace has three levels of burners. The results of taking two burners out
of service from opposite corners of the topmost tier are shown in the last
column of Table 6-18. It is seen that although the dampers are positioned
for uniform air distribution, the excess air level was higher than the
minimum value obtained with low excess air operation. The NO level was
rt
also higher than that obtained with LEA operation. However, some tests at
reduced load (around 140 MW) showed that N0¥ emissions could be reduced
A
down to about 70 ppm with all burners in the top tier on air only. The
minimum excess oxygen level under these conditions was approximately 3.5
percent; any further reduction caused excessive carbon monoxide emissions.
OSC operation was not recommended for this boiler as LEA firing was capable
of reducing NO emissions to values below the regulatory requirements, and
^
because OSC operation resulted in higher required excess oxygen levels, with
associated loss in efficiency.
Pacific Gas and Electric Company has modified its Combustion
Engineering, tangentially fired, 675 kg/s (5.36 x 10 Ib/hr) of steam,
Pittsburg No. 7 Boiler for low NO operation (Reference 6-9). The
/\
modifications involved installing OFA ports, capable of injecting 20 percent
of total air, and introducing F6R to the windbox capable of reducing oxygen
in the combustion air down to 17 percent. PG&E encountered a number of
problems with low NO operation of this boiler. The baseline NO
n /\
emissions of this unit at full load amounted to approximately 750 ppm which
is relatively high for a boiler of this type. The amount of flue gas
recirculation required to reduce the NO emissions to the local limit of
A
175 ppm caused excessive reheat steam temperatures and subsequent load
curtailment. These high rates of FGR combined with OSC operation often led
to high convective section tube and steam temperatures. FGR increased mass
flowrates, the increased velocities giving rise to higher heat transfer
coefficients.
6-76
-------
I
-J
to
CO
OJ
u
X
120
i/i (Si
* £ 110
X •—•
o
0)
100 -
90 -
4
6.0 -
5.0 -
4.0 -
3.0 -
9 0 _
3 0. £ .U —1
^^^
Normal operation (baseline)
Low excess air operation
100
120
T
140
160
Load (MW)
i
180
200
220
240
Fiaure 6-13. Comparison of NOX emissions and minimum excess oxygen
levels under baseline and low excess air conditions
for South Bay Unit No. 4 (Reference 6-8).
-------
0)
10
|
£1
3
<4-
«/>
ta
03
•U
Q.
-* E
(7) U O)
1 ^ '*"'
00 ^
60.
^ 50-
"E 40-
30-
^^
^
700.
650.
600-
550.
i
ro
o
o
in
2400
2000 -
1600 -
1200 _
800 -
£
800 -
700 -
600 -
500 -
2.2% reduction
3.1% reduction
80
Normal operating
(baseline)
Low excess air operation
100
I
120
140
i
160
180
200
220
240
Load (MW)
Figure 6-14. Comparison of gas consumption and stack temperature under baseline
and low excess air conditions for South Bay Unit No. 4 (Reference 6-8).
-------
Staging resulted in a lengthened combustion zone which increased the
furnace outlet gas temperatures. This phenomenon is common in gas-fired
units with small furnaces where the combustion zone fills the entire
furnace. In some cases the heat transfer rates and temperatures may be high
enough to cause tube failures or excessive steam temperatures which exceed
steam desuperheater capacities. In such cases boiler derating may be
required. In the case of the Pittsburg No. 7 Boiler, maximum load capacity
was reduced by 25 percent. The load curtailment on Pittsburg No. 7 has
recently been overcome by removing the capability of circulating flue gas
through the furnace hopper. In other cases, where a significant amount of
reheat attemperation is required, a loss in cycle efficiency will result.
Some other adverse effects caused by low NOV operation at the
A
Pittsburg No. 7 Boiler were: fan, duct, and building vibrations, high water
wall panel outlet tube temperatures, and reduction in load change response.
In summary, tangential gas-fired boilers can be modified for NOX
reduction using traditional NO control techniques. In cases where the
A
baseline emissions are not much higher than the desired levels, simple
techniques such as low excess air operation may be used. Excess air levels
can be minimized by ensuring uniform fuel/air distribution at the burners.
In cases where baseline emissions are high, FGR, OSC, or a combination of
the two may be required. In some cases, adverse effects such as vibrations,
high tube and steam temperatures, and reduced load response capability will
be encountered. In certain cases, especially with small furnaces and high
volumetric heat release rates, boiler derating may occur. For new boilers
with larger furnaces and factory- equipped OFA systems, there should be no
adverse effects associated with low NO operation.
6.11 HORIZONTALLY OPPOSED GAS-FIRED BOILERS
Very little process data are available on horizontally opposed
gas-fired boilers. However, boilers are often designed to accept both gas
and oil fuels. Thus, much of Section 6.7 on horizontally opposed oil-fired
boilers would also be pertinent here, especially the details concerning
hardware and control modifications which are essentially similar for oil-
and gas-fired boilers.
Pacific Gas and Electric Company has reported its experience with
converting six of its horizontally opposed boilers to low NO operation
A
(References 6-9 and 6-14). The Moss Landing Boilers Nos. 6 and 7 were the
6-79
-------
first among P6&E boilers to be modified for reduction of NO emissions.
A
These boilers were manufactured by Babcock and Mil cox Company, have 48
burners each divided into four levels, and can produce up to 640 kg/s
(5.1 x 10 Ib/hr) of steam. The baseline NO emissions on these boilers
/\
averaged over 1400 ppm at full load. Various techniques were tried to
reduce NO levels. The only techniques which resulted in the substantial
reductions desired were off stoichiometric firing, by taking the top level
of burners out of service and a combination of flue gas recirculation and
off stoichiometric firing. OSC firing alone gave NO reductions of
A
81 percent. OSC combined with FGR resulted in a reduction in N0x of
94 percent from baseline.
Table 6-19 gives a comparison of process data for the Moss Landing
Boiler No. 7 when operated under OSC and a combination of OSC with various
degrees of FGR to windbox (Reference 6-11). Unfortunately, no corresponding
baseline data were available with matching operating conditions. It is seen
that the power requirements of the FGR fan increase substantially as the
amount of FGR to the windbox is increased. At about 7 percent FGR to the
windbox the fan power increased by approximately 10 percent over that
required for FGR to the hopper. When FGR to the windbox was increased to
19 percent, the fan power requirements increased by 66 percent. The furnace
pressure also increased as FGR to the windbox increased due to the higher
furnace mass flowrate. The original furnace trip, which was set at 5.2 kPa
(21 inch H20), had to be raised to 6.0 kPa (24 inch H20) under low N0x
operation. This is very close to the boiler maximum design pressure of
6.7 kPa (27 inch H20).
At high rates of FGR to windbox, attemperation of reheat system was
required. This was partly due to the increased mass flowrates which tended
to increase heat transfer coefficients in the convective section. Under
normal operating procedure, reheat steam spray attemperation is generally
avoided due to associated cycle efficiency losses. Typically under baseline
operation of these boilers, superheat and reheat steam temperatures are
controlled by a combination of flue gas recirculation to the hopper,
proportioning dampers and spray attemperation. With FGR directed to the
windbox for NO control, it could no longer be employed to control steam
A
temperatures. Some limitations on damper control were also encountered due
to the high furnace pressures.
6-80
-------
TABLE 6-19.
CTt
1
00
COMPARISON OF MOSS LANDING BOILER NO. 7 UNDER OFF STOICHIOMETRIC COMBUSTION
AND COMBINED OFF STOICHIOMETRIC COMBUSTION AND FLUE GAS RECIRCULATION
(Reference 6-11)
Process Variables
Load
Burhner Firing Pattern
Gas Recirculatlon to Ulndbox
02 In Wlndbox
Overall Excess 02
Mean Steam Flow
SH Attemp. Sprsy Flow
RH Attemp. Spray Flow
SH Steam Pressure
RH Steam Pressure
SH Steam Temperature
RH Steam Temperature
Furnace Pressure
Air Heater Temperatures:
A1r In
Air Out
Gas In
Gas Out
Flue Gas Recirculatlon:
Fan Current Consumption
GR to Air Foil
GR to Hopper
NOX (3* 02 base)
CO
m
Percent
Percent
Percent
kg/s (106 Ib/hr)
kg/s (H>3 Ib/hr)
kg/s (103 Ib/hr)
HPa (pslg)
MPa (pslg)
K (°F)
K (OF)
kPa (Inch H20)
K (0F)
K (°F)
K (0F)
K (OF)
Amps
Percent
Percent
ppm
ppm
BOOS
Operation
733
Upper row BOOS
0.0
21.0
1.6
6S5 (5.20)
19 (150)
0
25.7 (3720)
4.4 (630)
806 (992)
810 (999)
5.15 (20.7)
299 (79)
564 (555)
625 (666)
400 (260)
106
0
100
223
178
BOOS + 7 Percent
FGR to Ulndbox
734
Upper row BOOS
6.9
19.8
1.8
649 (5.15)
24.4 (194)
0
25.7 (3720)
4.4 (630)
808 (994)
811 (1000)
5.23 (21.0)
300 (80)
565 (558)
628 (671)
401 (262)
116
100
0
148
196
BOOS + 14 Percent
FGR to Ulndbox
734
Upper row BOOS
13.8
18.8
1.7
649 (5.15)
21.8 (173)
0
25.7 (3720)
4.4 (630)
808 (995)
810 (999)
5.85 (23.5)
300 (80)
574 (574)
635 (684)
403 (266)
138
100
0
103
55
BOOS + 19 Percent
' 'FGR to Ulndbox
733
Upper row BOOS
19.0
18.1
1.9
649 (5.15)
16.0 (127)
9.7 (77)
25.7 (3720)
4.4 (630)
805 (990)
811 (1000)
5.77 (23.2)
299 (79)
574 (573)
633 (680)
402 (266)
176
100
0
73
28
-------
From Table 6-19 it is seen that when FGR is increased to 19 percent,
about 10 kg/s (80,000 Ib/hr) of reheat spray flow was required. PG&E has
estimated that this results in a 0.8 percent loss in cycle efficiency.
Removing some reheater surface would overcome this problem, but was not
attempted in this case because it would have resulted in even higher
efficiency losses when the boilers switched to oil fuel.
It should be noted that the data in Table 6-19 were taken under clean
boiler conditions. With gas fuels NO emissions are very sensitive to
A
boiler wall conditions. This poses a significant problem in boilers which
alternate between oil and gas fuels. In the Moss Landing Boiler the NO
emissions standards of 125 ppm could not be met when switching back to gas
fuel after a few days of oil burning, even though the ash content of the
fuel oils used was only about 0.02 percent by weight. Only a complete water
washing of the furnace and convective passes after each period of oil
burning could resolve the problem, but this solution was considered
impractical when frequent switching was required. Switching from oil to gas
also caused problems of high reheat and superheat temperatures as the higher
furnace exit gas temperatures associated with low NO operation were
A
further exacerbated by decreased heat absorption in a dirty furnace. In
some of the horizontally opposed PG&E boilers the reheat spray water limit
was approached on gas fuel after only nominal oil firing. Furthermore, in
some boilers the superheater tube temperature limit of 850K (1070°F) was
being closely approached and required close monitoring. One boiler was
curtailed to about 50 percent of full load when switching back to gas fuel
due to superheater tube temperature limits being exceeded. A series of
upper wall tube failures have occurred in that boiler. Also, some boilers
have been operating near the furnace pressure limit with FGR, aggravated by
slagging after periods of oil burning.
The baseline NO emissions at full load from the Moss Landing
A
Boilers, as mentioned earlier, averaged over 1400 ppm. The baseline
emissions at full load of the other horizontally opposed boilers reported in
Reference 6-9 ranged from about 770 ppm for the Pittsburg Boilers Nos. 5 and
6 to approximately 425 ppm for the Contra Costa Boilers Nos. 9 and 10. The
Pittsburg and Contra Costa Boilers produce 272 kg/s (2.16 x 106 Ib/hr) of
steam as compared to the rated capacity of 640 kg/s (5.1 x 10 Ib/hr) for
the Moss Landing Boilers. All boilers were modified in a similar manner for
6-82
-------
NO control. All were retrofitted to allow F6R to the windbox. In cases
^
where FGR to the hopper existed for steam temperature control, the FGR fans
were replaced with larger fans. In others where no FGR capability existed,
new fans were installed. In all boilers the FGR systems were capable of
diluting the oxygen in the windbox to 17 percent. In the Moss Landing
Boilers, OSC operation was carried out by injecting as much as 17 percent of
the total air through the top row of burners (BOOS). In the other boilers
OFA ports were retrofitted to allow up to 20 percent of the total air to be
introduced through the ports. In all cases the techniques of combined OSC
and FGR were effective in reducing NO levels down to around 125 to
175 ppm when the boilers were clean. Although the baseline NO emissions
/\
from these boilers span a wide range indicating a wide range of flame
intensities and surface heat rates, the problems encountered in low NO
J\
operation of the boilers were remarkably similar. As mentioned earlier,
higher convective section and upper furnace temperatures resulted due to
higher furnace exit gas temperatures and increased heat transfer
coefficients. Furnace exit gas temperatures usually rose with staged
firing, as combustion in this mode takes place over a larger part of the
furnace, and in some cases filled the whole furnace. Heat absorption
profiles may no longer peak in the lower half of the furnace and the gas
temperature profiles change accordingly. Convective heat transfer
coefficients increase due to the higher mass flowrates through the boiler
with FGR. The higher mass flowrates have also resulted in duct and furnace
vibrations, flame instability problems, and high furnace pressures. The
furnace vibration and flame instability problems were resolved in some cases
by installing new gas spuds and flame retainers especially developed for
this purpose. Duct vibration problems required reinforcement of the FGR
ducts and installation of splitter vanes. Higher furnace pressures have
necessitated raising the furnace trip settings, in some cases very close to
the upper design limit of the furnace.
In summary, horizontally opposed gas-fired boilers can be
successfully modified for low NO operation even in cases where the
A
baseline emissions are quite high. A combination of OSC and FGR operation
has resulted in maximum NO reduction. As is the case with other types of
>v
boiler design, the NOX reductions attainable will be generally less and
not continuously attainable for boilers utilizing both gas and oil fuels. A
6-83
-------
number of problems have been associated with these modifications. High
reheat, superheat, and upper furnace wall temperatures have occurred, which
may result in increased tube failure. Flame instabilities, boiler vibration
and high furnace pressures are other potential adverse effects experienced
with OSC + FGR firing. High tube and steam temperatures may be particularly
troublesome in boilers which switch between oil and gas firing. Losses in
cycle efficiency up to 1 percent have occurred if reheat steam required
attemperation. No derating of boilers was reported, but is possible in
cases where furnace pressure limits, tube temperature limits, or maximum
attemperation capacities are exceeded. No data were available on boiler
efficiency. However, as minimum excess air requirements have not been
reported to increase with low NO operation on gas-fired boilers, the
A
boiler efficiencies are not expected to be affected.
6.12 SINGLE WALL GAS-FIRED BOILERS
The 100 MW, San Diego Gas and Electric Company, Encina Units No. 1,
2, and 3 discussed in Section 6.8 were also tested for low NO operation
A
with natural gas fuel (Reference 6-13). The reduction in NO emissions
A
for gas fuel were also carried out in two steps. The first was designed to
meet the 1971 San Diego APCD regulation of 225 ppm NO for natural
gas-fired utility boilers. The second step was to reduce NO levels to
rt
125 ppm to meet 1974 standards.
The 1971 standards were met by combustion modification similar to
those used for oil firing. The boilers were fired off stoichiometrically by
taking 2 burners out of service from a total of 10. Fuel flow was
terminated to burner Nos. 2 and 4 in the top row of burners and the air
registers on these burners were opened 100 percent. Of the remaining
burners there were indications that the four wing burners received less air
than others. In order to attain uniform air distribution to the active
burners, the wing burner registers were opened 100 percent and the rest of
the active burner registers were throttled down to 70 percent open. A
sketch of the burner and register configurations are shown at the bottom of
Figure 6-15.
Taking two burners out of service necessitated an increase in gas
flow to the active burners to maintain maximum load. This was achieved by
increasing the gas delivery pressure through raising the gas pressure
regulator setting at each unit. This was not expected to create any adverse
6-84
-------
400 „
300 -j
<\J
o
ro
200 -
0.
Q.
X
O
100
20
40
I I
60 80
Load (MW)
Normal operation
Two-BOOS operation
100
120
Burner No. 12345
Register (% Open) 70 70 70 70 70
Burner No.
Register (% Open)
OOOOO
O O O O O
6 7 8 9 10
70 70 70 70 70
Normal operation
1 2
100 100
3
70
4
100
5
100
o & o » o
OOOOO
6 7 8 9 10
100 70 70 70 100
Modified operation fuel
flow to Nos. 2 & 4
burners terminated
Figure 6-15.
Comparison of NOX with normal and two-BOOS
operation with natural gas fuel for Encina
Unit No. 1 (Reference 6-13).
6-85
-------
effect on boiler operation. On one of the boilers some load pickup tests
were carried out. It was found that combustion modification did not in any
way affect unit response.
A number of tests were carried out to establish the minimum excess
air levels under low NO operation which would ensure that carbon monoxide
s\
levels would not exceed 100 ppm. Curves for excess oxygen levels as a
function of load were drawn up for each boiler. In general, the recommended
levels were conservative as the corresponding CO levels had a maximum value
of 30 ppm. Even so, the recommended curves for excess 0~ under low NOX
operation fell below the excess 02 curves for normal (baseline)
operation. Figure 6-16 gives the excess 02 curves under normal and low
NO operation for the Encina No. 1 Boiler. The corresponding NO levels
A A
are shown in Figure 6-17.
Table 6-20 gives a comparison of some process variables under
baseline and operation with two burners out of service. There were no major
changes except for an increased imbalance between the north and south ducts
under two-BOOS operation. This imbalance was noted on this boiler only and
was attributed to plugging of holes on the burner rings. No other major
problems were encountered. There was some flame carryover to the
superheater sections but it did not result in problems with high tube
temperatures or tube wastage. Increased attention to ring burners, tube
walls, and convective tubes was recommended for the low NO operation.
X
These could be readily incorporated in the normal furnace maintenance
program.
The 1974 NOV regulations (125 ppm) were met by taking three burners
A
out of service. The burner pattern which best reduced NO emissions was
A
slightly different from two-BOOS operation with oil firing for this unit.
The burner pattern and register settings which gave optimum results are
shown at the bottom of Figure 6-17. As with two-BOOS operation, tests were
run to establish recommended excess 02 levels as a function of load. The
results are shown in Figure 6-18 for the Encina No. 1 Boiler. The
recommended excess 02 levels were not much different from two-BOOS
operation for this unit. The same is true for the other units. The Encina
units are largely airflow limited. An increase in excess air level,
therefore, generally leads to a reduction in maximum load. However, as the
excess air requirements do not increase with OSC operation in these units
6-86
-------
6 -,
5 -
Normal operation
4 -
C
O)
o
S-
0)
Q.
CM
O
(ft
to
o>
Two-BOOS
Operation fT
.
2 -^
20
I
40
60 80
Load (MM)
100
120
*As read in control room
Figure 6-16.
Comparison of excess 02 for normal and two-BOOS
operation with natural gas fuel for Encina
Unit No. 1 (Reference 6-13).
6-87
-------
130
120 _
110 _
100 -
90 -
80
70
60
* 50
c
o
5 40
30
20
10
0
CM
O
c.
>>
Q.
CL
X
O
1974 NOX regulation
Maximum
Airflow
Recommended
Excess
02
Minimum Excess 03
12345
8
10
OOOOO
55 70 70 70 55
70 100 100 100 70
Burner
Pattern
Register
Settings
20 30 40 50 60 70 80 90 100
Load (MW)
110 120
Figure 6-17. NOX emissions versus load for gas fuel with seven-burner
operation for Encina Unit No. 1 (Reference 6-13).
6-*
-------
TABLE 6-20. COMPARISON OF GAS-FIRED ENCINA UNIT NO. 1, OPERATED UNDER
BASELINE CONDITIONS AND WITH TWO BURNERS OUT OF SERVICE
(Reference 6-13)
Process Variables
Load
Control Room 03
Burners Out of Service
Steam Temperature
Indicated Gas Flow
Indicated Airflow
Supply Gas Pressure
Burner Gas Pressure
Furnace Draft Pressure
AH Gas Out Temperature
AH Gas in Temperature
RC Fan
FD Fan
ID Fan
Measured NOX
Measured CO
MW
Percent
K (OF)
(Meter setting,
arbitrary units)
MPa (psi)
MPa (psi)
kPa (inch H20)
K (OF) N
S
K (°F) N
S
Amps
Amps N
S
Amps N
S
ppm N
S
ppm N
S
Baseline
Operation
98
2.0 to 2.2
None
811 (1000)
920/1100
58/70
0.160 (23.0)
0.076 (11.0)
-0.1 (-0.5)
453 (355)
450 (350)
653 (715)
650 (710)
0
80
80
115
115
320
375
30
30
Two- BOOS
Operation
98
1.6 to 1.8
Nos. 2 & 4
811 (1000)
860/1100
58/70
0.186 (27.0)
0.125 (18.1)
-0.1 (-0.5)
455 (360)
444 (340)
658 (725)
650 (710)
0
80
81
106
108
150
160
40
50
6-89
-------
7.0
6.0-1
5.0-
•K
S-
t! 4.0-
c
0)
o
s_
01
Q.
-------
when fired with gas, no derating of the units was needed. The NOX
emissions from Encina No. 1, corresponding to the excess 02 levels given
in Figure 6-18, are shown in Figure 6-17.
Table 6-21 gives a comparison of some process variables for the
Encina No. 1 Boiler under two-BOOS and three-BOOS operation. No major
changes in the variables were experienced as the degree of off
stoichiometric firing was increased. Table 6-21 does not, however,
represent normal boiler operation as the tests were conducted prior to a
major overhaul including servicing and cleanup. The No. 2 air register was
not functioning properly before the overhaul. Thus, Table 6-21 should not
be compared directly with Table 6-20.
As three-BOOS operation with natural gas involved operating with the
three center top burners out of service, the boiler performance changed in
terms of the effect of excess air on final steam temperatures. Increasing
overall excess air tended to reduce steam temperatures, in contrast with
usual practice where increasing airflow results in higher steam
temperatures. Apparently, this was due to the cooling effect of the air
from the BOOS ports in the superheater and reheater. However, no problems
with maintaining steam temperatures at design levels were anticipated, as
OSC operation with gas fuel did not lead to higher excess air requirements.
The Encina units have provisions for flue gas recirculation to
maintain adequate transfer coefficients in the convective sections. It was
found that FGR was not required above 80 MW with three-BOOS operation
compared to about 90 MW with two-BOOS operation. At high loads with OSC
firing some pressure pulsing occurred in the corners of the firebox even
though the furnace probe indicated a stable negative furnace pressure. This
was attributed to irregularities in the bulk gas flow dynamics, and it was
recommended that care be exercised when opening the observation ports on the
operating level. No other adverse effects were observed. The boiler
efficiency was not notably affected by OSC operation. In summary, the data
indicates that no significant operational or maintenance problems are likely
to occur with OSC operation of front wall gas-fired boilers of the type
studied here.
6.13 TURBO FURNACE GAS-FIRED BOILERS
A limited amount of process data are available on gas-fired turbo
furnace boilers. In general, the available NO control techniques for
6-91
-------
TABLE 6-21. COMPARISON OF GAS-FIRED ENCINA UNIT NO. 1, OPERATED WITH
TWO AND THREE BURNERS OUT OF SERVICE PRIOR TO OVERHAUL
(Reference 6-13).
Process Variables
Load
Control Room 03
Burners Out of Service
SH Steam Temperature
RH Steam Temperature
Steam Flow
Indicated Gas Flow
Indicated Airflow
Attemperator Temp. In
Attemper ator Temp. Out
Furnace Draft
Flue Gas Recirculation
AH Gas In
AH Gas Out
Ringleman Smoke
Chart No.
Measured NOX
Measured CO
MW
Percent
K (°F)
K (°F)
kg/s (103 Ib/hr)
(Meter settings,
arbitrary units)
K (°F) N
S
K (OF) N
S
kPa (inch HgO)
K (°F) N
S
K (°F) N
S
ppm N
S
ppm N
S
Two-BOOS
Operation
100
2.2 to 2.6
Nos. 2 & 4
811 (1000)
803 (985)
88.5 (702)
950
60
685 (775)
678 (760)
647 (705)
672 (750)
-0.13 (-0.54)
Off
625 (665)
639 (690)
439 (330)
442 (335)
0.34
220
207
0
0
Three-BOOS
Operation
100.8
2.5 to 3.5
Nos. 2, 3,
& 4
811 (1000)
816 (1010)
90.1 (715)
940 to 980
64
694 (790)
686 (775)
672 (750)
675 (755)
-0.05 (-0.2)
Off
669 (745)
655 (720)
455 (360)
450 (350)
0.4
111
125
0
0
6-92
-------
these boilers are the same as for oil-fired turbo furnaces. The NO
A
control techniques for which process data are available on turbo furnaces
include OFA, FGR, airflow adjustment, reduced air preheat, and water
injection.
South Bay Unit No. 3, a Riley Stoker turbo furnace unit owned by
San Diego Gas and Electric Company, was tested extensively for reductions in
NO emissions (Reference 6-8). The boiler can be fired with both oil and
A
gas, has 12 burners, and has a maximum continuous rating of 145 kg/s
(11.5 x 105 Ib/hr) of steam. After balancing the airflow to each of the
two windboxes, the settings on the velocity dampers and directional vanes
were varied to reduce NO emissions. It was found that minimum NO
X A
emissions were obtained when the velocity dampers on the part of the burner
below the fuel guns were completely closed and the velocity dampers above
the fuel guns were fully open. Apparently this arrangement simulated an
overfire air effect. The NO emissions were further decreased when the
A
directional vanes above the fuel guns were directed upwards at an angle of
30 degrees or higher relative to the direction of fuel injection as was the
case for oil firing. The first two columns in Table 6-22 show a comparison
of some process data with the boiler at partial load under baseline and
adjusted airflow conditions. It is seen that a substantial reduction in
NO emissions occurs under the adjusted airflow conditions, although part
of the reduction may be attributed to the slightly lower air preheat
temperature and excess oxygen level. The excess oxygen in the adjusted
airflow test was reduced to a minimal level as can be gauged by the higher
carbon monoxide generation and plume smoking condition.
Although a significant decrease in NO emissions occurred by
A
adjusting damper and vane settings, the reduction was not sufficient to meet
statutory limits especially at higher loads. Consequently, water injection
by means of a bank of nozzles into the preheated combustion air was tested
as a NOX control measure. The results of water injection for two
different injection rates are shown in the last two columns of Table 6-22
for partial load. The results of the NO control techniques, viz.,
A
airflow adjustment and water injection, are shown in Figure 6-19 for the
whole range of boiler loads. The amount of water required to maintain NOV
A
emissions below the legal limit is also shown as a function of load. The
increasing amount of water injection with load decreases the boiler
6-93
-------
TABLE 6-22.
COMPARISON OF GAS-FIRED SOUTH BAY UNIT NO. 3 UNDER BASELINE
AND LOW NO CONDITIONS UNDER PARTIAL LOAD (Reference 6-8)
Process Variables
Load
Excess Oxygen
Steam Flow
Natural Gas Flow
Hater Injection:
Flow-ate
Hater/Fuel Ratio
Velocity Dampers
Top
Bottom
Directional Vanes
Upper
Lower
Pressures:
Steam drum
Burner Natural Gas
Windbox
Furnace
Temperatures:
SH Steam
RH Steam
AH Air In
AH Air Out
AH Gas In
AH Gas Out
F.D. Fan Current
Emissions:
NOX (at 3X 02)
CO
Ringleman Smoke Density
MW
%
kg/s (105 Ib/hr)
nm3/s (105 scfh)
kg/s (103lb/hr)
kg/kg
X open
X open
degree
degree
MPa (psi)
HPa (psi)
kPa (In. H20)
kPa (in. HZO)
K (°F)
K (°F)
K (°F)
K (°F)
K (°F)
K (°F)
AMPS
pom
ppm
ppm
Baseline
137
1.6
113 (9.0)
9.75 (12.4)
0
0
70
70
0
0
14.9 (2160)
0.086 (0.5)
1.6 (6.5)
1.2 (5.0)
810 (999)
798 (978)
296 (73)
569 (565)
638 (690)
408 (275)
122
238
30
0
Airflow
Adjustment
132
1.2
107 (8.5)
9.91 (12.6)
0
0
100
0
up 45
0
14.7 (2130)
0.083 (12.0)
1.6 (6.6)
1.0 (4.1)
807 (994)
807 (994)
296 (73)
541 (515)
624 (665)
391 (245)
122
187
100
0
(Slight plume)
Water Injection
134
3.6
109 (8.65)
9.83 (12.5)
3.65 (29.0)
0.508
100
0
up 45
0
14.2 (2060)
0.083 (12.1)
2.3 (9.1)
1.8 (7.3)
809 (997)
806 (991)
297 (75)
436 (325)
648 (708)
405 (270)
150
135
10
0
129
3.4
108 (8.6)
9.75 (12.4)
5.86 (46.5)
0.821
100
0
up 45
0
14.0 (2025)
0.083 (12.1)
0.9 (3.8)
1.8 (7.3)
809 (997)
814 (1007)
297 (76)
429 (313)
655 (720)
408 (275)
98
102
0
-------
500
400
Mode
Normal
Modified
Directional
Vanes
Velocity
Dampers
Normal
CJ
o
Horizontal 60-80° open
Uppers -- 45° up Uppers — 100% open
Lowers — horiz. Lowers -- closed
300
••= 200 -
01
0)
X
o
100 -
Modified
Vane/
damper
positions
Modified vane/damper
positions and
water injection
emissions
X
Water injection
flow rate
- 100
50
600
800
-••15
• -10
• • 5
-I- 0
IB
i.
s.
Ol
4-1
ID
3:
1000
10
1200
3
1400
1800
10
m3/s
Gas fuel flowrate
12
14
Figure 6-19.
NOX emissions for gas-fired South Bay
Unit No. 3 (Reference 6-8).
-------
efficiency, necessitating an approximately 10 percent increase in fuel
consumption at full load. However, no flame stability problems were
encountered even with water to fuel ratios as high as 1.3 by weight.
Water injection reduces NOV emissions, at least partially, by
A
decreasing the combustion air temperature. The decrease in combustion air
temperature with increasing water injection rates can be seen in
Table 6-22. Some tests were also carried out with reduced air preheat by
bypassing some of the air and gas flow in the air heater. It was found that
the effect on NO emissions due to a decrease in combustion air
A
temperature using air heater bypass was similar to that obtained by water
injection with the same decrease in temperature. With oil fuels, however,
it was found that water injection was much more effective than air heater
bypass. For this reason, water injection was recommended over RAP by air
heater bypass for this boiler. Water injection, however, is considered only
to be an interim NO control measure until low NO techniques which
X A
result in less severe boiler performance penalties, such as OFA, can be
installed.
Pacific Gas and Electric Company has also reported NO reduction
A
modifications to its turbo furnace Potrero Boiler NO. 3-1 (Reference 6-9).
This boiler is a turbo furnace capable of burning both oil and gas and
generating 189 kg/s (1.5 x 10 Ib/hr) of steam. The furnace was
retrofitted with OFA ports designed to handle up to 25 percent of the
combustion air and FGR capable of reducing windbox oxygen content to
17 percent. Convective section modifications were also made to compensate
for the change in absorption profiles incurred with FGR and OFA. Part of
the reheater surface was removed in order to avoid excessive reheat steam
attemperation, and the economizer was replaced by a larger fin tube
economizer to improve the efficiency of the unit.
The baseline NO emissions at full load for Potrero No. 3-1
amounted to approximately 530 ppm. The use of OFA alone resulted in a
reduction of 50 percent in NO emissions. Operation with OFA and FGR
reduced N0¥ levels down to approximately 175 ppm. Some problems, however,
rt
arose with combined OFA and FGR operation. Tube metal and steam temperature
limits were approached at high loads resulting in increased superheater tube
failures as noted above for oil firing. Boiler load was limited to
95 percent of full rated value and the unit was removed from automatic
6-96
-------
dispatch operation due to the problems with superheater temperatures at high
loads.
In summary, the usual NO control techniques used with gas fuels
/\
such as OFA and FGR were successful in controlling NO emissions from
r\
turbo furnace boilers. These techniques may, however, be associated with
problems such as high convective section temperatures. Due to the
flexibility in controlling the airflow at the burners inherent in the turbo
furnace design, the airflow may be adjusted to create an overfire air
effect. This may in some cases result in significant NO reductions.
/\
Water injection and reduced air preheat were also successful in reducing
N0x emissions. However, the high increased fuel consumption penalty
associated with these techniques make them unattractive except as an interim
NO control measure.
6.14 SUMMARY OF PROCESS ANALYSES
A summary of the impact of low NO operation on boiler operation
^
and performance is given in this section. Details of process analyses for
each combination of furnace/fuel type have been discussed in the preceding
sections. There were some furnace/fuel combinations, however, where there
were insufficient data to allow for an adequate treatment of all applicable
low NO techniques. This section attempts to integrate the data on
J\
various boilers by considering each fuel type separately regardless of
furnace type. Thus, the major NO control techniques used for each fuel
J\
and the effects of low NO operation on the boiler are discussed below.
y\
It should be noted that the various test conditions discussed in the
preceding sections do not necessarily reflect current operating procedures
for any one specific boiler. Generally, utilities are continuously seeking
ways of increasing boiler efficiencies while achieving low emissions.
6.14.1 Coal-Fired Boilers
The effects of low NO operation on coal-fired boilers are
/\
summarized in Table 6-23. The most commonly applied low NO techniques
/\
for coal-fired boilers are low excess air (LEA) and off stoichiometric
combustion (OSC). Low NO burners are also being installed on some new
^
units and have been found to be effective. Other techniques which have been
tested but are less commonly employed are flue gas recirculation (FGR),
which has been found to be relatively ineffective, and water injection (WI),
which is not preferred because of efficiency losses. The major concerns
6-97
-------
TABLE 6-23. EFFECT OF LOW N0v OPERATION ON COAL-FIRED BOILERS
cr>
MD
00
Boiler
Tangenti al
Barry No. 2
Columbia
NO. 1
Huntlngton
Canyon No. 2
Barry No. 4
Navajo No. I
Comnanche No. 1
Opposed Wall
Harllee Branch
No. 3
Four Corners
No. 4
Hatfleld No. 3
Low NOX
Technique
BOOS
OFA
OFA
OFA
LEA. BOOS
LEA, BOOS, OFA
OFA
LEA, BOOS
LEA, BOOS
Water Injection
BOOS
FGR
Efficiency
Unaffected
Unaffected
Unaffected
Unaffected
Unaffected
Unaffected
Unaffected
0.6X average
decrease
0.6X increase
(excluding WI)
0.3X decrease
0.4X decrease in
boiler effi-
ciency. Some
decrease in cycle
efficiency due to
RH attemperation.
Corrosion
Measured 75*
increase, but
within normal
range
Measured 70*
increase, but
within normal
range
No change
Measured 25X
decrease, but
within normal
range
No significant
change
No significant
change
No significant
change
Slight Increase
No significant
change
..a
Load
Capacity
20X derate
Unaffected
Unaffected
Unaffected
20X or more
derate with
BOOS
Unaffected
Unaffected
Up to 17X
derate
with BOOS
Up to 25X
derate
with BOOS
10X derate
Unaffected
Carbon Loss
in Flyash
Slight increase
Slight Increase
Slight increase
Slight Increase
SOX average
decrease
No change
30X average
decrease
130X average
Increase
SOX average
decrease
SOX average
1 ncrease
120X average
Increase
Dust Loading9
100X increase
100X increase
..a
50 average
Increase
40X average
increase
20X average
decrease
10* average
increase
15X average
decrease
Unaffected
Unaffected
Part. Size
Distribution*
..a
—
No change
No significant
change
~
—
Other Effects,
Comments
Minor changes in heat
absorption profile
SH attemperation
increase by 70X
Minor changes in heat
absorption profile
SH attemperation
increased over 200X
Minor changes in heat
absorption profile
SH attemperation
increased by 70X
Minor changes in heat
absorption profile
No SH attemperation
requi red
No slagging or foul-
ing. No significant
increase in tube. tem-
peratures. Increase
in POMs by SOX.
Stable flames and
uniform combustion.
Increase in RH
attemperation. No
increase in POM
emissions. No signi-
ficant increase in
tube temperatures.
Denotes that investigated
-------
TABLE 6-23. Concluded
Boiler
E.G. Gas ton
No. 1
FU Unit A
Single Uall
Widows Creek
No. 5 (TVA
test)
Widows Creek
No. 5 (Exxon
test)
Widows Creek
No. 6
Mercer Station
No. 1 (wet
bottom)
Crist Station
No. 6
FW Unit B
FW Unit C
Turbo Furnace
Big Bend No. 2
Low NOX
Technique
LNB, LEA, BOOS
BOOS
BOOS
LEAS. BOOS
LEA, BOOS
LEA, Biased
firing
LEA. BOOS
BOOS
OFA
LEA, BOOS
«1r vane
adjustment
Efficiency
0.3% decrease
on average (LNB
baseline)
Unaffected
IX increase
It average
increase
Unaffected
Unaffected
0.4X decrease
•^ •
Unaffected
Unaffected
Corrosion
No significant
increase
Results of tests
Inconclusive
No significant
1 ncrease
-
No significant
Increase
--
-
Load
Capacity
Up to 30X
derate
(LNB with
BOOS)
Up to 25X
derate
Unaffected
Unaffected
Unaffected
Unaffected
Up to 15X
derate
Z5X derate
Unaffected
Up to 40X
derate with
with BOOS
Carbon Loss
in Flyash
130X average
increase (LNB
baseline)
85X increase
SOX increase
30X average
decrease
70X average
Increase
SOX average
Increase
60X increase
Unaffected
Unaffected
~
Oust Loading3
15X average
increase (LNB
baseline)
No significant
increase
15X average
decrease
20X average
decrease
10X average
Increase
SOX increase
-
Part. Size
Distribution*
Shift towards
smaller par-
ticles (LNB.
with or with-
out BOOS)
a
.
—
No significant
change
--
~
Other Effects,
Comments
Unit retrofitted
with low NOX
Baseline, LEA and
BOOS tests with LNB
compared to baseline
tests on sister
boiler with no LNB.
Severe slagging and
hazy flames filling
furnace at burner
stoichiometries
below 95X
Severe slagging and
hazy but stable flames
at burner
stoichiometries below
95X
NSPS unit with larger
fire box and factory
installed OFA ports.
No problems with
slagging reported.
I
IX>
VO
'Denotes not Investigated
-------
regarding low NO operation on coal-fired boilers have been the effects on
A
boiler performance, load capacity, furnace wall tube corrosion and slagging,
carbon loss, particulate loading and size distribution, other pollutant
emissions, heat absorption profile, and convective section tube and steam
temperatures.
Low excess air firing has become common operating practice in many
utility plants as it improves boiler efficiency. Reducing stoichiometry at
burners reduces both thermal as well as fuel NO . However, it is usually
A
difficult to reduce excess air levels to values much below 10 or 15 percent
in coal-fired boilers without excessive carbon monoxide or smoke
generation. To reduce NO emissions to meet statutory requirements, it is
A
often necessary to reduce burner stoichiometry down to 100 percent or
lower. This can be accomplished by OSC using overfire air (OFA), burners
out of service (BOOS), or biased burner firing (BBF). Minimum excess air
requirements under OSC are usually higher than with LEA. In most cases,
however, the excess air requirements under baseline conditions are
comparable to those with OSC. The efficiency of the boiler, therefore,
remains unaffected if unburned carbon loss does not increase appreciably.
In some cases when, due to nonuniform fuel/air distribution or other causes,
the excess air requirement increases substantially with OSC, a significant
decrease in efficiency may occur. From Table 6-23, it is seen that
efficiency decreases of up to 1 percent may occur under OSC. It is also
seen that the same boiler (Widows Creek No. 5) tested at a different time
under LEA and BOOS showed an average increase in efficiency by 1 percent.
It should be emphasized that optimal boiler conditions are very important
both in NO reduction and in minimizing potential adverse effects.
A
Uniform fuel and air distribution to the burners is especially important for
OSC operation if operation at reasonably low excess air levels is to be
achieved.
Many new boilers now come factory-equipped with OFA ports. Older
boilers can be retrofitted with OFA ports or can operate with minimal
hardware changes under BOOS or biased firing. Burners out of service
usually involves firing the higher level burners on air only while biased
firing involves firing upper level burners fuel lean and lower level burners
fuel rich. The optimal BOOS or biased firing pattern for low NO must
A
normally be determined by trial and error although removing upper level
6-100
-------
burners from service is generally most effective. The BOOS technique is
normally implemented by shutting off one or more pulverizers supplying these
upper levels. If the other pulverizers cannot handle the extra fuel to
maintain the total fuel flow constant, boiler derating will be required.
From Table 6-23, it is seen that boiler derating of 10 to 25 percent is not
uncommon with BOOS firing. Biased firing may reduce or eliminate the amount
of derating a boiler has to suffer. However, this type of firing has not
been tested sufficiently to establish its effectiveness as a NO control
/\
technique.
The possibility of increased corrosion has been a major cause for
concern with OSC operation. Furnaces fired with certain Eastern U.S.
bituminous coals with high sulfur contents may be especially susceptible to
corrosion attack under reducing atmospheres. Local reducing atmosphere
pockets may exist under OSC operation even when burner stoichiometry is
slightly over 100 percent. The problem may be further aggravated by
slagging as slag generally fuses at lower temperatures under reducing
conditions. The sulfur in the molten slag may then readily attack tube
walls. Severe slagging has been observed in some boilers operating at
burner stoichiometries below 95 percent. A number of short-term corrosion
tests have been carried out by inserting air cooled corrosion coupons at
various locations adjacent to the water walls. The results of the tests are
not quite conclusive as the rates measured by the coupons, even under
baseline conditions, do not correspond to normal corrosion rates. The
coupons can, however, be used to determine relative corrosion under baseline
and low NO conditions. In general, it has been found that no significant
/\
acceleration in corrosion rates occurs under OSC conditions. Nevertheless,
because of the wide scatter in data, the issue cannot be considered resolved
until definitive results from long-term tests with measurements on actual
water wall tubes are available.
Increased carbon loss in flyash may occur with OSC if complete
burnout of the carbon particles does not occur in the furnace. High carbon
loss will result in decreased boiler efficiency and may also cause
electrostatic precipitator (ESP) operating problems. From Table 6-23, it is
seen that increases in carbon loss vary over a wide range and can be as high
as 70 to 130 percent in some cases. However, increased carbon loss is not
perceived as one of the major problems associated with OSC operation. If
6-101
-------
the carbon content in flyash increases to levels where it threatens to
impair the operation of dust collection systems, the unburned carbon can
usually be easily controlled by increasing the overall excess air level in
the furnace. Although this will tend to increase stack heat losses, the
decrease in boiler efficiency will be partially compensated for by reduced
unburned carbon losses.
Increased particulate loading with OSC may be a source of problems if
baseline loadings are close to acceptable limits. Installing larger or more
efficient dust removal devices may be necessary. The problem can be
particularly severe if the particle size distribution shifts towards smaller
sizes because the efficiency of many dust collectors, such as ESPs,
decreases in the 0.1 to 1.0 ym range. From Table 6-23 it is seen that dust
loading changes can vary widely. In some cases, dust loading may double
with OSC operation, although from the few size distribution data available
no shift in distribution is evident. It is suspected that increased dust
loading may occur due to completion of combustion at a higher elevation in
the furnace. More particles thus tend to be entrained in the stream instead
of settling to the furnace hopper bottom. It should be noted, however, that
most of the particulate loading measurements were carried out at the
economizer outlet and do not necessarily reflect stack outlet conditions.
Extension of the combustion region to higher elevations in the
furnace may result in potential problems with excessive steam and tube
temperatures. However, among the numerous short-term OSC tests conducted no
such problems have been reported. In some tests where furnace and
convective section tube temperatures were measured directly, no significant
increase was found. Changes in heat absorption profiles were also found to
be minor, thus indicating no need for addition or removal of heat transfer
surfaces. Superheater attemperator spray flowrates tripled in one case due
to OSC operation, but in all cases were well within spray flow capacities of
the units. Reheater attemperator spray flowrates did not show any increase
due to OSC operation so that cycle efficiencies were not affected.
The effect of OSC operation on gaseous pollutants other than NO
J\
has undergone limited investigation. In one study where polycyclic organic
matter (POM) was measured, an increase of about 30 percent was reported with
OSC. The accuracy of POM measurement is, however, currently of the same
order as the measured increase so that no conclusions can be drawn at
6-102
-------
present. Carbon monoxide (CO) emissions usually increase rapidly once
burner stoichiometry or excess air levels are reduced below a certain
level. This minimum level usually differs from boiler to boiler and also
varies with load. Boiler operators usually establish recommended excess air
levels as a function of load for each boiler. The recommended values are
usually slightly higher than the minimum values to give the operator a
margin of safety especially under rapidly changing load situations. When a
boiler is operated under OSC, the recommended excess air levels must be
reestablished as a function of load due to the higher overall excess air
requirements. With proper care carbon monoxide generation should not
increase significantly over baseline values. Unburned hydrocarbons (UHC)
also should not exhibit any significant increase as CO is usually more
sensitive to excess air levels than UHC. Total SO emissions should not
A
be significantly affected by OSC operation. SOg conversion may actually
be inhibited under air lean conditions.
Many new wall fired coal boilers are being fitted with low NOX
burners (LNB). These burners are designed to reduce NO levels to meet
A
statutory requirements either alone or in some cases in combination with OFA
ports. The LNB technique has the advantage of eliminating or decreasing the
need for reducing or near reducing conditions near furnace walls. Corrosion
problems associated with reducing atmospheres should thus not arise with
this system. Although the LNB flames can be expected to be less turbulent
and hence longer than flames from normal burners, the combustion zone will
probably be extended less further up the furnace than the OSC. Potential
changes in heat absorption profile and excessive steam and tube temperatures
are, therefore, less likely to occur.
As fuel and airflows are controlled more closely in LNB-equipped
systems, nonuniform distribution of fuel/air ratios leading to excessive CO
generation or high excess air requirements should be eliminated. Boiler
efficiencies should, therefore, not be affected by installation of LNB.
However, Table 6-23 shows that the efficiency of one boiler decreased
slightly when retrofitted with LNB. The decrease in efficiency was mainly
due to the large increase in unburned carbon loss. Particulate loading also
increased slightly with LNB, and there was a distinct shift towards smaller
size particles. Still, more testing is required to check whether these
changes were isolated instances or whether they form a pattern with LNB
6-103
-------
operation. It should be noted that the decrease in efficiency and increases
in carbon loss and particulate loading were not greater than those
encountered with OSC operation. Corrosion rates are inferred from tests
with corrosion coupons showed no significant increase with LNB. Some BOOS
tests were also carried out on the LNB-equipped boiler. A substantial
decrease in NO emissions resulted below those already achieved with LNB
A
alone. However, the boiler was derated by up to 30 percent. Other
potential problems associated with OSC could also arise with this type of
firing.
Flue gas recirculation to the windbox has been tested as a NO
A
control technique for coal-fired boilers (Reference 6-5). The technique
inhibits thermal NO formation but is not very effective in controlling
A
fuel NO . The technique has not been used widely on coal-fired units.
A
The tests on Hatfield No. 3 showed that OSC was indeed much more effective
in controlling NO than FGR. Table 6-23 summarizes some of the effects of
A
FGR operation on that unit. The increase in carbon loss averaged
120 percent, although there were wide variations in the measured values.
Load capacity and dust loading remained unaffected. There was a slight
decrease in boiler efficiency attributable to the power consumption by the
FGR fans. There was no significant increase in tube temperature and POM
emissions remained essentially unchanged. Stable flames and uniform
combustion were observed throughout the tests even at high recirculation
rates (up to 15 percent at full load and 34 percent at reduced loads).
Reheat steam spray attemperation increased at high recirculation rates which
would result in a loss in cycle efficiency. Higher convective section heat
transfer rates may be expected with FGR as the higher gas mass flowrates
over the tubes tend to increase the convective coefficients. No corrosion
measurements were made so that the effect of FGR on corrosion is not known.
Corrosion due to chemical attack is not expected to be a major problem with
FGR. However, tube erosion may increase as the higher gas velocities may
result in greater particle impact on exposed surfaces.
Some data were available on the effect of water injection on NO
emissions. Water injection, however, results in a significant deterioration
of boiler performance. It has therefore not been recommended as a long-term
NO control measure for coal-fired boilers.
A
6-104
-------
It should be emphasized that the effects of NO control, in many
A
cases, will be critically dependent on boiler operating conditions. Factors
such as boiler cleanliness and uniform air and fuel distribution can have a
significant effect on the impacts of NO controls on both emissions and
^
boiler operations. It is therefore important that adequate maintenance
procedures are instituted. In some cases, normal maintenance and overhaul
schedules may have to be modified. In addition, when potential problems
such as tube corrosion and high tube temperatures are expected, the boiler
operator will have to pay closer attention to tube conditions and watch for
evidence of incipient failure. In a few cases hardware modifications may be
indicated, e.g., removal of reheater or superheater surface if attemperation
requirements become excessive. Furthermore, if attemperation leads to a
significant decrease in cycle efficiency, removal of reheater surface may be
indicated. Still, with proper design of retrofit systems and adequate
maintenance programs, low NO operation should not result in a substantial
/\
increase in operational problems over normal boiler operation.
6.14.2 Oil-Fired Boilers
The effects of low NO operation on oil-fired boilers are
A
summarized in Table 6-24. The most common low NO techniques tested for
yx
oil-fired boilers are low excess air (LEA), off stoichiometric combustion
(OSC), and flue gas recirculation (FGR). Other techniques which have been
tested but are less commonly employed are water injection (WI) and reduced
air preheat (RAP). The major concerns regarding low NO operation on
^
oil-fired boilers are effects on boiler performance, load capacity,
vibration, and steam and tube temperatures.
Low excess air is currently employed in many utility boilers due to
the beneficial effect it has on efficiency. Improvements in boiler
efficiency up to 5 percent have been reported in addition to lower fan power
consumption due to the smaller volume of air and gas flows. Still, LEA may,
in some cases, result in lower steam temperature which will adversely affect
cycle efficiency. To obtain the minimum possible excess air levels, it is
necessary to ensure uniform air and fuel flows. This often requires
adjusting air dampers and vanes and may also necessitate cleaning or
replacing burner tips. It is usually, however, very difficult to reduce
excess air levels much below 10 percent without raising carbon monoxide or
smoke emissions. Low excess air is, therefore, usually effective as a NOV
6-105
-------
TABLE 6-24. EFFECT OF LOW NOX OPERATION ON OIL-FIRED BOILERS
I
o
Boiler
Tangential
South Bay No. 4
Pittsburg No. 7
SCE tangential
boilers
Opposed Hall
Moss Landing
Nos. 6 and 7
Ormond Beach
Nos. 1 and 2
SCE BtU Units
Sewaren Station
No. 5
FU Unit C
Low NOX
Technique
LEA
BOOS
RAP
OFA and FGR
BOOS and FGR
OFA and FGR
BOOS and FGR
Hater Injection
BOOS and FGR
LEA. BOOS
BOOS and FGR
Efficiency
5S increase
Decrease in efficiency
compared to LEA due to
increased excess air
requirements
Unaffected due to
special preheater
design
..a
—
Increased excess air
requirements resulting
in decreased efficiency
Increased excess air
requirements resulting
In decreased efficiency
Increased sensible and
latent stack losses
FGR reduced minimum
excess air require-
ments increasing
unit efficiency
Load
Capacity
..a
—
Slower startups
and load changes
—
—
10 to 15* derate
due to maxed FD
fan capacity
"
15X derate to
vibration and
limited FO fan
capacity
Vibration and
Flame Instability
..a
—
FGR fan vibration
problems
--
FGR fan and duct
vibration, furnace
vibration problems.
Associated flame
instability.
Flame Instability
and associated
furnace vibration
Boiler vibration
problems
Flame instability
and associated
furnace vibration
Steam and Tube
Temperatures
..a
—
High water wall tube
temperatures
--
—
—
"
Other Effects, Comments
No adverse effects reported.
Fan power consumption
reduced.
No other adverse effects
reported
Limited tests. NOX
control effectiveness not
demonstrated.
No adverse effects reported
High furnace pressures.
Increased FGR and forced
draft fan power assumption.
Flame detection problems
due to change In flame
characteristics
Limited tests carried out
with HI at partial loads.
Excess air requirements
Increased.
Flame detection problems
due to change In flame
characteristics
Tests carried out at partial
loads. No adverse effects
reported. Participate load-
ing and size distribution
unaffected.
OFA ports very effective
in controlling NOX
Denotes not investigated
-------
TABLE 6-24. Concluded
Boiler
Single Wall
Endna Nos. 1,
2 and 3
Turbo
South Bay No. 3
Potrero No. 3-1
Low NOX
Technique
LEA and BOOS
(2 burners
on air only)
BOOS
(3 burners on
air only)
Airflow
adjustments
Water Injection
Reduced air
preheat
OFA and FGR
Efficiency
Increased unit effi-
ciency. Some adverse
effect on cycle effi-
ciency due to lower
steam temperatures.
Increased excess air
requirements resulting
1n reduced efficiency
Slight reduction in
EA resulting in slight
increase in efficiency
6% decrease at full
load
Reduction In effi-
ciency greater than
that with water
injection
Higher excess air re-
quirements, but addi-
tion of economizer
surface expected to
improve efficiency
Load
Capacity
-_a
5* derate due to
maxed ID fan
capacity
—
—
5X derate due to
excessive tube
temperatures
Vibration and
Flame Instability
__a
In most tests no
flame Instability
or blowoff noted
~
No flame instability
noted even at high
rates of WI
Side to side
windbox oxygen
cycling
Steam and Tube
Temperatures
Decrease in SH & RH
steam temperature
Intermittent flame
carryover to SH
inlet but tube
temperature limits
not exceeded
a
—
Tube and steam tem-
perature limits ap-
proached. Increased
SH tube failures.
Other Effects, Comments
No other adverse effects
reported
No abnormal tube fouling,
corrosion or erosion noted.
Increased tendency to smoke
and obscure flame zone.
No adverse effects reported
No other adverse effects
reported
Limited tests
Increased tendency to smoke
required higher minimum ex-
cess 0? levels. RH surface
removed to avoid excessive RK
steam attemperation. Larger
economizer Installed to
compensate for RH surface
removal .
CT>
I
'Denotes not investigated
-------
control technique when the baseline emissions are only slightly higher than
the statutory limits.
Off stoichiometric combustion has been found to be an effective
technique for NOV control from oil-fired boilers. The technique inhibits
J\
both thermal and fuel NO formation. Many new boilers come equipped with
A
OFA ports as standard equipment. Older boilers may be retrofitted with OFA
ports or operated with BOOS or biased firing. As with LEA it is important
that a uniform air and fuel distribution be maintained at all burners in
order to minimize CO and smoke emissions. Boiler operation with OSC
generally increases the minimum excess air requirements which may result in
a loss in boiler efficiency. In extreme cases when the boiler is operating
close to the limits of its fan capacity, boiler derating may be required.
Derates of as much as 15 percent have been reported due to the lack of
capability to meet the increased airflow requirements at full load.
In many cases, BOOS operation in oil-fired boilers has been found to
be more effective in controlling NOV than OFA firing. The BOOS technique
/\
involves firing a few burners, usually from the top rows, on air only,
although the optimal BOOS pattern which will result in maximum NO
A
reduction must usually be determined by trial and error. The fuel flow to
the rest of the burners thus increases if load is to remain constant. In
some cases, it has been necessary to enlarge the burner tips in order to
accommodate these increased flows.
No flame instabilities or boiler vibrations have been noted with BOOS
firing, nor are they expected with any type of OSC operation alone.
However, OSC operation results in an extended combustion zone. In some
cases flame carryover to the convective section may occur. However, in one
case where intermittent flame carryover occurred, no excessive tube
temperatures were recorded. Also no abnormal tube fouling or corrosion was
encountered. In another test, particulate loading and size distributions
were measured under OSC operation. No significant differences were found
from baseline values. OSC operation does usually result in hazy flames and
obscure flame zones. Thus, new flame scanners and detectors are often
required due to the change in flame characteristics. Except for the above,
no other major adverse effects have been reported in the numerous short and
medium tests conducted with OSC on oil-fired boilers.
6-108
-------
Flue gas recirculation can also be employed to reduce mainly thermal
NO emissions from oil-fired boilers. Implementing FGR for NO control
X X
usually requires retrofit hardware modification to boilers as FGR is more
effective when it is recirculated to the windbox than to the furnace
hopper. Some boilers come equipped with FGR to the hopper for steam
temperature control. These then require booster fans to introduce the flue
gas to the windbox. In boilers with no original FGR capability, FGR fans
and ducting must be installed along with appropriate splitter vanes and
mixing devices.
There are a number of potential problems which can occur with FGR
retrofit operation. The most common problems, such as FGR fan and duct
vibrations, can usually be avoided by good design. Other problems such as
flame instability, which can lead to furnace vibrations, are caused by the
increased gas velocity at the burner throats. Modifications to the burner
geometry and design such as enlarging the throat, altering the burner tips,
adding diffuser plates or flame retainers, and in one case, providing
tertiary airflow around the oil gun may then be required. These
modifications are usually made by trial and error for each boiler and are
often very time consuming. If the problem of excessive boiler vibration and
flame instabilities persists at high loads, the boiler may have to be
derated.
Other problems associated with FGR are high tube and steam
temperatures in the convective section. The increased mass velocities which
occur with FGR cause the convective heat transfer coefficient to rise. This
may, in extreme cases, lead to tube failures, exceeding attemperator spray
flow limits, or loss in cycle efficiency due to excessive reheat steam
attemperation. Increased mass flowrates in the furnace may also cause
furnace pressures to increase beyond safe limits. Flue gas recirculation
usually, however, has an advantage of not increasing minimum excess air
levels. Boiler efficiency is therefore relatively unaffected except for the
power consumed by the FGR or booster fans.
There is a paucity of data on boilers operated with FGR alone.
Boilers are usually tested with OSC first to check whether the NO
A
reductions are sufficient to meet regulations. If not, FGR capability is
added. The combination of OSC and FGR is very effective in reducing NO
/\
emissions. However, the problems associated with each technique are also
6-109
-------
combined. Tube and steam temperature problems in the upper furnace are
particularly aggravated, as both OSC and FGR tend to increase upper furnace
temperatures and heat transfer rates. Otherwise the comments for OSC and
FGR alone also apply to their combined operation. Boiler efficiencies
usually decline slightly with combined OSC and FGR firing due to higher EA
requirements and greater fan power consumption.
Water injection has been tested in a few instances as a NO control
J\
technique. It is generally relatively simple to implement. Water is
sprayed directly into the combustion air by a bank of nozzles installed in
the air duct downstream of the preheater. In one series of tests it was
found to be as effective as FGR. However, WI carries a heavy penalty in
reduced efficiency due to stack latent heat losses. Excess air requirements
often increase, contributing to a decrease in efficiency. In one case an
increase in fuel flow of 6 percent was required to maintain full load.
However, no flame instabilities or other adverse effects were noted. For
the above reasons, though, WI is generally used as an interim N0¥ control
/\
measure until a permanent, less energy wasteful technique can be initiated.
Reduced air preheat has also been tested as a NO control measure.
A
On oil fuels the test results on its effectiveness are mixed. RAP usually
leads to severe losses in efficiency due to increased stack gas
temperatures. On one boiler, it was estimated that for the same reduction
in NOY, WI resulted in lower efficiency losses than RAP. However, in
y\
special cases where the air heater and boiler are of unique design, a boiler
designed to incorporate a steam coil preheater instead of an air/gas heat
exchanger, RAP may be employed without any theoretical decrease in boiler
efficiency. As tests conducted with RAP have been limited in nature, no
data are available on the effect of RAP on other aspects of boiler operation.
Operating with low NO burners would seem to be a promising NO
^ ^
control technique for wall fired boilers. Many of the deficiencies
associated with other techniques would be eliminated. Several manufacturers
in the U.S. are in the process of developing and testing LNB for oil-fired
boilers. However, no test data have been released to evaluate the
effectiveness of the burners. Nevertheless, once LNB for oil-fired burners
are commercialized, they will probably become one of the more important
NO control measures employed for wall fired boilers.
6-110
-------
In certain boilers, such as turbo furnaces, burner air dampers and
vanes can be adjusted to create airflow patterns in the furnace resembling
an overfire air effect. By balancing airflows, it may also be possible to
reduce excess air requirements at the same time. As the reduction in NO^
levels is generally small, this technique, like LEA, is useful when baseline
NO emissions are close to regulatory limits.
/\
As with coal-fired boilers, before low NO techniques are
X
instituted on an oil-fired boiler, it is important to assure that it is in
good operating condition. Uniform burner air and fuel flows are essential
for optimal NO control. Retrofit NO control systems must be designed
A A
and installed properly to minimize potential adverse effects. Despite these
precautions, in some cases inevitable problems will occur, such as flame
instability or high tube temperatures. In some of these cases problem
shooting by trial and error and certain hardware modification will be
required to resolve the problems. In other cases, increased vigilance will
be needed on the part of the boiler operator, and an accelerated schedule of
maintenance and overhaul may be required. Changes to the boiler safety and
control systems may also be required, such as installation of new flame
scanners and modification of combustion control for new minimum excess air
levels. In some cases the boiler will be unable to function on automatic
control requiring manual operation. Some boilers may have their startup
procedures and load pickup responses altered due to FGR fan preheating
requirements, etc. Very many of the problems can now be avoided because of
hindsight and experience. Thus, retrofit systems can now be designed and
installed with care to avoid any potential adverse effects. New units with
built-in OFA and FGR systems or LNB should function without problems.
6.14.3 Gas-Fired Boilers
The effects of low NO operation on gas-fired boilers are
A
summarized in Table 6-25. The low NO techniques used and their effects
A
are very similar to those for oil-fired boilers. Usually there is no
distinction between oil- and gas-fired boilers as they are often designed to
switch from one fuel to the other according to availability. Since the
NO control method, the effects of low NOV operation, and the boilers
A A
themselves are similar for gas and oil, a detailed discussion of gas-fired
boilers will not be given here. Most of the above discussion of applicable
NO control measures to oil-fired boilers and potential problems resulting
A
6-111
-------
TABLE 6-25. EFFECT OF LOW NO OPERATION ON GAS-FIRED BOILERS
/\
Boiler
Tangential
South Bay No. 4
Pittsburg No. 7
Horizontally
Opposed
Moss Landing
Nos. 6 and 7
Pittsburg
Nos 5 and 6
Contra Costa
Nos. 9 and 10
Single Wall
Encina Nos. 1,
2 and 3
Low NOX
Technique
LEA
BOOS
OFA and FGfi
OFA and FOR
OFA and FGR
OFA and FGR
BOOS
(2 and 3
burners out
of service)
Efficiency
2 to 31 increase
Decrease in efficiency
compared to LEA due to
increased excess air
requirements
— a
0.81 decrease in cycle
efficiency due to RH
steam at temper at ion
Low EA levels were
possible even with
BOOS, resulting in
increased efficiency
Load
Capacity
..a
25X derate due to
excessive steam
temperatures.
Slower load
change response
Load curtailment
to 501 after oil
burns due to SH
tube temperature
limits being
exceeded
No derate. Load
pickup response
not affected
Vibration and
Flame Instability
a
Fan and duct
vibration problems
Furnace and duct
vibration problems.
Flame instability.
FGR fan and duct
vibrations. Flame
instability problems.
FGR duct vibrations
Some pressure
pulsing at
corners of
firebox
Steam and Tube
Temperatures
..a
High tube and RH
steam temperatures
RH spray and SH tube
temperature limits
approached after oil
burns upper wall tube
failures
Upper water wall
tube failures
High SH and RH steam
temperatures. SH
tube temperature
limits being
approached.
Some flame carryover
to SH but no
problems with high
tube temperature or
tube wastage
Other Effects, Comments
No adverse effects reported
No other adverse effects
reported
Furnace pressure limit
approached. FGR fan power
requirements increased by
as much as 66t. Problems
associated with switching
to gas after oil burning
could be eliminated only
with complete water washing
of furnace.
Boiler initially restricted
to manual operation due to
problems with flame insta-
bility on automatic control
Furnace pressure limits
approached after oil firing.
FGR fan preheating required
to reduce vibrations on cole
boiler startups.
No other adverse effects
reported
ro
Denotes not investigated
-------
TABLE 6-25. Concluded
Boiler
Turbo
South Bay No. 3
Potrero No. 3-1
Low NOX
Technique
Air flow
adjustments
Water injection
OFA and FOR
Ef f i c i ency
Slight reduction in
EA resulting in slight
improvement in
efficiency
1W decrease at full
load
Installation of larger
economizer expected to
improve efficiency
Load
Capacity
..a
—
5* derate due to
problems with high
temperatures
Vibration and
Flame Instability
__a
No flame instability
noted even at high
rates of WI
Side to side
windbox oxygen
cycling
Steam and Tube
Temperatures
__a
—
Tube metal and steam
temperature limits
reached at high
loads
Other Effects, Comments
No adverse effects reported
No other adverse effects
reported
Hardware modifications
included partial RH surface
removal to avoid excessive
RH steam at temper at ion.
Larger economizer then
installed to compensate for
smaller RH surface.
Denotes not investigated
-------
applies. Some effects specific to gas-fired boilers alone are treated
briefly below.
NO emissions oftentimes are difficult to control after switching
n
from oil to gas firing. Residual oil firing tends to foul the furnace due
to the oil ash content. Thus, NOY control measures which have been tested
/\
on a clean furnace with gas may be found inadequate after oil firing due to
the changed furnace conditions. These problems can be resolved by complete
water washing of the furnace after any oil burns. This is not very
practical, however, especially if oil to gas fuel switching occurs
frequently.
Boilers fired with gas usually have higher gas temperatures at the
furnace outlet than when fired with oil. Gas flames are not very luminous
and therefore radiate less energy to the furnace walls than oil flames. The
upper furnace and convective section inlet surfaces are thus subject to
higher temperatures with gas firing. These temperatures may increase
further when the combustion zone is extended due to OSC. Furthermore, heat
transfer rates in the convective section will rise with increased mass
velocities due to FGR. Upper furnace and convective section tube failures
and excessive steam temperatures are therefore more likely to occur when OSC
and FGR are implemented on gas-fired boilers. The situation may be
aggravated further if switching from gas fuel occurs after oil burns as
fouling will further reduce furnace absorption and, hence, increase gas
temperatures. Excessive steam temperatures or attemperation can be
corrected by partial removal of superheater or reheater surface. Excessive
tube temperatures will usually result in a derating of the system.
6-114
-------
REFERENCE FOR SECTION 6
6-1. Selker, A. P., "Program for Reduction of NOX from Tangential
Coal-Fired Boilers, Phase II and Ila," EPA-650/2-73-005a and 5b,
NTIS-PB 245 162/AS and NTIS-PB 246 889/AS, June 1975 and August 1975.
6-2 Crawford, A. R., E. H. Manny, and W. Bartok, "Field Testing:
Application of Combustion Modifications to Control NOX Emissions
from Utility Boilers," EPA-650/2-74-066, NTIS-PB 237 344/AS,
June 1974.
6-3. Burrington, R. L., et al., "Overfire Air Technology for Tangentially
Fired Utility Boilers Burning Western U.S. Coal," EPA-600/7-77-117,
NTIS-PB 277 012/AS, October 1977.
6-4. Crawford, A. R., et al., "The Effect of Combustion Modification on
Pollutants and Equipment Performance of Power Generation Equipment,"
in Proceedings of the Stationary Source Combustion Symposium.
Volume III, Atlanta, EPA-600/2-76-152c, NTIS-PB 257 146/AS, June 1976.
6-5. Thompson, R. E., et al., "Effectiveness of Gas Recirculation and
Staged Combustion in Reducing NOX on a 560 MW Coal-Fired Boiler,"
EPRI Report No. FP-257, NTIS-PB 260 582, September 1976.
6-6. Unpublished data supplied by G.A. Hollinden, Tennessee Valley
Authority, Chattanooga, TN, July 1977.
6-7. Crawford, A. R., et al., "Field Testing: Application of Combustion
Modification to Power Generating Combustion Sources," in Proceedings
of the Second Stationary Source Combustion Symposium, Volume II,
New Orleans, EPA-600/7-77-073b, NTIS-PB 271 756/9BE, July 1977.
6-8. Unpublished data supplied by Meinzer, R. P., Jr., San Diego Gas &
Electric Company, San Diego, December 1977.
6-9. Barr, W. H., F. W. Strehlitz, and S. M. Dalton, "Modifying Large
Boilers to Reduce Nitric Oxide Emissions," Chem. Eng. Prog.,
Volume 73, No. 7, pp. 59 to 68, July 1977.
6-10. Norton, D. M., K. A. Krumweide, C. E. Blakeslee, and B. P. Breen,
"Status of Oil-Fired NOX Control Technology," in Proceedings of the
NOy Control Technology Seminar. San Francisco, EPRI SR-39,
February 1976.
6-11. Unpublished data supplied by E. J. Campobenedetto, Babcock and Wilcox
Co., Barberton, OH and W. H. Barr, and E. Marble, Pacific Gas &
Electric Co., San Francisco, February 1978.
6-12. Norton, D. M., W. P. Gorzegno, and B. P. Breen, "Modifications to
Ormond Beach Steam Generators for NOX Compliance," ASME Winter
Meeting, ASME 75-WA/Pwr-9, November 1975.
6-115
-------
6-13. Unpublished data supplied by R. P. Meinzer, Jr., San Diego Gas &
Electric, San Diego, October 1977.
6-14. Barr, W. H., and D. E. James, "Nitric Oxide Control — A Program of
Significant Accomplishments," ASME Winter Annual Meeting, New York,
ASME 72-WA/Pwr-13, November 1972.
6-15. Campobenedetto, E. J., "The Dual Register Pulverized Coal Burner —
Field Test Results," Presented at Engineering Foundation Conference
on Clean Combustion of Coal, Rindge, NH, July 31 to August 5, 1977.
6-16. Unpublished data supplied by E. J. Campobenedetto, Babcock & Wilcox
Co., Barberton, OH, November 1977.
6-17. Rawdon, A. H., and S. A. Johnson, "Application of NOX Control
Technology to Power Boilers," in Proceedings of the American Power
Conference. Volume 35, pp. 828-837, 1973.
6-18. Rawdon, A. H., and S. A. Johnson, "Control of NOX Emissions from
Power Boilers," presented at the Annual Meeting of the Institute for
Fuel (Australian Membership), Adelaide, Australia, November 1974.
6-19. Hinrichs, J. M., and R. E. Floyd, "Low Excess Air Burner Modification
— 230 MW Unit," presented to Pacific Coast Electric Association
Engineering & Operating Conference, Los Angeles, CA, March 17-18, 1977.
6-20. Personal communication with R. Meinzer, Jr., San Diego Gas & Electric
Company, San Diego, CA, October 1979.
6-116
-------
SECTION 7
COST OF COMBUSTION MODIFICATION CONTROLS
It is generally agreed by boiler manufacturers and utility companies
alike, that the reliable estimation or projection of NO control costs for
/\
utility boilers is a difficult task indeed. Control equipment needs and
control costs are highly dependent on an individual boiler's characteristics,
as well as on individual installation and operational problems
(References 7-1 through 7-6). Therefore, in this study, control costs for
typical boilers were analyzed. They should be taken as such — typical
costs and not necessarily the norm for all cases.
In Section 7.1, previously reported cost estimates are reviewed and
cost analysis needs identified. Based on the need for a standardized cost
analysis procedure for comparing the cost effectiveness of controls, Section
7.2 develops the cost calculation procedure used in this study. Section 7.3
analyzes in detail typical retrofit control costs, based on preliminary
design studies, equipment vendor quotes, and engineering estimates. The
incremental costs of implementing controls to new boilers meeting current
NSPS are presented in Section 7.4, based on the latest design estimates from
a major boiler manufacturer.
In all the control cases considered, the projected control costs are
documented as thoroughly as available data allow. However, it should be
reiterated that the presented numbers should only be considered as
representative of typical cases — there is no such thing as a standard
boiler or control application. Furthermore, there are still unanswered
questions from long term operation with controls, such as possible increased
corrosion, slagging, and associated maintenance costs.
7-1
-------
7.1 BACKGROUND
One of the earliest efforts at assigning costs to combustion
modification control techniques for utility boilers was attempted by Esso
Research and Engineering in 1969 (Reference 7-7). Since 1969, however, it
has been shown that the effectiveness of control techniques among boilers
varies widely and requires continuing cost effectiveness evaluations on an
individual boiler basis. As an example of cost variations for combustion
modifications among individual existing units, several case studies from
Pacific Gas and Electric are presented in Table 7-1 (Reference 7-8). The
numbers shown are the costs incurred by PG&E during a recent program to
bring eight oil-fired units into compliance with local NO emission
*\
regulations. For the most part, the conversions involved the combination of
windbox flue gas recirculation and overfire air ports. Although the average
cost of the modifications was about $10/kW, in 1975 dollars, they ranged
from $1.8/kW to $17/kW.
Another West Coast electric utility company, the Los Angeles
Department of Water and Power (LADWP), has had extensive experience in
implementing NO control techniques on its gas- and oil-fired boilers.
/\
The techniques currently utilized by the Department include burners out of
service, overfire air, and low excess air. Table 7-2 shows the NOV
A
control installation costs incurred by LADWP for four different units
(Reference 7-9). The figures for the BOOS technique reflect the R&D costs
that preceded the retrofit. The very low expense associated with OFA on on
the B&W 235 MW unit was due to the base year of that estimate (1964 to
1965), and to the fact that this modification was included in the original
design. For the most part, the LADWP boilers were modified without much
difficulty, and the associated costs probably represent the lower limits of
the costs for the three NO reduction techniques implemented.
/\
The modification costs presented by PG&E and LADWP were only gross
estimates. Many different organizations were involved in the retrofit
efforts cited, and as a consequence, cost sharing and accounting often
obscured the true costs. Furthermore, research and development costs, which
could only be crudely estimated, significantly raised or lowered cost
figures, depending on whether or not they were included. Finally, the rush
to meet new local air pollution regulations often increased control
implementation costs greatly (References 7-4 and 7-5).
7-2
-------
TABLE 7-1. 1975 INSTALLED EQUIPMENT COSTS FOR EXISTING PG&E RESIDUAL
OIL-FIRED UTILITY BOILERS (Reference 7-8)
Unit Name
Pittsburg
No. 7
Pittsburg
Nos. 5 and 6
Nos. 9 and 10
Contra Costa
Potrero No. 3
Moss Landing
Nos. 6-1 and 7-1
Design Type
CE tangential
fired, divided
B&W opposed wall
BAH opposed wall
Riley turbo furnace
B&W opposed wall
Year
Online
1972
1964
1965
1972
1967, 1968
Capacity
(MM)
730
330 (each)
345 (each)
206
750 (each)
Modification
Cost
($106)
6.2
(6.9)a
7.8 (both)
(8.7)a
6 (both)
(6.7)a
3.5
(3.9)a
2.8
(3.1)a
$AW
8.5
(9.4)a
11.8
(13.1)a
8.7
(9.7)a
17
(18.9)a
1.8
(2.0)a
. J
Year
Modified
1975
1975
1975
1975
1971
Type of Modification
Windbox FGR, Overfire Air
• Two new 5000 hp FGR fans
• FGR ducting (17X FGR)
• NO port installation
• No new burner safeguard system
Windbox FGR, Overfire Air
• Transferred two FGR fans from other units
• FGR ducting (17* FGR)
i New hopper
• NO port installation; one for each
burner column
• New burner safeguard system; computer,
N0x control board, 0? controls on
dampers, flame scanners
Windbox FGR, Overfire Air
• New FGR fans (1 each) (17S FGR)
• Nominal amount of new ducting to windbox
• NO port installation
Windbox FGR, Overfire Air
• New FGR fan (17X FGR)
• NOX port installation, nominal amount of
ducting
• New burner safeguard system, N0x control
board, computer
Windbox FGR, Overfire Air
• Existing temperature control FGR fans
replaced with larger fans
• New flame scanners
co
1977 dollars in parenthesis
-------
TABLE 7-2. LADWP ESTIMATED INSTALLED 1974 CAPITAL COSTS FOR NOX REDUCTION TECHNIQUES
ON GAS- AND OIL-FIRED UTILITY BOILERS (Reference 7-9)
Unit
Capacity
(HW)
180
235
235
350
Unit
Type
CE Single Wall
CE Single Wall
B&W Opposed Wall
B&W Opposed Wall
NOX Reduction
Technique
BOOS
BOOS
BOOS
Overfire air
LEA
BOOS
Overfire Air
LEA
Implementation
Method
Retrofit
Retrofit
Retrofit
Original Design
Retrofit
Retrofit
Retrofit
Retrofit
Estimated Cost,
$ 103
69.4 (84.2)b
28.9 (35.1)
75.2 (91.3)
14. Oa (17.0)
28.9 (35.1)
266.0 (323.0)
101.0 (122.0)
28.9 (35.1)
S/kW
0.38 (0.46)b
0.16 (0.19)
0.32 (0.39)
0.06 (0.07)
0.12 (0.15)
0.76 (0.92)
0.29 (0.35)
0.08 (0.10)
a!964-65 base year
"1977 dollars in parenthesis
-------
In another study, Lachapelle (Reference 7-10) estimated costs for
operating under low excess air conditions. Generally, no significant
additional cost for modern units or units in good condition is required for
reducing excess air. However, some older units may require modifications
such as altering the windbox by adding division plates, separate dampers and
operators, fuel valving, air register operators, instrumentation for fuel
and airflow, and automatic combustion controls. Table 7-3 shows estimated
investment costs for LEA firing on existing utility boilers (Reference
7-10). These costs are guidelines which can vary depending on the
modifications that are required. As unit size increases, the cost per kW
decreases since the larger units typically have inherently greater
flexibility and may require less extensive modification.
TABLE 7-3. 1974 ESTIMATED INVESTMENT COSTS FOR LOW EXCESS AIR FIRING
ON EXISTING BOILERS NEEDING MODIFICATIONS (Reference 7-10)
Unit Size
(Electrical Output)
(MM)
1000
750
500
250
120
Investment Cost,
$/kW
Gas and Oil
0.12
(0.15)a
0.16
(0.19)
0.21
(0.25)
0.33
(0.40)
0.53
(0.64)
Coal
0.48
(0.58)a
0.51
(0.62)
0.55
(0.67)
0.64
(0.78)
0.73
(0.89)
a!977 dollars in parenthesis
7-5
-------
The use of low excess air firing reportedly increases boiler
efficiency by 0.5 to 5 percent. Additional savings may result from
decreased maintenance and operating costs, so any investment costs can be
offset by savings in fuel and operating expenses.
The best documented control costs to date have been those of Selker
and Blakeslee (References 7-11 and 7-12). Costs for the combined use of
overfire air ports and low excess air firing for both new and existing units
are summarized from these studies in Figures 7-1 and 7-2. Capital costs
were projected over a unit size range of 25 to 1000 MW. Figure 7-1 applies
to new unit designs with heating surfaces adjusted to compensate for the
resultant changes in heat transfer and rates. Figure 7-2 applies to
existing units with no change in heating surface, as these changes must be
calculated on an individual unit basis. Cost ranges for existing units vary
more widely than for new units, since variations in unit design and
construction can either hinder or aid the installation of a given NO
A
control system. It can be noted from Figures 7-1 and 7-2 that the average
(not the range of) modification cost on a per kilowatt basis is not a strong
function of equipment size. In other words, for the purposes of cost
estimation, there are no significant economies of scale since the "error
band" in the original estimate is so broad. This fact will be put to use in
the cost estimations of Section 7.3.
In addition to the increased capital costs for including OFA in new
or existing units, Selker and Blakeslee reported differential operating
costs for 500 MW new and existing boilers, as shown in Table 7-4 (Reference
7-12). To put these operating costs in perspective, they can be compared to
the percent increase in generating costs shown at the bottom of Table 7-4.
Except for the case of older units, the difference in operating cost is
below 0.1 percent of annual cost.
The results of Selker and Blakeslee, though valuable, are for a
particular control case only, overfire air for tangential coal-fired
boilers. The most recent cost estimates are those of Krippene for oil- and
gas-fired boilers (Reference 7-13). Table 7-5 gives the estimates for
investment costs and total annual cost. Unfortunately, the initial
investment cost of controls, including hardware requirements and costs, were
not documented.
7-6
-------
1.00 -
0.75
»- 0.50
V)
o
o
0.25
0.00
NEW UNITS INSTALLATION COSTS
4 WINDBOX FURNACES
8WINOBOX FURNACES
200
400 600 600
UNIT SIZE, MW
1000
Figure 7-1. 1975 capital cost of OFA on new tangential
coal-fired boilers (Reference 7-11).
1.50 -
1.25
5 LOO
V.
••»
,_• 0.75
V)
O
o
0.50
0.25
0.00
EXISTING UNITS MODIFICATION COSTS
4 WINDBOX FURNACES
6 WINDBOX FURNACES
200
400 600
UNIT SIZE, MW
800
1000
Figure 7-2. 1975 capital cost of OFA on existing coal-fired
boilers (Reference 7-11).
7-7
-------
TABLE 7-4. 1975 DIFFERENTIAL OPERATING COSTS OF OFA ON NEW AND EXISTING TANGENTIAL COAL-FIRED UTILITY
BOILERS (Reference 7-12) (Net Heat Rate 10 MJ/kWh, March 1975 Equipment Costs)*
-•J
I
CO
Capital Cost ($/kW)
Annual Capital Cost ($)
Annual Fuel Cost ($)
Labor and Maintenance ($)e
Total Annual Costf ($)
Electricity Cost (mills/kWh)9
Increase (X)
Increase (mills/kUh)f
New
Plant
Without
Overfire Air
500.00
40,000,000a
18,000,000C
8,100,000
66.100,000
24.481
~
~
New
Plant
With
Overfire Air
500.20
40,016,000
18,000,000
8,100,000
66,116,000
24.487
0.024
0.006
Recent
Existing
With Added
Overfire Air
500.70
40,056,000
18,000, 000d
8,100,000
66,156,000
24.502
0.086
0.021
Older
Existing
Without
Overfire Air
250.00
20,000,000b
9,000,000
8,100,000
37,100,000
13.741
—
~
Older
Existing
With Added
Overfire Air
250.70
20,056,000
9,000,000
8,100,000
37,156,000
13.762
0.153
0.021
aAnnual fixed charge rate of 16 % x 500 $/kW x 500,000 kW
b!6 X x 250 $/kW x 500,000 kW
C0.66 J/GJ coal cost x 5,400 hr/yr x 500,000 kW x 10 MJ/kWh
d0.33 $/GJ coal cost x 5,400 hr/yr x 500,000 kW x 10 MJ/kWh
eLabor and maintenance cost of 3.0 mills/kWh
f5,400 hr/yr at 500 MM — 2,700 GWh/yr
^Cost at plant bus bar; transmission and distribution not included
*To convert to 1977 dollars, multiply 1975 dollars by 1.2 factor
-------
TABLE 7-5. COSTS FOR NO.. EMISSION CONTROLS ON ELECTRIC POWERPLANTS USING GAS- AND OIL-FIRED
STEAM GENERATION EQUIPMENT (1977 DOLLARS) (Reference 7-13)
Investment ($/kW)
Fixed Capital Charges ($/kM-yr)
Operation and Maintenance Cost ($/kW-yr)
Fuel Cost Penalty ($AH-yr)
Annual Cost ($/kW-yr)
Existing Powerplants
Staged
Combustion9
0.4 to 2.0
0.08 to 0.4
0.02 to 0.1
0.31
0.41 to 0.81
Staged Combustion
Plus FG Recirculation
(Best Effort Basis)
7.0 to 11.0
1.4 to 2.2
0.35 to 0.55
1.24
2.99 to 3.99
New Powerplants
Staged
Combustion3
0.5 to 1.50C
0.1 to 0.3
0.025 to 0.075
0.15 to 0.31
0.275 to 0.685
State of the Art
NO Control
5.0 to 12.0
1.0 to 2.4
0.25 to 0.6
0.62 to 1.24
1.87 to 4.24
I
UD
aF1xed capital charges = 201, O&M costs = 5X, fuel penalty
= (0.12 - 0.25X) x S2.85/GJ x 5000 hrs/yr
Fixed capital charges * 20%, O&M costs = 5X, fuel penalty
= (0.5 - l.OX) x S2.85/GJ x 5000 hrs/yr
cTo meet current EPA NO emission requirements: I.e., 86 ng/J, gas; 129 ng/J, oil
Net plant heat rate = 9.71 MJ/kWh
-------
7.2 COST ANALYSIS PROCEDURES
Given the background of control cost estimation discussed in the
previous section, there is an evident need for a systematic, well documented,
up to date cost analysis of typical controls for representative boiler
design/fuel classifications. In this way, the cost effectiveness of
controls can be compared from boiler to boiler on an even basis.
Therefore, the use of accepted estimation procedures for costing
NO control implementation in current dollars was employed in this study,
A
with heavy reliance on discussions with boiler manufacturers, equipment
vendors, and utilities. For the case of retrofit control costs, preliminary
design work was performed to allow estimation of hardware and installation
needs, as well as engineering requirements. The analysis was applied to a
number of cases to give a range of retrofit control costs. For the cost of
NO controls in new boilers, the services of two major suppliers, the
A
Babcock & Wilcox Company and the Foster Wheeler Energy Corporation, were
enlisted. Their estimates are presented in Section 7.4.
For the analysis of the cost of controls, regulated public utility
economics were adopted. These are governed by the following principles
(Reference 7-14).
• Permitted revenue - (current operating disbursements +
depreciation + interest paid on debt) = taxable income
• Taxable income x effective tax rate = income taxes
• Permitted revenue = current operating disbursements +
depreciation + income taxes + (fair return x rate base)
Permitted revenue is often called revenue requirement by utilities. Thus,
the latter term is adopted in the following. Based on the revenue
requirement approach, an annualized cost methodology was developed, adapted
from that used by the Tennessee Valley Authority in evaluating the cost of
power plant projects for EPA (Reference 7-15) and EPRI (Reference 7-16).
This procedure has been generally accepted in the industry (References 7-17
through 7-19).
For the present application, the additional revenue requirement
represents the incremental cost of operating a boiler under controlled
conditions over and above the cost of operating the same boiler uncontrolled.
In other words, the revenue requirement takes into account the initial
investment, the annual capital charges resulting from that investment, and
7-10
-------
all direct operating costs such as operation and maintenance. Once the
revenue requirement RR(n) for each year n that the utility operates the
control up to N years (the remaining lifetime of the boiler) is obtained, an
annualized cost, or a discounted level annual cost can be evaluated. Using
basic economics (Reference 7-14):
Annualized Revenue Requirement =
30 *
J)N -
The first term of the product is the capital recovery factor, which
recognizes the time value of money by discounting at an annual cost of
capital of j x 100 percent (effective interest rate). The effective
interest rate j is given by:
j = bi + (1 - b)r
where b is the debt/equity ratio, i is the interest rate on this borrowed
money (debt), and r is the rate of return to equity. According to the
Edison Electric Institute (Reference 7-20), the debt/equity ratio for the
utility industry has been relatively constant over recent years and b = 0.5
is a good estimate. The interest rates i and r were taken as 0.08 and 0.12,
respectively (References 7-15 and 7-16).
With the annualized revenue requirement or annualized cost approach,
the details of calculating RR(n) will be presented. The revenue
requirements for each year n are given by the sum of direct operating costs
and indirect operating costs.
RR(n) = DOC + IOC(n)
where DOC is given by the sum of the following incremental costs:
0 Fuel penalty under controlled conditions
t Fuel credit (for unused fuel if forced to derate)
• Raw materials
7-11
-------
• Conversion costs
— Additional operating personnel
— Additional utilities requirements
— Additional maintenance
— Required analyses
• Annual royalties (if any)
• Purchased power (if forced to derate)
and IOC(n) is given by the sum of the following incremental costs:
• Capital charges
— Depreciation
— Insurance
— Replacement costs
— Cost of capital and taxes
t Capital charges of lost capacity (if forced to derate)
• Overhead
-- Administrative overhead
— Plant overhead
Indirect operating costs represent overhead as well as the capital charges
due to the initial investment and any lost capacity. This lost capacity
charge will be discussed later in this section. The initial investment is
given by the sum of the following costs:
• Engineering design and supervision
• Engineering fee
• Hardware requirements
• Installation labor and supervision
• Construction facilities
• Service facilities
• Utilities facilities
• Construction field expense
• Contractor's fee
t Construction contingency
• Initial charges (such as licensing fees, if any)
• Startup costs
The appropriate equations or estimation procedures for calculating all of
the above cost factors are presented in Table 7-6.
7-12
-------
TABLE 7-6. COST ANALYSIS CALCULATION ALGORITHM'
Cost Factor
Calculation Equation
Reference
I
»-•
co
Initial Investment, II
Engineering Design & Supervision, DS
Engineering Fee, EFEE
Hardware, TM
Installation Labor & Supervision, TL
Construction Facilities, CF
Service Facilities, SF
Utilities Facilities, UF
Construction Field Expense, CFE
Contractor's Fee, CON
Construction Contingency, CTN
Initial Charges, 1C
Startup Costs, SC
Indirect Operating Costs, IOC(n)
Capital Charge, CC(n)
Depreciation, D
Insurance, IN
Replacements, RE
Cost of Capital and Taxes, CCT(n)
II = (DI + IND + SC + 1C), as per below
DS estimated from preliminary design work
EFEE = 0.08 x DS
TM from preliminary design work
TL from preliminary design work and engineering estimate
CF = 0.05 x (TL + TM + UF + SF)
SF = 0.05 x (TL + TM)
UF = 0.03 x (TL + TM)
CFE * 0.13 x (TL + TM + CF + SF + UF) = 0.13 x DI
CON * 0.07 x DI
CTN = 0.11 x DI
1C from input data (e.g., licensing fees, usually none)
SC = 0.10 x (DI + DS + EFEE + CFE + CON + CTN)
- 0.10 x (DI + IND)
IOC(n) = CC(n) + CCLOST(n) + OH
CC(n) - D + IN + RE + CCT(n)
D - II/N
IN = 0.005 x II
RE - 0.004 x II
CCT(n). = [ib'+ r(l - b) + ^p-^-f (l ~ b>r ]
where t = effective tax rate
= s + (l-s)f
and s * state tax rate
f " federal tax rate
and ODB * II - (n-l)D
This report, Section 7.3
Engineering estimate
Vendor quotes
This report, Section 7.3
TVA (References 7-15, 7-16)
TVA (References 7-15, 7-16)
TVA (References 7-15, 7-16)
TVA (References 7-15, 7-16)
TVA (References 7-15, 7-16)
TVA (References 7-15, 7-16)
This report, Section 7.3
TVA (References 7-15, 7-16)
Straight line depreciation
TVA (References 7-15, 7-16)
This report, Section 7.2
aA glossary of cost analysis terms appears in Appendix E.
-------
TABLE 7-6. Concluded
Cost Factor *
Calculation Equation
Reference
Capital charges of Lost Capacity,
CCLOST (n)
Overhead
Administrative overhead, OHA
Plant overhead, OHP
Direct Operating Costs, DOC
Fuel Penalty, AF
Fuel Credit, FC
Raw materials, RH
Conversions Costs
Additional operating personnel, OLS
Additional utilities, UC
Additional maintenance, M
Required analyses, A
Annual royalties, AROY
Purchased Power, PP
Calculated analogously to CC(n), only use
no x 2BATE 1n place of n
where 110 = Initial investment of boiler
ORATE = Power derate with controls, if necessary
KW = Power rating of boiler before control
OHA = 0.10 x OLS
OHP = 0.20 x (OLS + UC + M + A), as indicated below
AF = HYR x HRATE x (KW - DRATE) x FCOST
x FPEN
where HYR = Annual operating hours
FC = HYR x HRATE x DRATE x FCOST
RM from input data
where HRATE = Heat rate of boiler
FCOST = Fuel cost
OLS from engineering estimate
UC from engineering estimate
M = 0.05 x (TL + TM)
A from engineering estimate
AROY from input data
PP = DRATE x HYR x PPR
where PPR = purchased power rate
This report. Section 7.2
TVA (References 7-15, 7-16)
TVA (References 7-15, 7-16)
Engineering estimate
Engineering estimate
Engineering estimate
Reference 7-4
TVA (References 7-15, 7-16)
Reference 7-4
Reference 7-4
Engineering estimate
Annualized Cost to Control, ARRU
ARRU =
DOC + IOC(n)
(KW - DRATE)
This report, Section 7.2
Engineering estimates are based on process analyses 1n Section 6, and design analyses in Section 7.3.
-------
The calculation equations of Table 7-6 are self-explanatory, but
perhaps a few comments are in order. The cost of capital and taxes per year
can be calculated as follows:
Tax Deductible
*
Taxable Income = (Return to Equity) + (Interest onx Borrowed Money)
/
/
Tax Deductible /
^
+ (Depreciation) + (Money for Taxes)
Now the total tax, T, is given by:
T = Federal + State Tax
= t x (Taxable Income)
where t = effective tax rate
= s + (1 - s)f
s = State tax rate
f = Federal tax rate
since State taxes are deductible from Federal taxes.
Combining the equations,
T = -j r (Return to Equity)
= y4-£ (1 - b)r . ODB(n)
where the outstanding depreciation base ODB(n) in year n is given by:
n
ODB(n) = II - ^ D(n)
n=l
= II - (n - 1)D
assuming straight line depreciation.
7-15
-------
Therefore, the cost of capital and taxes in year n is given by:
CCT(n) = fib + r(l - b) + -^~t (1 ' b)r| * ODB(n)
and should be annualized as:
(1 * J)N - 1
CCT(n)
Another point of note in the cost analysis is the accounting of lost
capacity if a utility boiler is forced to derate due to the controls
implemented and the utility cannot compensate elsewhere for the lost power.
For example, if a utility is forced to apply BOOS as a control technique on
a coal-fired boiler, the unit may have to be derated by as much as
20 percent. The cost of purchased power to make up their lost capacity,
less any savings from unused fuel (due to derating of the boiler), should be
charged to the cost of that control. Furthermore, the control technique
should be held accountable for a prorated portion of the capital charge of
the original boiler based on the fractional loss in boiler capacity, i.e.,
/Annual Capital Charge
\ of Lost Capacity
\ _ /Annual Capital Charge\ /
I ~ \ of Boiler / x \
Lost Capacity \
Orig. Total Capacity)
7.3 RETROFIT CONTROL COSTS
Representative costs for retrofitting and operating typical existing
boilers under NO control are presented in this section. Costs are given
^
in dollars per unit electrical output per operating year. As shown in
Section 7.1, on a per unit kW basis, average control costs are not a strong
function of unit size, but rather strongly dependent on the characteristics
of the particular unit in question. Still, for the purposes of this cost
analysis, typical unit sizes are chosen in Section 7.3.1. Appropriate
representative controls are also selected. Section 7.3.2 goes into the
7-16
-------
details of control equipment hardware and operating costs, while
Section 7.3.3 gives the results of annualizing the cost to control.
7.3.1 Selection of Representative Boilers
The three major utility boiler firing designs (tangential, single
wall, and horizontally opposed wall firing) and the three primary fuels
(coal, oil, and natural gas), give nine basic boiler/fuel classifications.
Of course, many units are designed to burn more than one fuel. This is
particularly true for gas and oil fuels.
The Environmental Protection Agency's Energy Data System
(Reference 7-21) was used to obtain relative installed population size
distributions of the nine boiler/fuel classifications. Using this system, a
typical unit size was determined for each category. Further comparisons
were then made to determine which cases would be examined in more detail in
retrofit design studies.
The cases selected for further study were a tangential coal-fired
unit to power a 225 MW turbine generator, a 540 MW horizontally opposed
coal-fired unit, and a 90 MW front wall gas- and oil-fired unit. Primary
considerations in making these selections included:
• The trend toward coal firing, particularly in larger size units,
emphasizes tangential and horizontally opposed firing designs
• Many units are capable of burning oil and gas, especially in the
smaller size ranges. Single wall (front or rear) fired units are
common in this application
The methods used to control NO emissions are fuel dependent. In
n
retrofit applications, the number of control methods available may be
limited. The discussion of control techniques in Sections 4 through 6
addressed the limitations of the various control methods. It was noted that
the retrofit application of NO controls can impose serious operating
J\.
problems on the user in that the control methods often cause a significant
departure from the original design operating characteristics of the unit
(References 7-8 and 7-23). Noting that control needs, and therefore,
control costs are more dependent on the fuel fired than on the boiler
7-17
-------
equipment type, the following typical boilers and controls were chosen for
detailed analysis:
Boiler/Fuel Type NCL Control
^•«»^^_^_
Tangential/Coal OFA
Opposed Wall/Coal OFA
Opposed Wall/Coal Low NO Burners
A
Opposed Wall/Coal BOOS
Single Wall/Oil and Gas BOOS
Single Wall/Oil and Gas OFA and FGR
Overfire air and low NO burners were selected as the retrofit
A
control methods for coal firing. Burners out of service is not necessarily
recommended for coal-fired units, but is included to demonstrate the
prohibitively high cost of derating a unit, as is often the case for
pulverized coal units. Burners out of service, and flue gas recirculation
through the burners combined with overfire air were selected as the retrofit
control methods for the single wall oil- and gas-fired unit. These methods
have been shown to be effective in retrofit applications, as discussed
earlier in this report.
7.3.2 Retrofit Design Analysis
In the retrofit design analysis, three representative units were
selected, retrofit NOX controls were chosen, and estimates of labor,
materials, and equipment required to install the NO control equipment
A
were performed.
Using the aforementioned Energy Data System (Reference 7-21), a
listing for each fuel and firing type of boiler was obtained. A weighted
average of unit capacity was then used to select a typical unit capacity.
This size and firing arrangement was then approximated and used as the basis
for the design study. It should be noted that in the design of a boiler,
there are many variables to be considered. There are no "standard" utility
boilers.
7.3.2.1 Tangential Coal
A 225 MW unit was selected as a representative tangential pulverized
coal-fired unit with overfire air selected as the retrofit NO control
A
technique. The model used for the study was a single furnace design with
7-18
-------
five levels of burners or fuel admission nozzles, one set per corner. With
a plant cycle efficiency of 37.1 percent (9.71 MJ/kWh or 9200 Btu/kWh heat
rate), a heat input rate per nozzle of 30 MW (1.02 x 108 Btu/hr) would be
required. The design excess air at full load was assumed to be 20 percent.
The overfire air ports were designed to handle a maximum of 20 percent of
the total airflow.
It was assumed that there were no major obstructions in routing the
ductwork from the hot combustion air (secondary air) duct to the overfire
air ports. It was also assumed that there were no major problems with
access to the work areas. It should be noted that in retrofit
installations, problems with obstructions and access are frequently
encountered. These problems can increase installation and material costs
significantly.
The cost estimates were based on vendor quotes and engineering
estimates. Installation costs were based on the Richardson Rapid Method
(Reference 7-22) and assumed an average labor rate for a composite crew of
$15.30/hr. These estimates are listed in Tables 7-7 through 7-9. It will
be noted that in these and subsequent analogous tables, numbers presented
have not been rounded off in order to minimize errors in the cost code
calculations. Obviously, they should be taken to only two significant
figures. The final control costs presented in Section 7.3.3 have been
rounded to two significant figures. Drawings of the tangential coal model
for the design study are shown in Figures 7-3 through 7-5. A discussion of
annualized retrofit control costs including investment and operating costs
is given in Section 7.3.3.
7.3.2.2 Opposed Mali Coal
A 540 MW opposed wall coal-fired boiler was selected as a
representative unit. Overfire air was selected as the retrofit N0¥
A
control technique as in the tangential coal design. The model used for the
study was a single furnace design with 48 burners, 24 on the front wall and
24 on the rear wall. The burners were arranged in four horizontal rows of
six burners each. With a plant cycle efficiency of 37.1 percent
(9.71 MJ/kWh or 9200 Btu/kWh heat rate), a heat input rate per burner of
30.4 MW (1.03 x 108 Btu/hr) would be required. The overfire air ports
were designed to handle 20 percent of the total airflow. The excess air
level at full load was assumed to be 15 percent.
7-19
-------
TABLE 7-7. COMPONENT COST ESTIMATE: RETROFIT OFA FOR TANGENTIAL
COAL-FIRED BOILER (1977 DOLLARS)
Component Description and Quantity Required
Expansion Joints, 38" x 72" x 12"
Eight Required
Control Dampers, 38" x 72"
Four Required
Tilting Air Nozzles, 28" x 16"
Eight Required
Hot Air Duct 38" x 72" x 120"
Four Required
Materials (Tubes, Fittings and Supports)
Total Component Cost
Type of
Quote
WQa
WQb
WQb
WQb
EE
Amount ($)
13,312
5,060
6,760
5,760
1,339
31,903
aTate-Reynolds Co., Inc.
bKanawha Manufacturing Co.
WQ — Written quote
EE — Engineering estimate
TABLE 7-8. INSTALLATION COST ESTIMATE: RETROFIT OFA FOR
TANGENTIAL COAL-FIRED BOILER (1977 DOLLARS)
Component Installed
Overfire Air Ports
Ducts, Expansion Joints,
and Dampers
Total Installation Estimate
Estimated Hours
2965
565
3530
Cost Estimate @
45,362
8,647
$15.30/hr
54,009
7-20
-------
TABLE 7-9. RETROFIT OFA FOR TANGENTIAL COAL-FIRED BOILER
(1977 DOLLARS)
I. Design Estimate
Estimated
Category Hours
1. Designer @ $9/hr 600
2. Engineer $ $12/hr 240
3. Supervision @ 10% of 1 and 2
4. Overhead @ 110% of 1, 2, and 3
5. General and Administrative
9 25% of 1, 2, 3, and 4
6. Fee @ 8% of 1, 2, 3, 4, and 5
II. Construction Estimate
Estimated
Category Hours
1. Labor 9 $15.30/hr (Table 7-8) 3530
2. Supervision @ 10% of 1
3. General and Administrative
0 25% of 1 and 2
4. Construction Field Expense
5. Contractor's Fee
6. Construction Contingency
7. Construction Facilities
8. Service Facilities
9. Utilities Facilities
III. Component Cost Estimate (Table 7-7)
Subtotal (I + II + III)
Startup Costs 9 10% of I, II, & III
Total Initial Investment
Estimated
Costs
5,400
2,880
828
10,019
4,782
1.913
25,822
Estimated
Cost
54,009
5,401
14,852
15,651
8,427
13,243
5,733
5,308
3,185
125,809
31,903
183,534
18,353
201,887
7-21
-------
DIMENSIONS IN METERS (FEET)
Figure 7-3. Retrofit overfire air for typical tangential
coal-fired boilers.
7-22
-------
DIMENSIONS IN METERS (FEET)
—
5j ^*^
O.H (l)
^vf
THV .< U J
1
. •**
" r
L
/-*|
I
rl
1
f
I
Figure 7-4. Typical overfire air port arrangement for
tangential coal-fired boilers.
7-23
-------
-
-1
PFTAlL P
4.4 « i u>
(Hv - i — 1
T.Ut K *7 L
( * « ^ X '
- — <"•<•=•
(Vll)
-------
Twelve overfire air ports were added, one port above each vertical
row of burners. It was assumed that there were no major obstructions in
routing the necessary ductwork. The ducts were attached as simple
extensions to the front and rear windboxes. It was also assumed that there
would be no major problems with access to the work areas. It should be
noted that in retrofit installations, access and obstruction problems are
frequently encountered. These factors may increase installation and
material costs significantly.
Component and installation costs were estimated as described in
Section 7.3.2.1. These estimates are listed in Tables 7-10 through 7-12.
Drawings of the opposed wall coal-fired model boiler for the design study
are shown in Figure 7-6. A discussion of annualized retrofit control costs
including investment and operating costs is given in Section 7.3.3.
The retrofit installation of low NO burners on this opposed wall
A
coal-fired unit was also considered. It was assumed that low NO burners
X
could be installed in place of the existing burners with no modifications to
burner openings in the furnace walls. It was also assumed that existing
coal conveying equipment, flame safeguard equipment, burner register drives,
and igniting equipment could be utilized. The assumptions of good access
and few obstructions cannot be justified here. To remove and replace the
burners, the burner front piping and portions of the windbox would have to
be removed for access. Considering these factors, the cost estimates for
the retrofit installation of low NO burners is shown in Tables 7-13
A
through 7-15.
7.3.2.3 Single wall Oil and Gas
A single wall unit designed to fire both oil and gas was selected
because many units are capable of burning either fuel. Also, the same
retrofit NO control techniques are effective with either fuel. A
A
representative unit size of 90 MW was chosen. The model used for the study
was a single furnace design with six burners on the front wall arranged in
two rows of three burners each. With a plant cycle efficiency of 37 percent
(9.71 MJ/kWh or 9200 Btu/kWh), a heat input rate per burner of approximately
40.5 MW (1.38 x 108 Btu/hr) would be required.
The retrofit NO control cases chosen were (1) burners out of
A
service, and (2) flue gas recirculation through the burners combined with
overfire air. These control methods are effective for both gas- and
7-25
-------
TABLE 7-10. COMPONENT COST ESTIMATE: RETROFIT OFA FOR OPPOSED
WALL COAL-FIRED BOILER (1977 DOLLARS)
Component Description and Quantity Required
Expansion Joints, 36" Diameter
Twelve Required
Segmented Elbows, 36" Diameter
Twelve Required
Round to Square Transitions
Twelve Required
Control Dampers, 48" x 48"
Twelve Required
Materials (Tubes, Fittings, Supports)
Type of
Quote
WQa
WQb
WQb
WQb
EE
Amount ($)
21,156
14,340
13,296
15,348
1,933
Total Component Cost 66,073
aTate-Reynolds Co., Inc.
Kanawha Manufacturing Co.
WQ — Written Quote
EE — Engineering Estimate
TABLE 7-11. INSTALLATION COST ESTIMATE: RETROFIT OFA FOR OPPOSED
WALL COAL-FIRED BOILER (1977 DOLLARS)
Component Installed
Over fire Air Ports
Ducts, Expansion Joints,
and Dampers
Total Installation Estimate
Estimated Hours
3523
2090
5613
Cost Estimate 9
$15.30/hr
53,898
31,975
85,873
7-26
-------
TABLE 7-12. INITIAL INVESTMENT ESTIMATE: RETROFIT OFA FOR OPPOSED
WALL COAL-FIRED BOILER (1977 DOLLARS)
I . Design Estimate
Estimated
Category Hours
1. Designer @ $9/hr 700
2. Engineer @ $12/hr 300
3. Supervision @ 10% of 1 and 2
4. Overhead @ 110% of 1, 2, and 3
5. General and Administrative
9 25% of 1, 2, 3, and 4
6. Fee § 8% of 1, 2, 3, 4, and 5
II. Construction Estimate
Estimated
Category Hours
1. Labor @ $15.30/hr (Table 7-11) 5613
2. Supervision § 10% of 1
3. General and Administrative
@ 25% of 1 and 2
4. Construction Field Expense
5. Contractor's Fee
6. Construction Contingency
7. Construction Facilities
8. Service Facilities
9. Utilities Facilities
III. Component Cost Estimate (Table 7-10)
Subtotal (I + II + III)
Startup Costs @ 10% of I, II, & III
Total Initial Investment
Estimated
Cost
6,300
3,600
990
11,979
5,717
2,287
30,873
Estimated
Cost
85,879
8,588
23,617
27,148
14,618
22,972
9,944
9,208
5,525
207,499
66,073
304,445
30,444
334,889
7-27
-------
la.tag.cfr
i
DIMENSIONS IN METERS (FEET)
Figure 7-6. Retrofit overfire air for typical opposed wall coal-fired boilers.
-------
TABLE 7-13. COMPONENT ESTIMATE: RETROFIT LOW NO BURNERS FOR OPPOSED
WALL COAL-FIRED BOILER(1977 DOLLARS)X
Component Description and Quantity Required
Low NO Burner, Complete
48 Reqfiired
Type of
Quote
EE
Amount ($)
336,000
EE — Engineering estimate based on discussion with equipment
manufacturer.
TABLE 7-14. INSTALLATION- COST ESTIMATE: RETROFIT LOW NO BURNERS
FOR OPPOSED WALL COAL-FIRED BOILER (1977 DOLLARS)
Component Installed
Low NO Burners
J\
Estimated Hours
15,360
Cost Estimate @ $15.30/hr
235,008
7-29
-------
TABLE 7-15. INITIAL INVESTMENT ESTIMATE: RETROFIT LOW NOX BURNERS FOR
OPPOSED WALL COAL-FIRED BOILER (1977 DOLLARS)
I. Design Estimate
Category
1. Designer @ $9/hr
2. Engineer @ $12/hr
3. Supervision @ 10% of 1 and 2
4. Overhead @ 11056 of 1, 2, and 3
5. General and Administrative
@ 25% of 1, 2, 3, and 4
6. Fee 9 8% of 1, 2, 3, 4, and 5
II. Construction Estimate
Category
1. Labor @ $15.30/hr (Table 7-14)
2. Supervision @ 10% of 1
3. General and Administrative
G> 25% of 1 and 2
4. Construction Field Expense
5. Contractor's Fee
6. Construction Contingency
7. Construction Facilities
8. Service Facilities
9. Utilities Facilities
III. Component Cost Estimate (Table 7-13)
Subtotal (I + II + III)
Startup Costs @ 10% of I, II, & III
Total Initial Investment
Estimated Estimated
Hours Cost
400 3,600
150 1,800
540
6,534
3,119
1,247
16,840
Estimated Estimated
Hours Cost
15,360 235,008
23,501
64,627
97,170
52,322
82,221
35,593
32,957
19,774
643,173
336,000
996,013
99,601
1,095,614
7-30
-------
oil-fired units. The overfire air ports were designed for 25 percent of the
total air and recirculated gas flow. The flue gas recirculation system was
designed to handle 25 percent of the flue gas normally produced.
Recirculating this flue gas into the burner windbox results in a minimum
windbox 02 level of approximately 17 percent.
Three overfire air ports were added, one port above each vertical row
of burners. Again, it was assumed that there were no major obstructions in
routing the necessary ductwork. The ducts were attached as simple
extensions to the windbox. The flue gas recirculation system was similarly
added. It was further assumed that there was reasonable access to the work
areas. As noted in Section 7.3.2.1 and 7.3.2.2, access and obstruction
problems are frequently encountered and have the effect of increasing
installation and material costs significantly. Other assumptions were that
there was adequate forced draft fan capacity and that ductwork and furnace
strength were adequate with the addition of gas recirculation.
Component and installation cost estimates using the methods described
in Section 7.3.2.1 are listed in Tables 7-16 through 7-18. Drawings of the
single wall unit used in the study are shown in Figure 7-7. A discussion of
annualized retrofit control costs including investment and operating costs
is given in Section 7.3.3.
7.3.3 Annualized Retrofit Control Costs
Based on the retrofit control design analysis of Sections 7.3.2, and
the assumptions made in the cost analysis algorithm of Section 7.2, typical
retrofit control costs were generated. The results based on 1977 dollars
are given in detail in Tables 7-19 through 7-24. It is assumed here that
low excess air represents standard operating procedure. As discussed in
Section 7.1, any investment costs for this control are usually offset by
savings in operating efficiency.
It was assumed that all retrofit installations could be completed
during normal outage periods, and hence downtime need not be costed. As
shown in Section 7.3.2, this assumption is a good one (installation time 6
weeks or less) for all the retrofit cases considered with the exception of
low NO burner installation. For low NOV burner retrofit, which is
« A
estimated to require 12 weeks, installation will have to be scheduled during
a major overhaul of the boiler.
7-31
-------
TABLE 7-16. COMPONENT COST ESTIMATE: RETROFIT OFA AND FGR FOR
TYPICAL SINGLE WALL OIL- AND GAS-FIRED BOILER (1977 DOLLARS)
Component Description and Quantity Required
Flue Gas Recirculating Fan, Housing, Motor,
Turning Gear, Switchgear, Inlet Damper
Controls and Instrumentation
Expansion Joints, FGR Ducts
Five Required
Dampers, FGR Ducts
Three Required
Segmented Elbows, 36" Diameter, OFA
Three Required
Expansion Joints, 36" Diameter, OFA
Three Required
Round to Square Transitions, OFA
Three Required
Control Dampers, 48" x 48", OFA
Three Required
Materials (Tubes, Fittings, Supports,
Concrete, Reinforcing)
Ductwork
Type of
Quote
VQa
VQb
EE
EE
WQC
WQd
WQC
WQ
EE
EE
Amount ($)
110,000
14,495
8,320
3,837
3,585
5,289
3,324
3,837
4,165
38,002
Total Component Cost 194,854
aWestinghouse, Sturtevant Div.
bBailey Controls Co.
cKanawha Mfg. Co.
dTate Reynolds Co. Inc.
VQ — _Verbal Quote
WQ -- Written Quote
EE — Engineering Estimate
7-32
-------
TABLE 7-17. INSTALLATION COST ESTIMATE: RETROFIT OFA AND FGR FOR TYPICAL
SINGLE WALL OIL- AND GAS-FIRED BOILER (1977 DOLLARS)
Component Installed
Overfire Air Ports
OFA Ducts, Expansion
Joints, and Dampers
FGR Fan Foundation
FGR Fan and Motor
FGR Ductwork, Dampers,
Expansion Joints
Crane Rental
Total Installation Estimate
Estimated Hours
1166
934
45
304
1863
—
4612
Cost Estimate 9 $15.30/hr
17,835
14,285
5,280
4,651
28,502
2,500
73,053
7-33
-------
TABLE 7-18. INITIAL INVESTMENT ESTIMATE: RETROFIT OFA AND FGR FOR TYPICAL
SINGLE WALL OIL- AND GAS-FIRED BOILER (1977 DOLLARS)
I. Design Estimate
Estimated
Category Hours
1. Designer (3 $9/hr 800
2. Engineer @ $12/hr 290
3. Supervsion @ 10% of 1 and 2
4. Overhead @ 110% of 1, 2, and 3
5. General and Administrative
(a 25% of 1, 2, 3, and 4
6. Fee (P 8% of 1, 2, 3, 4, and 5
II. Construction Estimate
Estimated
Category Hours
1. Labor @ $15.30/hr (Table 7-17) 4,612
2. Supervision @ 10% of 1
3. General and Administrative
@ 25% of 1 and 2
4. Construction Field Expense
5. Contractor's Fee
6. Construction Contingency
7. Construction Facilities
8. Service Facilities
9. Utilities Facilities
III. Component Cost Estimate (Table 7-16)
Subtotal (I + II + III)
Startup Costs ? 10% of I, II, & III
Total Initial Investment
Estimated
Cost
7,200
3,480
1,068
12,923
6,168
2,467
33,306
Estimated
Cost
70,564
7,056
19,405
43,028
23,169
36,408
15,761
14,594
8,756
238,741
194,854
466,901
46,690
513,591
7-34
-------
ftml
DIMENSIONS IN METERS (FEET)
Figure 7-7. Retrofit OFA and FGR for typical single wall oil- and
gas-fired boilers.
7-35
-------
TABLE 7-19. RETROFIT CONTROL COST: OVERFIRE AIR FOR EXISTING
TANGENTIAL COAL-FIRED BOILER (1977 DOLLARS)
MAXIMUM CONTINUOUS RATING (MW) : 225.
TYPICAL BASELINE NOX EMISSION IPPM «T 3* 02) : .
TYPICAL CONTROLLED NOX EMISSION (PPM AT 3* O2) I 310.
DERATE REQUIRED (KW) I NONE
FUEL PENALTY (PERCENTI ! .00
ANNUALIZCO LOST CAPACITY CAPITAL CHARGE (S/KW-YRI : NONF
ANNUALIZCO PURCHASED POWER PENALTY IS/KW-YHI : MONE
INITIAL INVESTMENT (S/KWI i .90
ANNUALIZED JKOlREcT OPERATING COST «*/KU-YR> : .21
ANNUAL1ZEO DIRECT OPERATING COST (t/KU-TR) I .32
ANNUALIZED COST TO CONTROL ll/KW-YP) « .53
INITIAL INVESTMENT (1)
ENGINEERING DESIGN 1 SUPERVISION
ENGINEERING FEE
HARDWARE
INSTALLATION LABOR S SUPERVISION
CONSTRUCTION FACILITIES
SERVICE FACILITIES
UTILITIES FACILITIES
CONSTRUCTION FIELD EXPENSE
CONTRACTORS FEE
CONSTRUCTION CONTINGENCY
INITIAL CHARGES
STARTUP COSTS
TOTAL INITIAL INVESTMENT
23908.
1913.
31903.
74262.
5733.
5508.
M«S.
15651.
8427.
132«3.
0.
18J53.
201887.
A'JNUALIZEO OPERATING rOST <»/YR)
INDIRECT OPERATING COSTS
CAPITAL CHARGES
DEPRECIATION «OTS.
INSURANCE 1009.
REPLACEMENT COSTS 808.
COST OF CAPITAL » TAXES 23055.
CAPITAL CHARGES OF LOST CAPACITY IIF DERATE)
DEPRECIATION n.
INSURANCE 0.
REPLACEMENT COSTS 0.
COST OF CAPITAL 1 TAXES 0*
OVERHEAD
ADMINISTRATIVE OVERHEAD 0.
PLANT OVERHEAD 14U32.
DIRECT OPERATING COSTS
FUEL COST PENALTY 0.
FUEL CREDIT IFOR UNUSED FUEL IF DERATE) I 0.)
RAW MATERIALS 0.
CONVERSION COSTS
ADDITIONAL OPERATING PERSONNEL 0.
ADDITIONAL UTILITIES REQUIREMENTS 66850.
ADDITIONAL MAINTENANCE 5308.
RE8UIREO ANALYSES 0.
ANNUAL ROYALTIES 0.
PURCHASED POWER IIF DERATE* o.
TOTAL ANNUALIZED OPERATING COSTS 119538.
AMNUALIZEO COST TO CONTROL (S/Kw-YR) .S>J
7-36
-------
TABLE 7-20. RETROFIT CONTROL COST: OVERFIRE AIR FROM EXISTING OPPOSED
WALL COAL-FIRED BOILER (1977 DOLLARS)
MAXIMUM CONTINUOUS RATING : 5to.
TYPICAL BASEL P-E NOX EMISSION Iff* AT S» O2 > !
TYPICAL CONTROLLED NOX EMISSION (PJ-M AT s» oz> :
DERATE REOUIREO (MW) I NONE
FUEL PENALTY (PERCENT) : .28
ANNUALIZEO LOST CAPACITY CAPITAL CHARGE <»/KU-YR)
ANNUAL i ZED PURCHASED POWEH PENALTY (»/KW-YR> :
INITIAL INVESTMENT <$/KWI | .62
ANNuALizES INDIRECT OPERATING COST IS/KW-YRI :
ANNUALIZEO DIRECT OPERATING COST ll/KW-YRI :
ANNUALIZED COST TO CONTROL (S/Ku-YR) : .69
7KS.
sso.
NONE
.if.
.52
INITIAL INVESTMENT ($)
ENGINEERING DESIGN £ SUPERVISIOr
FEE
HARDWARE
INSTALLATION LABOR & SUPERVISION
CONSTRUCTION FACILITIES
SERVICE FACILITIES
UTILITIES FACILITIES
CONSTRUCTIOM FIELD EXPENSE
CONTRACTORS FEE
CONSTRUCTION CONTINGENCY
INITIAL CHAPGES
STARTUP COSTS
TOTAL INITIAL INVESTKENT
ANNUALIZED OPERATING COST IS/YR)
INDIRECT OPERATING COSTS
CAPITAL CHARGES
66073.
9JOS.
5525.
1*616.
0.
50UUU.
INSURANCE
HEPLACfMFNT COSTS
COST OF CAPITAL » TAXES
CAPITAL CHARGES OF LOST CAPACITY IIP DERATE)
DEPRECIATION
REPLACEMENT COSTS
COST OF CAPITAL * TAXES
OVERHEAD
ADMINISTRATIVE OVERHEAD
PLANT OVERHEAD
DIRECT OPERATING COSTS
FUEL COST PENALTY
FUEL CREOIT (FOH UNUSED FUEL IF DERATE I
«AW MATERIALS
CONVERSIOK COSTS
ADDITIONAL OPERATING PERSnf.'NrL
ADDITIOMtL UTILITIES REOUlNEhENTS
ADDITIONAL MAINTENANCE
REOUIRED ANALYSES
ANNUAL ROYALTIES
PURCHASED POWER (IF DERATE)
TOTAL ANNUALIZEO OPERATING COSTS
ANNUALIZED COST TO CONTROL IS/KU-YR)
1SHO.
Q.
3SS72.
1150??.
0.)
O-
0.
o.
0'
o.
ST1505.
.69
7-37
-------
TABLE 7-21. RETROFIT CONTROL COST: LOW NOX BURNERS FOR EXISTING
OPPOSED WALL COAL-FIRED BOILER (1977 DOLLARS)
MAXIMUM CONTINUOUS RATING |MW> : bUO.
TTPICAl BASELINE NOX EMISSION
0.
0-
21A350.
.10
7-38
-------
TABLE 7-22. RETROFIT CONTROL COST: BURNERS OUT OF SERVICE FOR EXISTING
OPPOSED WALL COAL-FIRED BOILER (1977 DOLLARS)3
MAXIMUM CONTINUOUS RATING :
ANNUALIZED DIRECT OPERATING CnST (t/KW-YRI :
ANNUALIZED COST TO CONTROL t*/*w-YR) : SO.l?
7*5.
510.
: 5.33
5.34
INITIAL INVESTMENT It)
ENGINEERING DESIGN * SUPERVISION
ENGINEERING FEE
HARDWARE
INSTALLATION LABOR I SUPERVISION
CONSTRUCT^' FACILITIES
SERVICE FACILITIES
UTILITIES FACILITIES
CONSTRUCTION FIELD EXPENSE
CONTRACTORS FEE
CONSTRUCTION CONTINGENCY
INITIAL CHARGES
STARTUP COSTS
TOTAL INITIAL INVESTMENT
ANNUALIZED OPERATING COST (*/YH)
INDIRECT OPERATING COSTS
CAPITAL CHARGES
DEPRECIATION
INSURANCE
REPLACEMENT COSTS
COST OF CAPITAL S TAXES
CAPITAL CHARGES OF LOST CAPACITY IIF DERATEi
DEPRECIATION
INSURArCF
REPLACEMENT COSTS
COST OF CAPITAL 1 TAXES
OVERHEAD
ADMINISTRATIVE OVERHEAD
PLANT OVERHEAD
DIRECT OPERATING COSTS
FUEL COST PENALTY
FUEL CREDIT IFOR UNUSED FUEL IF DERATE)
RAW MATERIALS
CONVERSION COSTS
ADDITIONAL OPERATING PERSONNEL
ADDITIONAL UTILITIES REOUTREMCNTS
ADDITIONAL MAINTENANCE
REQUIRED ANALYSES
ANNUAL ROYALTIES
PURCHASED POWER CIF DERATE)
TOTAL ANNUALJZfD OPERATING COSTS
ANNUALIZEO COST TO CONTROL (*/Kw-YR>
7-4U.
KOI.
tso.
S6139.
ins.
»127,
558000-
85700.
1593065.
0.
50.
9001S.
90*1760.
0.
0>
0.
J52.
0.
0.
19fc5f-000.
1S01?5«».
)0*12
^Assumes a 20 percent derate, which is typical when applying
BOOS on a coal-fired utility boiler.
7-39
-------
TABLE 7-23. RETROFIT CONTROL COST: BURNERS OUT OF SERVICE FOR EXISTING
SINGLE WALL OIL- AND GAS- FIRED BOILER (1977 DOLLARS)
MAXIMUM CONTINUOUS RATIN6 (MW| ; 90.
TYPICAL PASELINE HOX EMISSION (PPM AT s* 0!> i • 355 o11/470 gas
TYPICAL CONTROLLED NOX EMISSION (PPM AT S» 0?) : 210 011/235 gas
DERATE proi'lorn IMU) :
FUEL PENALTY IPERCENTI i .?*
AWNUALIJEn LOST CAPACITY CAPITA) CHARGE (S/KU-YSI : fc
ANNUALIZED PURCHASED POWER PENfll.TY (f/Kg-YK| : MOME
INITIAL INVESTMENT (»/KKI • .30
ANNUALT7ED Ik:PIRECT OPFHATTNR COST (%/KU-YRI t .05
ANNIJALI2ED PTRECT PPE"»TINR CnST It/KU-TH) : .44
COST Tn COMT«OL H/KU.TI') : .49
INITIAL INVESTMENT <»>
ENGINEERING DESIGN a SUPERVISION
ENGINEERING FEE
HARDWARE
INSTALLATION LABOR I SUPERVISION
CONSTRUCTION FACILITIES
SERVICE FACILITIES
UTILITIES FACILITIES
CONSTRUCTION FIELD EXPENSE
CONTRACTORS FEE
CONSTRUCTION CONTINGENCY
INITIAL CHARGES
STARTUP COSTS
TOTAL INITIAL INVESTMENT
18105.
o.
3S66.
IPZ.
166.
101.
11)6.
?Tnn<».
ANHUALIZEO OPERATING TOST (f/YR)
INDIRECT OPERATING COSTS
CAPITAL CHARGES
DEPRECIATION
INSURANCE
REPLACEMENT COSTS
COST OF CAPITAL a TAXES
CAPITAL CHARGES OF LOST CAPACITY (ir DERATE)
DEPRECIATION
INSURANCE
REPLACEMENT COSTS
COST OF CAPITAL I TAXES
OVERHEAD
ADMINISTRATIVE OVERHEAD
PLANT OVERHEAD
DIRECT OPERATING COSTS
FUEL COST PENALTY
FUEL CREDIT (FOR UNUSED FUEL IF DEBATE)
RAW MATERIALS
CONVERSION COSTS
ADDITIONAL OPERATING PERSONNEL
ADDITIONAL UTILITIES REQUIREMENTS
ADDITIONAL MAINTENANCE
RE9UI«€0 ANALYSES
ANNUAL ROYALTIES
PURCHASED POWER (IF DERATE)
TOTAL ANNUALIZED OPERATING COST*
ANNUALIZEO CPST TO CONTROL (S/KU-TPI
10PO.
us.
106.
JOS'*.
0.
30.
3«12S.
0.)
0.
0.
0.
168.
0.
0.
0.
UJT33.
7-40
-------
TABLE 7-24. RETROFIT CONTROL COST: FLUE GAS RECIRCULATION AND OVERFIRE
AIR FOR EXISTING SINGLE WALL OIL- AND GAS-FIRED BOILER
(1977 DOLLARS)
MAXIMUM CONTINUOUS RATING i 90.
TYPICAL BASCLINt NOX EMISSION (PPM AT S* 021 : 355 011/470 qaS
TYPICAL CONTROLLED NOX EMsSION IPPM AT i* O2) | 155 oil/115 gas
DERATE REQUIRED (My) : NONE
FUEL PfNALTt (PERCENT) I .50
ANNUAL I ZED LOST CAPACITY CAPITAL CHARGE (1/KU-Ya) : NONE
ANNUALIZEO PURCHASED POWER PENALTY «»/KW-YR> : HONE
INITIAL INVESTMENT (S/KWI : 5.71
ANNUALIZED INDIRECT OPERATING COST <»/KU-YM> t 1.14
ANNUAUI2CD DIRECT OPERATING COST (»/KW-TNI : 1.91
AHNUALIZED COST TO CONTROL (s/Ku-YR) : 3.05
INITIAL INVESTMENT <»>
ENGINEERING DESIGN 1 SUPERVISION
ENGINEERING FEE
HARDWARE
INSTALLATION LABOR a SUPERVISION
CONSTRUCT^! FACILITIES
SERVICE FACILITIES
UTILITIES FACILITIES
CONSTRUCTION FIELD EXPENSE
CONTRACTORS FEE
CONSTRUCTION CONTIK'GENCY
INITIAL CHARGES
STARTUP COSTS
TOTAL INITIAL INVESTMENT
30638.
19UK50-
9TD2S.
1ST61.
1S02*.
2M69.
36108.
0.
16690.
51S587.
ANNUALIZED OPERATING COST (S/YR)
INDIRECT OPERATING COSTS
CAPITAL CHARGES
DEPRECIATION
INSURANCE
REPLACFMfNT COSTS
COST 0* CAPITAL I TAXES
CAPITAL CHARGES OF LOST CAPACITY IIF DERATE)
DEPRECIATION
INSURANCE
REPLACEMENT COSTS
COST OF CAPITAL * TAXES
OVERHEAr
ADMIMST°»TIVE OVERHEAD
PLANT OVERHEAD
DIRECT OPERATING COSTS
FUEL COST PENALTY
FUEL CRFDlT (FOR UNUSED FUEL IF DERATE)
RAW MATERIALS
CONVERSION COSTS
ADDITIONAL OPERATING PERSnHNfL
ADDITIONAL UTILT.TICS-RCOUTRCFENTS
ADDlTlOK'AL MAINTENANCE
REQUIRED ANALTSES
ANNUAL ROYALTIES
PURCHASED POWER (IF DERATE I
TOTAL ANNUALIZEO OPERATING COST*
2666.
?P5»-
ilfcil.
0.
1A669.
ANNUALIZEO COST TO CONTROL
5.05
7-41
-------
All cost input data and assumptions are listed in Appendix E. For
control cost projection purposes, the results shown should be considered
valid to only two significant figures. Obviously, the cost code input
figures of Appendix E and the intermediate results in Tables 7-19 through
7-24 were not rounded off in the computer code to minimize errors in the
calculations. It will be noted that the final figure, $/kW-yr, for each
control case has been rounded to the two significant figure accuracy.
It should be reiterated that the results presented are only
representative typical retrofit control costs. They represent retrofitting
relatively new boilers, say 5 to 10 years old, with at least 25 years of
service remaining. In any event, these relatively new boilers would likely
be the first to be controlled under any proposed retrofit emissions
regulations for existing boilers.
Although the control hardware, engineering and installation costs
for the retrofit cases considered are well documented, the initial
investments could, in selected cases, be doubled if accessibility problems
and startup difficulties are severe. Another key point of the analysis is
that any loss in boiler efficiency due to a NO control, cited in the
^
tables as a fuel penalty, would result in a severe cost penalty. For
example, in the case of OFA for the typical 540 MW opposed wall unit treated
in Table 7-20, a 0.25 percent lost in unit efficiency resulted in an annual
cost of $113,000 or $0.21/kW-yr based on a 7000 hour operating year. This
is almost a third of the total control cost. Thus, there is a definite need
for careful, long-term monitoring of control behavior to unequivocally
determine any losses in boiler efficiency or additional maintenance
requirements.
The results in Tables 7-19 through 7-24 represent the best
projections to date based on discussions with equipment manufacturers and
vendors, retrofit design studies, and detailed process analyses. They are
in basic agreement with the work of Selker and Krippene discussed in
Section 7.1 if adjustments to constant year/dollars are made. The operating
costs presented here for tangential coal-fired NO control are somewhat
^
higher than Selker's estimates, and are thus conservative. However, the
main thrust in this analysis was to compare control costs for a variety of
applications, all on an equal and well documented basis.
7-42
-------
Based on the favorable process analysis results presented in
Section 6, it is evident from an examination of Tables 7-25 and 7-26 that
OFA and LNB are the preferred, cost-effective NO controls for coal
A
firing. For very high level of NO control of coal-fired units (170
A
ng/J), both OFA and LNB would be required. For more moderate levels of
control, LNB would seem to be less expensive and more cost-effective than
OFA in reducing NO .
A
Table 7-26 also presents projected retrofit control requirements for
alternative NO emissions levels. Control requirements are recommended to
A
achieve a given NO emission level. These requirements and techniques
y\
combined with the cost to control column, complete the cost effectiveness
picture. Since this study has been completed (1978), manufacturers have
acquired more long term experience with low NO burners, and are now
recommending LNB over OFA even for retrofit applications (Reference 7-24).
In any event, the choice of retrofitting LNB or OFA must be decided on a
case-by-case basis, based on fuel/furnace design considerations. For
example, although LNB may appear to be preferable operational-wise, as well
as cost-wise, the existing furnace may not be of the proper design or size
to accommodate the larger, less turbulent flame. In that case, OFA may be
more suitable. Another example would be a furnace firing a high slagging
potential coal; OFA would not be attractive because it could increase that
slagging potential.
Burners out of service was treated in the cost analysis not as a
recommended control technique for coal firing but to show the prohibitively
high cost of derating. As detailed in Table 7-22, this high cost was due
principally to the need to purchase make up power from elsewhere and to
account for the lost capacity of the system through a lost capacity capital
charge.
As far as moderate control for oil- and gas-fired units, off
stoichiometric combustion via BOOS appears to be the preferred route, as
indicated in Tables 7-25 and 7-26. Initial investment is minimized since
there are no associated major hardware requirements, only engineering and
startup costs. To reach the next level of NO control, 86 ng/J, FSR and
OFA installation would seem to be in order. However, the increase in cost
from $0.49/kW-yr for BOOS to $3/kW-yr for FGR + OFA does not make the
option attractive. Besides, from a regulatory point of view, requirement of
7-43
-------
TABLE 7-25. SUMMARY OF RETROFIT CONTROL COSTS3 (1977 DOLLARS)
Boiler /Fuel Type
Tangential /Coal -Fired
OFA ^ 2:^5 M'Oe
Opposed Wall /Coal-Fired
OFA
LNB
BOOS
Single Wall/Oil- and Gas-Fired
8005 -TO M».
FGR/OFA
Initial
Investment
(SAW)
0.90
0.62
2.03
0.08
0.30
5.71
Annual i zed Indirect
Operating Cost
($/kW-yr)
0.21
0.16
0.34
5.34
0.05
1.14
Annuali zed Direct
Operating Cost
($/kW-yr)b
0.32
0.52
0.06
24.78
0.44
1.91
Total to Cost
Control
($/kW-yr)b
0.53
0.69
0.40
30.12
0.49
3.05
•p.
-P*
aBased on assumptions given in text and costs input parameters listed in Appendix E.
bBased on 7000 h operating year. Typical costs only.
cAssumes twenty percent derate required.
-------
TABLE 7-26. PROJECTED RETROFIT CONTROL REQUIREMENTS FOR ALTERNATE
NOX EMISSIONS LEVELS
Fuel/N0x Emission Level
ng/J (lb/10 Btu)
Coal
301 (0.7)
258 (0.6)
215 (0.5)
172 (0.4)
Oil
129 (0.3)
86 (0.2)
Gas
129 (0.3)
86 (0.2)
43 (0.1)
Recommended Control
Requirement3
OFAd
OFAd
LNB
OF A + LNB
BOOS
FGR + OFA
BOOS
FGR + OFA
FGR + OFA
Cost to Control
$/kW-yrb»c
0.50 to 0.70
0.50 to 0.70
0.40 to 0.50
0.95 to 1.20
0.50 to 0.60
-3.00
0.50 to 0.60
-3.00
~3.00
LEA considered standard operating practice.
Typical installation only; could be significantly higher.
C1977 dollars.
As manufacturers acquire more experience with LNB, they are now
recommending LNB over OFA.
7-45
-------
the emission level achievable with FGR + OFA would not be particularly
attractive since oil- and gas-fired units with BOOS would already have very
low NO emissions, 129 ng/J, compared to coal-fired units. Furthermore,
rt
with impending fuel shortages, oil- and gas-fired units will be eventually
phased out.
7.4 CONTROL COSTS FOR NSPS BOILERS
Estimating the incremental costs of NO controls for NSPS boilers
A
is in some respects an even more difficult task than costing retrofits.
Certain modifications on new units, though effective in reducing NO
A
emissions, were originally incorporated due to operational considerations
rather than from a control viewpoint. For example, the furnace of a typical
unit designed to meet 1971 NSPS (301 ng/J, 0.7 lb/106 Btu) has been
enlarged to reduce slagging potential. But this also reduces NO due to
A
the lowered heat release rate, as established in Section 4. Thus, since the
design change would have been implemented even without the anticipated NO
A
reduction, the cost of that design modification should not be attributed to
NO control.
A
Babcock & Wilcox has estimated the incremental costs of NO
^
controls on an NSPS coal-fired boiler (Appendix A). The two units used in
the comparison were identical except for N0x controls on the NSPS unit
which included:
• Replacing the high turbulence, rapid mixing cell burner with the
limited turbulence dual register (low NO ) burner
/\
• Increasing the burner zone by spreading the burners vertically to
include 22 percent more furnace surface
• Metering and controlling the airflow to each row of burners using
a compartmented windbox
To provide these changes for NO control, the price increase was about
A
$1.75 to $2.50/kW (1977 dollars). If these costs are annualized according
to the format of Section 7.2, they translate to 0.28 to 0.40 $/kW-yr.
Comparing these costs with the retrofit costs (0.40 to 0.70 $/kW-yr
for LNB or OFA) presented in Section 7.3 and considering the better NO
A
control anticipated with NSPS units, it is certainly more cost-effective to
implement controls on new units. Furthermore, fewer operational problems
are expected with factory installed controls.
7-46
-------
Foster Wheeler has provided the NOX Environmental Assessment
Program with a detailed design study aimed at identifying the incremental
costs of NO control inclusive in NSPS units (Appendix B). Foster Wheeler
looked at these unit designs with the following results:
Boiler Design Relative Cost
Unit 1: Pre-NSPS base design 100
Unit 2: Enlarged Furnace, no 114
active NO control
A
Unit 3: NSPS design; enlarged 115.5
furnace, low NO burner,
A
perforated hood, overfire
air, boundary air
Assuming the cost of a pre-NSPS coal fired boiler to be about $100/kW
in 1969, or $180/kW in 1977 construction costs (References 7-25 through
7-27), the incremental cost of active NOX controls (LNB plus OFA) is
$2.70/kW, or about $0.43/kW-yr annualized. The Foster Wheeler estimate
which includes both LNB and OFA, thus agrees quite well with the Babcock &
Wilcox estimate, which includes only LNB and associated equipment.
Recent emissions test data from the above manufacturers indicate that
control levels of 172 to 215 ng/J (0.4 to 0.5 lb/106 Btu) for coal-firing
may be possible with the combination of overfire air and low NOX burners
(References 7-28 and 7-29). However, the viability of long term controlled
operation for a variety of coals remains to be demonstrated.
7-47
-------
REFERENCES FOR SECTION 7
7-1. Personal communication, Vatsky, J., Foster Wheeler Energy
Corporation, Livingston, NJ, October 1977.
7-2. Personal communication, Campobenedetto, E. J., Babcock & Wilcox
Company, Barberton, OH, October 1977.
7-3. Personal communication, Sadowski, R., Riley Stoker Corporation,
Worcester, ME, October 1977.
7-4. Personal communication, Pepper, W., Los Angeles Department of Water &
Power, Los Angeles, September 1977.
7-5. Personal communication, Strehlitz, F., Pacific Gas & Electric
Company, San Francisco, September 1977.
7-6. Personal communication, Meinzer, R. P., and Gabrielson, E., San Diego
Gas & Electric Company, San Diego, September 1977.
7-7. Bartok, W., et al., "Systems Study of Nitrogen Oxide Control Methods
for Stationary Sources," Final Report — Volume II, Esso Research and
Engineering Company, prepared for NAPCA, NTIS-PB 192 789, November
1969.
7-8. Barr, W. H., Strehlitz, F. W., and Dalton, S. M., "Retrofit of Large
Utility Boilers for Nitric Oxide Emission Reductions -- Experience
and Status Report," presented at the 69th Annual AICHE Meeting,
Chicago, November 1976.
7-9. Personal communication — letter from Pepper, W., Los Angeles
Department of Water and Power to Acurex Corporation, May 1975.
7-10. Lachapelle, D. G., et al., "Overview of the Environmental Protection
Agency's NO Control Technology for Stationary Combustion Sources,"
presented at the 67th AIChE Annual Meeting, December 1974.
7-11. Selker, A. P., "Program for Reduction of NO from Tangential
Coal-Fired Boilers, Phase II and Ila," EPA-650/2-73-005a and b,
NTIS-PB 245 162/AS and NTIS-PB 246 889/AS, June 1975.
7-12. Blakeslee, C. E., and Selker, A. P., "Program for the Reduction of
NO from Tangential Coal-Fired Boilers, Phase I," Environmental
Protection Technology Series, EPA-650/2-73-005, NTIS-PB 226 547/AS,
August 1973.
7-13. Krippene, B. C., "Conventional NO Reduction Techniques for Oil and
Gas-Fired Boilers," presented at NOX Control Technology Workshop,
Pacific Grove, CA, October 1977.
7-48
-------
7-14. Grant, E. L., Ireson, W. 6., Leavenwouth, R. S., Principles of
Engineering Economy, Sixth Edition, Ronald Press Co., New York,
1975:
7-15. McGlamery, G. G., et al., "Detailed Cost Estimates for Advanced
Effluent Desulfurization Processes," EPA-600/2-75-006, January 1975.
7-16. Waitzman, D. A., et al., "Evaluation of Fixed-Bed Low-Btu Coal
Gasification Systems for Retrofitting Power Plants," EPRI Report
203-1, February 1975.
7-17. Ponder, W. H., Stern, R. D., and McGlamery, G. G., "SO Control
Methods Compared," The Oil and Gas Journal, pp. 60 to 66, December
1976.
7-18. Engdahl, R. B., "The Status of Flue Gas Desulfurization," ASME Air
Pollution Control Division News, April 1977.
7-19. Princiotta, F. T., "Advances in SO Stack Gas Scrubbing," Chemical
Engineering Progress, pp. 58 to 64, February 1978.
7-20. Edison Electric Institute, "Statistical Year Book of the Electric
Utility Industry for 1976," New York, EEI, October 1977.
7-21. Energy Data Systems, Environmental Protection Agency, Office of Air
and Waste Management, Office of Air Quality Planning and Standards,
Strategies and Air Standards Division.
7-22. Richardson Rapid System, Richardson Engineering Services, Inc.,
1977-1978.
7-23. Norton, D. M., et al., "Status of Oil-Fired NOX Control
Technology," in Proceedings of the NOX Control Technology
Seminar. EPRI SR-39, February 1976.
7-24. Vatsky, J., Foster Wheeler Energy Corporation, Letter to
K. J. Lim, Acurex Corporation, January 1980.
7-25. Olmsted, L. M., "18th Steam Station Cost Survey," Electrical World.
Volume 180, No. 9, pp. 39 through 54, November 1973.
7-26. "Economic Indicators," Chemical Engineering. Volume 85, No. 11,
pg. 189, May 1978.
7-27. Olmsted, L. M., "19th Steam Station Cost Survey," Electrical World.
Volume 184, No. 10, pp. 43 through 58, November 1975.
7-28. Vatsky, J., "Experience in Reducing NOX Emissions on Operating
Steam Generators," in Proceedings: Second NOyControl Technology
Seminar, pp. 7-1 through 7-17, EPRI FP-1109-SR, July 1979.
7-29. Barsin, J. A., "Pulverized Coal Firing NOX Control," in
Proceedings: Second NOV Control Technology Seminar,
pp. 8-1 through 8-22, EPRI FP-1109-SR, July 1979.
7-49
-------
SECTION 8
ENVIRONMENTAL ASSESSMENT
The evaluation of the effectiveness and impacts of NO combustion
/V
controls applied to utility boilers must also include an analysis of the
effect of these controls on incremental emissions of other pollutants as
well as NOV. Section 8.1 summarizes the demonstrated or predicted effects
/\
of controls on waste stream pollutant concentrations, including the latest
results from a coal-fired utility boiler field tested under this NOX EA
program. As a step toward quantifying how low NO firing affects the
^
environmental impact of a combustion source, a Source Analysis Model, SAM IA
(References 8-1 and 8-2), was applied to the results of that utility boiler
field test.
To complete the environmental assessment picture, Sections 8.2
through 8.5 summarize the highlights presented earlier in this report of
process impacts including energy impacts, economic impact, and control
effectiveness of combustion modifications, respectively. With these
analyses in hand, Section 8.6 concludes with control technology and R&D
recommendations.
8.1 ENVIRONMENTAL IMPACT
Modification of the combustion process in utility boilers for N0x
control in turn reduces the ambient levels of NO^, which is both a toxic
substance and a potential precursor for nitrate aerosols, nitrosamines, and
other elements of photochemical smog. These modifications can also cause
changes in emissions of other combustion generated pollutants. If
unchecked, these changes, referred to here as incremental emissions, may
have an adverse effect on the environment, in addition to effects on overall
system performance. However, since the incremental emissions are sensitive
to the same combustion conditions as NO , they may, with proper engineering,
3\
also be held to acceptable levels during control development so that the net
8-1
-------
environmental benefit is maximized. In fact, control of incremental
emissions of carbon monoxide, hydrocarbons, and particulate has been a key
part of all past NO control development programs. In addition, recent
/\
control development has been giving increased attention to other potential
pollutants such as sulfates, organics, and trace metals.
This section presents data obtained to date on the demonstrated or
predicted effects of combustion modification NO controls on incremental
emissions from utility boilers. Attention is focused on the flue gas
emissions, as the limited data base is concentrated in this area. Besides,
flue gas stream environmental impacts are expected to dominate over those of
liquid and solid effluent streams, as will be discussed later in this
section. Emission categories discussed in detail are incremental carbon
monoxide, vapor phase hydrocarbons, particulates, trace metals, sulfates,
and condensed phase organic compounds. Where appropriate, the results from
low NO testing of a 180 MW tangential coal-fired utility boiler will be
A
highlighted, as that test program emphasized the impact of controls on
incremental emissions and represents the latest reported field results.
Details of the boiler tested, the test program performed, and test results
obtained are presented in a separate report (Reference 8-1).
8.1.1 Carbon Monoxide Emissions
Since large quantities of CO in the flue gas of utility boilers mean
decreased efficiency, utility boilers are operated to keep CO emissions at a
minimum. Furthermore, if flue gas CO levels reach concentrations in excess
of 2000 ppm, the potential exists for severe equipment damage from potential
explosions in flue gas exit passages. Thus, the degree to which a NO
reduction technique is allowed to increase CO is limited by other than
environmental concerns. In general, a NO control method can be applied
A
until flue gas CO reaches about 200 ppm. Further application is then
curtailed.
NOX control effects on CO emissions are highly dependent on the
equipment type and the fuel fired. In utility boilers of newer design, it
is generally possible to achieve good NO reduction without causing
^
significant CO production. This is possible because newer burner and
furnace designs allow for better combustion air control and longer
combustion gas residence time. In addition, oil- and coal-fired boilers
usually emit very low CO levels during low NO combustion because smoke
A
8-2
-------
and soot production generally occurs with these fuels before significant CO
levels are attained. Since boiler operators strive to keep combustible
losses to a minimum, conditions which result in soot formation are avoided,
resulting in correspondingly low CO levels. A summary of the field data on
the effects on CO emissions of the more extensively implemented combustion
modifications are shown in Table 8-1. These data are discussed below for
each combustion NO control.
7\
As the data in Table 8-1 illustrate, lower excess air levels in
utility boilers can have profound effects on CO emissions. In virtually all
instances CO emissions increased significantly when excess 02 levels were
reduced 30 to 60 percent. Gas-fired boilers showed emission increases up to
400 percent when excess Op was lowered over this range, while oil-fired
boilers were less sensitive, and showed CO emission increases from 0 to 120
percent. However coal-fired boilers were the most sensitive to excess air
reductions. Reducing excess 0? by 40 to 60 percent gave 100 to 1,000
percent increases in CO emissions.
Off stoichiometric combustion has proven to be a very effective NOX
reduction technique for large steam generators. As noted in Section 3, it
can be implemented in a variety of ways including burners out of service,
overfire air ports, and biased firing. In all cases, the effectiveness of
off stoichiometric combustion in reducing NO emissions depends in large
A
part on the fraction of total combustion air that can be introduced into the
second combustion stage. It is in this second stage that complete
combustion of the fuel is achieved. CO emissions rise when this second
stage combustion does not go to completion prior to quenching in the
convective section. This is caused by a combination of the first stage
being too fuel rich and the mixing of second stage air being too slow for
the residence time provided. During development of retrofit or new design
controls, these parameters are usually selected so that CO emissions are
acceptable.
The effectiveness of off stoichiometric combustion in reducing NO
/\
formation while keeping CO emissions low is highly dependent on specific
equipment type. New utility boilers with multiburner furnaces are
especially amenable to this technique because it is generally not
8-3
-------
TABLE 8-1. REPRESENTATIVE EFFECTS OF NOX CONTROLS ON CO EMISSIONS
FROM UTILITY BOILERS (References 8-3 through 8-7)
NO Control
Low Excess Air
Off
Sto1ch1wnetr1c
Combustion
Flue Gas Red re ulafion
Load Reduction
Fuel
Natural Gas
Oil
Coal
Natural Gas
Oil
Coal
Natural Gas
Oil
Natural Gas
Oil
Coal
CO Emissions (ppm)a
Baseline
14
86
12
8
14
19
85
15
19
42
20
24
27
27
14
86
12
14
19
85
15
28
24
27
17
31
29
29
175
21
14
52
12
14
19
30
15
19
20
25
31
24
NOX Control
68
74
61
8
34
42
53
20
19
93
60
283
81
225
16
67
13
14
21
85
21
37
23
26
40
45
35
22
65
9
13
52
15
21
14
5
19
22
41
19
e
12
*3S 02. dry basis.
8-4
-------
difficult to adequately distribute secondary air and assure complete
combustion in these sources. Consequently, implementing off stoichiometric
combustion in utility boilers is expected to elicit little effect on
incremental CO emissions. This conclusion is certainly borne out by the
representative data presented in Table 8-1.
The use of flue gas recirculation (FGR) for NO control has, in
/\
practice, been restricted to gas- and oil-fired units. This technique is
ineffective in reducing fuel NO production, the predominant source of
A
NO in coal firing (Reference 8-8). When FGR is implemented, 10 to 30
A
percent of the total burner gas flow is recycled flue gas from the boiler
exhaust. Further FGR increases can cause flame instability due to reduced
flame temperatures and oxygen availability. Theoretically, FGR can lead to
increased CO emissions, but unacceptable flame instabilities usually occur
before the onset of CO or smoke production. Thus, as Table 8-1 shows, the
use of FGR has not caused increased CO emissions. On the contrary, CO
emissions have decreased in the cases shown.
Since load reduction in steam generators necessitates increased
excess air levels to maintain good furnace air-fuel mixing and steam
temperature control, increased CO emissions using this NO reduction
A
technique are not expected. In addition, the increased combustion gas
residence time afforded under reduced load would tend to facilitate complete
CO burnout. As Table 8-1 illustrates, CO emissions remain relatively
unchanged with reduced load.
8.1.2 Hydrocarbon Emissions
Field test programs studying the effectiveness of NO controls
A
often monitor flue gas HC emissions as a supplementary measure of boiler
efficiency. Therefore, some data on the effect of these controls on HC
emissions are available. Three recent test programs on utility boilers
routinely measured flue gas HC (References 8-3 through 8-5). However, in
virtually all tests, both baseline and low NO operation, hydrocarbon
A
emissions were less than 1 ppm (or below the detection limit of the
available monitoring instrument). Thus, it was concluded that HC emissions
are relatively unaffected by imposing preferred NO combustion controls on
A
large utility boilers. However, this conclusion is not altogether
8-5
-------
unexpected. The presence of unburned HC in flue gases implies poor boiler
operating efficiency, and NO controls which significantly decrease
efficiency have found little acceptance.
8.1.3 Particulate Emissions
Although gas-fired units produce negligible amounts of particulate,
oil- and coal-fired utility boilers currently emit approximately 38 percent
of the nationwide particulate and smoke emissions (Reference 8-7).
Potential adverse effects on these particulate emissions from NO
A
combustion controls could therefore have significant environmental impact.
Unfortunately the optimum conditions for reducing particulate formation
(intense, high temperature flames as produced by high turbulence and rapid
fuel-air mixing), are not the conditions for suppressing NO formation.
A
Therefore, most attempts to produce low NO combustion designs have been
A^
compromised by the need to limit formation of particulates. This compromise
has generally produced designs which maintain a well controlled, cool flame,
while still providing sufficient gas residence time to completely burn
carbon containing particles.
The NO combustion controls currently receiving the most widespread
/\
application in utility boilers are low excess air, off stoichiometric
combustion, and flue gas recirculation (for gas and oil). The altered
combustion conditions resulting from these modifications can be expected to
influence emitted particulate load and size distribution. For example,
smoke and particulate emissions tend to increase as available oxygen is
reduced (soot emissions increase and ash particles contain more carbon).
Thus the degree to which excess air can be lowered to control NOX is
usually limited by the appearance of smoke, especially in oil-fired units.
Of course, the extent to which excess air can be limited depends on
equipment types and design. Many modern burners can operate on as little as
3 to 5 percent excess air.
Similarly, the degree to which off stoichiometric combustion can be
employed is frequently limited by the degree to which the primary flame zone
can be stably operated fuel-rich, how well the second stage air mixes with
primary stage combustion products, and the residence time for combustion in
the second stage. Soot and carbon particles formed in the fuel-rich primary
stage tend to resist complete combustion downstream of the primary stage.
8-6
-------
On the other hand, flue gas recirculation on oil-fired units can
serve to decrease particulate emissions by providing more intimate mixing.
Kamo, et al. (Reference 8-9) have demonstrated that recirculation rates of
40 to 50 percent on a heater-sized oil-fired furnace reduced the smoke
number significantly.
Published data on the effects of NO reduction techniques on
/N
particulate emissions from utility boilers are scattered and insufficient
for indepth analysis. Table 8-2 summarizes the particulate emissions data
obtained during four recent field test programs which studied coal-fired
utility boilers (References 8-3, 8-4, 8-5, and 8-10). During the studies,
particulate measurements were recorded under baseline and low NO
A
conditions. Since these NO conditions were generally produced by a
/\
combination of low excess air and off stoichiometric combustion, the
individual effect of each technique on particulate emissions cannot be
determined. Nevertheless, the data do show that particulate emissions are
relatively unaffected by low NO firing.
A
The effects of low NO firing on carbon (or combustible) content of
/\
the particulate are also shown in Table 8-2. Although the data are quite
scattered, it appears that carbon losses increase for single wall- and
opposed wall-firing under low NO conditions, but decrease slightly for
^
tangential firing. However, the changes are small and may not be
significant.
The effect of low NO conditions on emitted particle size
/\
distribution have also been investigated to a limited extent
(References 8-3, 8-4, and 8-10). The data from a study of particle size
distribution in six boilers are summarized in Table 8-3. As the table
shows, no significant changes were noted in five of the boilers. For the
opposed wall coal-fired boiler, a distinct shift to smaller particles was
noted, but the author reported problems with the sampling and particle
sizing equipment in this test, so the data may not be significant
(Reference 8-4).
8.1.4 Trace Metals
Emissions of trace metals are a concern for combustion sources firing
coal and residual oil. They are a lesser problem in sources firing
distillate fuels since trace metal concentrations in distillate oils are
generally much lower than those in residual oils. Trace metals from
8-7
-------
TABLE 8-2. EFFECTS OF NOX CONTROLS ON PARTICULATE EMISSIONS FROM
COAL-FIRED UTILITY BOILERS (References 8-3, 8-4, 8-5 and 8-10)
Firing Mode
Single Wall
Single Wall
(wet bottom)
Opposed Wall
Tangential
Parti cu late Emissions
(wg/J)
Baseline
2.3
1.9
2.0-3.4
0.7-1.3
1.6-2.1
3.3-3.8
1.3-1.7
1.1-1.8
1.3-1.4
0.9-2.2
1.4
Low NOX
1.8-2.0
2.3
1.7-2.4
0.6-1.8
1.9-2.6
2.4-3.6
1.3-1.8
1.2-3.0
1.0-1.3
2.4-2.4
1.2-1.4
Percent Carbon
in Par ticu late
Baseline
9.1-13.0
5.1
5.9-6.3
1.3-2.2
1.1-2.7
0.5-0.7
2.8-5.5
0.9-2.0
0.6-0.7
24.2-25.8
2.7
Low NOX
6.2-8.1
8.2
8.5-12.4
1.7-5.8
3.4-5.7
0.2-0.5
6.7-11.8
0.8-1.5
0.2-0.6
14.8-18.8
2.3-2.8
References
8-4
8-5
8-5
8-10
8-4
8-5
8-5
8-4
8-4
8-5
8-3
8-8
-------
TABLE 8-3. EFFECT OF NOX CONTROLS ON EMITTED PARTICLE SIZE
DISTRIBUTION FROM UTILITY BOILERS
Equipment Type:
Fuel
Tangential
Coal
Tangential
Coal
Opposed Wall
Coal
Single Wall
Coal
(wet bottom)
Opposed Wall
Oil
Firing
Condition
Baseline
Low NOX
Baseline
Low NOX
Baseline
Low NOX
Baseline
Low NOX
Baseline
Low NOX
Average Weight Percent Particles of Size:
>2.5 pro
81.78
80.74
92.75
93.94
92.56
59.37
85.0
86.43
84.6
80.0
2.0 ym
9.12
8.91
2.97
1.89
2.59
10.77
3.5
5.27
0.9
2.0
1.5 pm
2.01
2.28
0.70
0.59
0.62
4.08
2.27
1.8
1.7
2.0
1.0 pm
2.64
2.92
0.97
0.86
0.96
5.89
2.23
1.97
1.3
2.5
0.5 urn
2.92
3.25
1.21
1.10
1.45
9.55
1.17
1.27
1.3
2.7
<0.5 urn
1.55
1.88
1.38
1.61
1.84
10.36
5.83
3.27
8.2
10.8
References
8-4
8-4
8-4
8-10
8-10
Tangential
Coal
Baseline
Low NOX
>10 pm
36.8
35.4
3-10 urn
40.3
42.1
1-3 urn
18.1
17.6
<1 urn
4.8
4.9
8-3
8-9
-------
stationary sources are emitted to the atmosphere with the flue gas either as
a vapor or condensed on particulate. The quantity of any given metal
emitted, in general, depends on:
• Its concentration in the fuel
• The combustion conditions in the boiler
t The type of particulate control device used, and its collection
efficiency as a function of particle size
• The physical and chemical properties of the element itself
For present purposes, the trace metal composition of the fuel is
considered a given quantity not subject to manipulation. Therefore although
composition has a controlling effect on the absolute trace metal emissions
from a combustion source, it is not considered as a factor to explain the
effects NO controls have on incremental trace metal emissions.
A
It has become widely recognized that some trace metals tend to
concentrate in certain waste particle streams from a boiler (bottom ash,
collector ash, flue gas particulate), while others do not (References 8-11
through 8-18). Based on this phenomenon, three classes of partitioning
metals have been defined (References 8-11 and 8-12).
t Class I: 20 metals (Al, Ba, Ca, Ce, Co, Eu, Fe, Hf, K, La, Mg,
Mn, Rb, Sc, Si, Sm, Sr, Ta, Th, and Ti). These are found in the
bottom ash or slag, the particle collector inlet flyash, and the
collector outlet flyash in approximately the same mass
concentrations.
• Class II: 9 metals (As, Cd, Cu, Ga, Pb, Sb, Se, Sn, and Zn).
These are not usually found in bottom ash or slag, but are found
in flyash. Mass concentrations in particle collector inlet
flyash are generally less than in collector outlet flyash.
t Class III: Hg, and possibly Se. These are usually emitted as
vapors in the flue gas.
Another set of elements (Cr, Cs, Na, Ni, U, and V) exhibits properties
intermediate between Classes I and II.
Other work has shown that the Class II metals, As, Cd, Pb, Sb, Se,
and Zn, along with Ni, Cr, and V become increasingly more concentrated in
flyash particles as particle size decreases (Reference 8-13). Cd, Pb, Ni,
Sb, Se, Sn, V and Zn all appear to have a mass mean diameter (HMD) of less
8-10
-------
than 1 ym in the atmosphere. The more common Class I metals, Fe, Al, and
Si, have MMDs of 2.5 to 7.0 ym (Reference 8-19).
The most logical explanation for this segregation behavior involves a
volatilization-condensation mechanism (Reference 8-11). In its simplest
form, the argument says that Class I metals have boiling points sufficiently
high that they are not volatilized in the combustion zone. Instead, they
form a melt of relatively uniform concentration, which becomes both bottom
ash or slag, and flyash. Thus, Class I elements remain in a condensed phase
throughout the boiler and show little partitioning with particle size. By
contrast, Class II metals have boiling points below peak combustion
temperatures, so they are volatilized in the combustion zone and do not
become incorporated in the slag. As combustion gases cool by traveling
through the boiler, these elements either form condensation nuclei or
condense onto other available solid surfaces (predominantly Class I mineral
particles). Since the available surface area to mass ratio increases as
particle size decreases, Class II elements concentrate in small particles.
This partitioning mechanism is further substantiated by observations that
certain Class II metals exhibit higher surface concentrations than bulk
concentrations in fine particles (Reference 8-20).
This simple mechanism described above does not fully account for all
experimental observations. For example, Ca and Cu behave as high boiling
point metals, whereas Rb, Cs, and Mg behave as volatile elements.
Therefore, the volatilization-condensation mechanism has been extended as
follows (Reference 8-11):
• Trace elements in coal are present as aluminosilicates, sulfides,
and organometallics
• On combustion, the aluminosilicates melt to form slag or bottom
ash, and flyash
• In the reducing atmosphere during initial stages of combustion,
metal sulfides are reduced to vapor phase metal; at the same time
the organic matrix of organometallics oxidizes, leaving
volatilized metal
• Volatilized metals may themselves become oxidized to less
volatile oxides
• As the combustion gas cools, these volatile species condense onto
available solid surfaces, and concentrate in small particles
8-11
-------
t Since slag and flue gas are in contact for only a short time,
little volatile condensation in slag occurs
This extended mechanism is indirectly supported by the fact that Class I
metals are largely geochemical lithophiles (readily associated with
aluminosilicate minerals), while Class II metals are largely chalcophiles
(readily incorporated into sulfide minerals).
In all mechanisms the Class III metals, Hg and to some extent Se,
remain vaporized through the stack and are emitted as flue gas vapor
components. Some 90 percent of Hg emissions (Reference 8-21) and about
20 percent of Se emissions (Reference 8-11) are emitted as vapors.
Regardless of the exact mechanism for the trace metal partitioning
phenomenon, the partitioning significantly influences trace metal emissions
from combustion sources with particulate control devices. All particle
collection devices are more efficient at collecting large particles than
small particles. Since Class II metals in flyash occur in smaller particles
than Class I metals, a larger fraction of the Class II elements introduced
into a boiler will be emitted from sources equipped with particulate control
units.
This behavior is illustrated by recent trace metal emissions data
from industrial boilers (Reference 8-22). Figure 8-1 shows the
concentration of several Class I metals measured in particle samples from
different points in a coal-fired industrial boiler. Figure 8-2 shows the
same profile for several Class II elements. As the partitioning theory
predicts, the concentration of Class I metals remains fairly constant
throughout the boiler. On the other hand, flyash concentrations of Class II
elements increase toward the flue gas exit. The expected increase in
concentrations in the collector effluent ash over collector inlet ash and
collected ash is quite significant.
By understanding trace metal partitioning and concentration in fine
particulate, it is possible to postulate the effects NO combustion
/\
controls will have on incremental trace metal emissions. Several NO
X
controls for boilers result in lowered peak flame temperatures (off
stoichiometric combustion, flue gas recirculation, reduced air preheat, load
reduction, and water injection). The volatilization-condensation theory
predicts that if the combustion temperature is reduced, less
8-12
-------
c
:
01
E
HI
5,000
2,500
Ash up-
stream of
collector
Ash down-
stream of
collector
Figure 8-1. Partitioning of Class I elements (Reference 8-22)
8-13
-------
in
§
(J
c
o
u
10
0
10
0
zoo
0
200
0
20
10
0
5
0
200
0
1,000
• • Arsenic
100 • - Copper
100
Z.5
100 • • Zinc
Coal
Cadmium
• Selenium
-£.
600 • Vanadium
*-f f £,
Furnace
bottom ash
S/JS
Ash up-
stream of
collector
/ / j.
Ash in
collector
Ash down-
stream of
collector
' / / /
/
Figure 8-2. Partitioning of Class II elements (Reference 8-22).
8-14
-------
Class II metal will initially volatilize, hence less will be available for
subsequent condensation. Under these conditions (lowered flame
temperature), it is expected that less Class II metal (the segregating trace
metals) will be redistributed to small particulate. Therefore, in boilers
with particulate controls, lowered volatile metal emissions should result.
Class I metal (the nonsegregating trace metals) emissions should remain
relatively unchanged. Since 8 of the 20 most toxic elements in air are
Class II metals, obtaining trace metal partitioning data should be given
high priority (Reference 8-23).
Lowered local 02 concentrations are also expected to affect
segregating metal emissions from boilers with particle controls. Lowered
Op availability decreases the possibility of volatile metal oxidation to
less volatile oxides. Under these conditions Class II metals should remain
in the vapor phase into the cooler sections of the boiler. More
redistribution to small particles should occur and emissions should
increase. Again, nonsegregating metal emissions should be unaffected. This
behavior is expected when low excess air is implemented. Other combustion
NO controls which decrease local Oo concentrations (off stoichiometric
A C.
combustion and flue gas recirculation) also reduce peak flame temperature.
For these, the effect of lowered combustion temperature might be expected to
predominate.
The effect of NO combustion controls on segregating metal
A
emissions from combustion sources without particle collection devices should
be marginal at best. Particle redistribution will not affect mass emissions
because all particulate produced is emitted from these sources. However,
since trace metal condensation on internal boiler surfaces may occur,
conditions which decrease the extent of Class II metal volatilization
(lowered peak flame temperature) might cause a slight decrease in
segregating metal emissions. Conversely, conditions which increase metal
volatility (low local 02 concentrations) may cause slight increases in
volatile metal emissions.
Trace metal sampling was performed at a coal-fired utility boiler
under baseline and low NO conditions (burners out of service or biased
A
burner firing) as part of the NO EA Program (Reference 8-3). Trace
A
element concentrations in the bottom ash were compared to those in the
flyash (boiler outlet) to determine if trace element stream partitioning
8-15
-------
could be observed. The trace elements were placed into three groupings
depending on whether; (EQ) the metal was partitioned about equally between
bottom ash and flyash (less than a factor of two difference, behavior
expected of Class I elements); (FA) the material was preferentially
concentrated (by a factor of two or greater) in the flyash (behavior
expected of Class II elements) or; (BA) the material was concentrated in the
bottom ash (by a factor of two or greater). As shown in Table 8-4, almost
all of the trace elements had concentrations which were enhanced at the
flyash inlet or partitioned equally between bottom ash and flyash streams.
In general, the partitioning tendencies found in this test agree with the
expectations discussed earlier in this section. However, low NO firing
n
implementation shows little, if any, effect on trace element partitioning
based on the concentration doubling criteria used.
The trace element data from the coal-fired utility boiler were also
examined to determine if partitioning occurs with particle size. Table 8-5
illustrates the presence of this effect at the electrostatic precipitator
inlet. The concentrations of Sb, Pb, Zn, Cl, F, sulfate, and ammonium were
found to be higher in the finer size particulates. In general, the trace
metal behavior is in accordance with partitioning theory. Again though, low
NOX firing seems to have little effect on the tendency to partition with
particle size.
In summary, based on the limited test data from one coal-fired
utility boiler, low NOX firing appears to have little effect on the
partitioning of trace elements between bottom ash and flyash, and little
effect on the segregation of trace species within experimental error
(Reference 8-3).
8.1.5 Sulfate Emissions
Ambient sulfate levels have recently become a matter of increasing
concern in regions with large numbers of combustion sources, notably
boilers, firing sulfur-bearing coal and oil. Although the direct health
effects of high ambient sulfate levels are currently unclear (References 8-24
and 8-25), recent thought suggests that sulfates may be more hazardous than
S02. For this reason, control of primary sulfate emissions is becoming a
concern even though primary sulfates (directly emitted) comprise only 5 to
20 percent of ambient sulfate on a regional basis (Reference 8-25).
8-16
-------
TABLE 8-4. TRACE ELEMENT PARTITIONING - BOTTOM ASH/FLYASH - IN
COAL-FIRED UTILITY BOILER (Reference 8-3)
Antimony
Arsenic
Barium
Beryllium
Bismuth
Boron
Cadmium
Chromium
Cobalt
Copper
Iron
Lead
Manganese
Mercury
Molybdenum
Nickel
Selenium
Te 1 1 er i urn
Thallium
Tin
Titanium
Uranium
Vanadium
Zinc
Zirconium
Chloride
Fluoride
Cyan i de
Nitrate
Sulfate
Ammon i urn
Baseline
X
FA
EQ
EQ
X
FA
X
EQ
EQ
EQ
EQ
FA
EQ
EQ
X
EQ
X
X
X
X
EQ
X
EQ
FA
BA
EQ
FA
X
X
FA
X
Burners Out of Service
BOOS I
X
FA
EQ
EQ
X
FA
X
EQ
EQ
EQ
EQ
FA
EQ
BA
X
EQ
X
X
X
EQ
EQ
EQ
FA
BA
EQ
EQ
X
BA
FA
FA
BOOS II
X
FA
FA
EQ
X
FA
X
EQ
EQ
EQ
EQ
FA
EQ
BA
X
EQ
X
X
X
X
EQ
X
EQ
FA
BA
EQ
BA
X
X
FA
FA
Biased Burner Firing
BIAS I
X
FA
EQ
EQ
X
X
X
EQ
EQ
FA
EQ
FA
EQ
FA
X
EQ
X
X
X
X
EQ
X
EQ
FA
BA
EQ
FA
X
X
FA
X
BIAS II
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
EQ - Material partitioned about equally between bottom ash and flyash
FA - Material preferentially concentrated in flyash
BA - Material preferentially concentrated in bottom ash
X - Insufficient data.
8-17
-------
TABLE 8-5. TRACE SPECIES PARTITIONING WITH PARTICLE SIZE — ESP INLET
OF A COAL-FIRED UTILITY BOILER (Reference 8-3)
Antimony
Arsenic
B ar i urn
Beryllium
Bismuth
Boron
Cadmium
Chromi urn
Cobalt
Copper
Iron
Lead
Manganese
Mercury
Molybdenum
Nickel
Se 1 en i urn
Te 11 er i urn
Thallium
Tin
Titanium
Uranium
Vanadium
Zinc
Zirconium
Chloride
Fluoride
Cyan i de
Nitrate
Sulfate
Ammonium
Baseline
2
2
0
0
X
X
X
0
0
0
0
2
0
0
X
0
X
2
X
0
0
X
0
2
0
2
0
X
X
0
X
Burners Out of Service
BOOS I
2
X
0
0
X
X
X
0
0
0
0
2
0
2
1
0
X
2
X
X
0
X
0
2
2
2
2
X
X
2
X
BOOS II
2
X
1
0
X
X
X
0
0
0
0
2
0
0
X
0
X
1
X
X
0
X
0
2
1
2
2
X
X
2
X
Biased Burner Firing
BIAS I
X
X
0
0
X
X
X
0
0
0
0
0
0
0
X
0
X
1
X
X
0
X
0
0
2
0
0
X
X
2
2
BIAS II
2
X
0
0
X
X
X
0
0
0
0
0
0
1
X
0
X
1
X
1
0
0
0
0
2
2
0
X
X
0
2
0 - No significant separation
1 - Concentration enhancement in >3 ym fraction
2 - Concentration enhancement in <3 ym fraction.
X - Insufficient data.
8-18
-------
Since approximately 98 percent of the sulfur introduced into a
utility boiler appears in flue gas as an oxide, applying NO controls
^
would have essentially no effect on total SO emissions. However, effects
^
on the emitted (SO, + particulate sulfate)/SO, ratio can be
0 t.
significant. Specifically, combustion conditions which limit local oxygen
concentrations would be expected to decrease the extent of S0~ to SO^
oxidation. Thus applying low excess air firing and off stoichiometric
combustion to control NOX should also lower SO., and sulfate emissions.
Confirming data, though sparse, do exist. Recent measurements have
demonstrated the expected dependence on sulfate emissions on boiler excess
air levels. Bennett and Knapp (Reference 8-26) have shown that particulate
sulfate emissions increase with increasing boiler excess Op in oil-fired
powerplants. Homolya, et al. (Reference 8-27) report a similar increase in
sulfate emissions as a percentage of total sulfur emissions with increasing
excess 0~ in coal-fired boilers. Their data, shown in Figure 8-3, show a
linear relationship between the sulfate fraction of emitted sulfur and
boiler excess 02- Still the data of Crawford, et al. (Reference 8-27)
from coal-fired utility boilers, Table 8-6, indicate that S03 emissions
are relatively unaffected by low NO firing, considering the accuracy of
A
the measurement techniques, at these low concentrations.
Sulfur emissions, under baseline and low NO (biased burner firing
A
and burners out of service) firing modes, were examined during the 180 MW
coal-fired tangential boiler test in the NO EA program (Reference 8-3).
y\
Unfortunately, due to limited coal supplies, the coal sulfur contents were
not constant throughout the test program, as noted in Table 8-7.
Nevertheless, the data do indicate that the SOo/SO, ratio is not
strongly affected by low NO firing. Furthermore, the data also show that
^
levels of sulfate in the ash samples are higher in the smaller size
particulates, regardless of firing mode. This is seen in the progressively
higher sulfate levels in the direction of flue gas flow and also in the
comparison of sulfate loading on different size particulate samples
collected by the Source Assessment Sampling System (SASS) train. Table 8-7
also indicates that low NO firing may also decrease the total sulfate
^
emissions in the flue gas.
8-19
-------
co
00
i
ro
o
CVI
O
CM
O
E
Q.
D.
"'c?
l/l
Q.
Q.
1550
1500
1450
1400 •
0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9
Boiler excess 0,
1.0 1.1 1.2 1.3 1.4
Figure 8-3. S02 conversion vs. excess oxygen in coal-fired utility
boilers (Reference 8-27).
-------
TABLE 8-6. SOX EMISSIONS FROM COAL-FIRED UTILITY BOILERS
(Reference 8-28)
Baseline
Low Excess Air
Off
Stoichiometric
Combustion
02 (X)
Boiler Exit
2.9
1.55
1.5
3.1
3.4
Stack
6.8
5.7
7.72
7.1
S02
so3
ppm Corrected to 3% 0?
944
948
1,000
Avg 974
1,010
968
Avg 989
28
13.5
35.9
Avg 25
14.0
13.9
Avg 14
8-21
-------
TABLE 8-7. SULFUR SPECIES FROM 180 MW TANGENTIAL COAL-FIRED UTILITY BOILER
Test
Heat Input Ml
Coal-Sulfur
Bottom Ash
Sulfate
Flue Gas - Cont. Monitor
S02 ppm
Mechanical Collector Ash
Sulfate
ESP Inlet - Method 8
Particulate - Sulfate
S02
S03
Units
%
pg/gm
at 3* O^
u9/gm
wQ/gm
pg/dscm
pg/dscm
(ng/J)
(ug/J)
(Pg/J)
(wg/J)
(pg/J)
(pg/J)
(pg/J)
Baseline
227.1
2.19
500
2059
1700
7800
4.4 x 10-6
1.6 x 10*
(0.834)
(7.46 x 10-4)
(1.75)
(7.75 x 10-3)
(1.1 x 10-2)
(1.55)
(5.6 x 10-3)
BIAS I
229.2
1.75
530
1527
1000
4500
3.4 x 106
1 x 104
(0.664)
(7.68 x 10-4)
(1.30)
(4.42 x 10-3)
(5.6 x 10-3)
(1.19)
(3.5 x 10-3)
BOOS II
209.8
2.13
400
1865
1300
3000
3.9 x 106
1 x 10*
(0.804)
(5.58 x 10-4)
(1.59)
(5.65 x 10-3)
(3.9 x 10-3)
(1.48)
(3.8 x 10-3)
CO
r\>
FVJ
-------
TABLE 8-7. Concluded
Test
ESP Inlet - SASS
10 + 3 vm - Sulfate
1 urn + filter - Sulfate
ESP Hopper Ash
Sulfate
ESP Outlet - Method 8
Particulate - Sulfate
S02
S03
ESP Outlet - SASS
10 + 3 pm - Sulfate
1 pm + fi Her - Sulfate
Units
pg/gm
pg/gm
pg/gm
yg/gm
pg/dscm
ug/dscm
pg/gm
yg/gm
(jjg/J)
(wg/J)
(pg/J)
(ng/J)
(wg/J)
(ng/J)
(yg/J)
(pg/J)
Baseline
4100
7300
5200
37800
4.2 x 106
1.2 x 104
7800
12000
(4.44 x 10-3)
(2.34 x 10-3)
(6.02 x 10-3)
(8.6 x 10-3)
(1.50)
(4.3 x ID'3)
(7.21 x 10-4)
(1.62 x 10-3)
BIAS I
1900
8400
4600
18200
3.5 x 106
1.1 x 104
5300
6400
(2.48 x 10-3)
(5.66 x 10-4)
(5.41 x 10-3)
(4.4 x 10-3)
(1.25)
(3.9 x ID'3)
(7.23 x 10-4)
(6.57 x 10-4)
BOOS II
3800
9200
4400
60500
4.2 x 106
1.2 x 104
—
4900
(3.62 x 10-3)
(2.55 x 10-3)
(4.71 x 10-3)
(1.1 x 10-2)
(1.57)
(4.5 x 10-3)
—
(7.83 x 10-4)
00
co
-------
In comparing the sulfate analyses of the SASS and EPA Method 8
samples, it should be noted that the Method 8 data include any condensed
material in the sampling probe wash while the SASS data account only for
sulfate on the particulate. Details of the test procedures and results are
given in Reference 8-3. These findings on sulfur emissions, though based on
limited data, are worth noting as the sulfur mass balance closure around the
180 MW unit was greater than 90 percent.
The use of post combustion ammonia injection for NO control could
rt
possibly lead to significantly increased primary sulfate emissions. Under
normal conditions, the pH of near plume liquid droplets is low,
approximately 3. At this pH, S02 solubility is low. However, if
sufficient quantities of a basic specie, such as ammonia, were present to
neutralize these droplets, SOp solubility would increase dramatically.
This could lead to significant amounts of sulfate production through
solution catalysis in the near plume (Reference 8-29). Further work is
needed in this area before any conclusions can be substantiated.
The problem of acid smut emissions, or sulfate fallout, also deserves
some discussion here. Sulfate fallout emissions of large, highly acidic
carbonaceous particulate have been experienced recently from several
residual oil-fired utility boilers in the U.S. This fallout is extremely
corrosive and since the acidic particulate is of large size (up to 100 m),
leads to fallout in the vicinity of the powerplant. Sulfate fallout is thus
of concern for potential impact on both human health and welfare. Acid
fallout has been experienced for many years in Europe due to the practice of
firing heavy oil units at lower levels of excess air than is common in the
U.S. (References 8-30 through 8-34).
Recently the problem has occurred when certain NO controls,
i A
notably off stoichiometric combustion combined with low excess air, are
implemented on residual oil-fired units. It also invariably occurs in
boilers which were originally designed to fire natural gas, but have been
converted to oil firing because of fuel availability problems.
The exact reasons for the appearance of acid fallout are not clearly
understood. However, it is clear that they are related to air heater design
and the resulting final flue gas temperature. Since natural gas contains
very little sulfur, acid mist condensation in and downstream of the air
heaters has never been a concern. Therefore, air heaters in gas-fired
8-24
-------
boilers have been designed to give lower flue gas temperatures than
corresponding air heaters in oil-fired units. However, when these same
gas-fired units are switched to oil firing, it is possible for flue gas
temperatures downstream of the air heater to approach the acid dew point.
In the absence of particulate emissions, flue gas sulfuric acid could then
condense and reevaporate through the ductwork and stack until ultimately
emitted as a finely dispersed mist.
The appearance of fallout when implementing NO controls which
enhance the production of soot particles suggests the possible next step in
the smut formation mechanism. In the presence of sufficient particulate,
flue gas sulfuric acid condenses onto particle surfaces in sufficient
amounts to cause particle agglomeration. Agglomerated particles then
deposit onto ductwork walls. These deposits continue to grow through
further agglomeration until they become large enough to fall off the wall.
Thus, emissions of large acidic particulate occur.
In light of the above, sulfate fallout emissions have been viewed as
a combined sulfate production problem and particulate production problem.
Attacks on the problem have included both reduction of acid formation and/or
condensation and suppression of carbon formation or agglomeration. Table
8-8 summarizes process modifications used or proposed in Europe and the U.S.
(References 8-31 through 8-34). It appears that incremental sulfate fallout
emissions can be suppressed if addressed during control development. The
potential for acid fallout emissions should be considered when implementing
NO controls on heavy oil-fired boilers with air preheaters and without
rt
particle collection devices.
In summary, the postulated, and in some cases demonstrated, effects
of most NO combustion controls on primary sulfates are to decrease
X
emissions or leave them unchanged. However, since there are insufficient
data to fully substantiate any real conclusion, it seems appropriate to
consider incremental sulfate emissions due to NO combustion modifications
J\
of questionable concern, except in the case of acid fallout and use of post
combustion ammonia injection. Because ammonia injection may significantly
increase near plume sulfate production through solution chemistry, its
effects on residual sulfate should be considered of definite concern.
8-25
-------
TABLE 8-8. SUMMARY OF PROCESS MODIFICATIONS TO REDUCE SULFATE FALLOUT
CO
ro
Principle
1. Suppress buildup
of acid smut
2. Prevent acid
condensation
3. Neutralize acid
smut
4. Suppress SO.,
formation
5. Reduce carbon
emissions
6. Particle
collection
Candidate Techniques
Frequent or continuous
soot blow
Reduced air preheat
Additives: dolomite,
limestone, MgO, NH^
Reduced excess air
Reduced load
Reduced catalytic
activity of superheater
Reduced sulfur in fuel;
mixed distillate/resid.
firing
Increased excess air
Better firebox mixing
Cyclone, ESP or baghouse
Size Range Affected
Large particles
Large particles
Large and small
particles
Large and small
particles
Large and small
particles
Large and small
particles
Large and small
particles
Large and small
particles
Large and small
particles
Large and small
particles
Comments
Acid smuts emitted in smaller, dispers-
able, size range; successfully tested at
Eastern Utility; promising option
Reduced efficiency; possible smut buildup
in stack at reduced size range
Reduces (50%) but doesn't eliminate acid
emissions; additives increase particle
loading
Increased efficiency; increased carbon
and CO emissions; limited by NO control
techniques
Not cost effective
Additive coating is partially effective;
operational problems
Distillate availability uncertain
Reduced efficiency; increased S03
Limited by N0x controls
Effective but costly
-------
8.1.6 Organic Emissions
The term organic emissions as used here is defined to mean those
organic compounds which exist as a condensed phase at ambient temperature.
Thus they are organics which are either emitted as "carbon on particulate"
or condense onto emitted particulate in the near-plume of a stack gas.
These compounds, with few exceptions, can be classified into a group known
variously as polycyclic organic matter (POM) or polynuclear aromatic
hydrocarbons (PNA or PAH).
POM production is generally only a minor concern in gas-fired
systems, of some concern in oil-fired sources, and of greater concern in
coal-fired equipment. Like CO and HC emissions, POM emissions are the
result of incomplete combustion. Since NO combustion controls can lead
to inefficient combustion, if not carefully applied (especially low excess
air and off stoichiometric combustion), applying these controls can
potentially lead to increased POM production.
Supporting data, however, are very limited, largely because of the
difficulty of sampling flue gas streams for POM and of accurately assaying
samples for individual POM species. Thompson et al., recently reported the
effects of staged combustion and flue gas recirculation on POM emissions
from a coal-fired utility boiler (Reference 8-8). Their data, shown in
Table 8-9, seem to indicate that POM emissions do increase with off
stoichiometric combustion, but are relatively unaffected by flue gas
recirculation alone. However, the authors state that the sampling and
laboratory analysis procedures used in obtaining the data varied over the
sample set. Thus, the conclusion that POM emissions may be increased with
low NOX firing should only be considered tentative. In another study,
Bennett and Knapp (Reference 8-26) attempted to investigate the effects of
boiler excess 02 on POM emissions from an oil-fired utility boiler. They
found that particulate carbon content increased with decreasing excess
Oo. However, because POM assay data varied widely, even for baseline
condition analyses, no conclusion regarding POM emissions was possible.
The organic analyses from low NO firing at the 180 MW unit tested
in the NO EA program yielded only general conclusions. There was not a
n
sufficient amount of organic material in any of the samples to permit
significant species identification. However, the analyses do show that
total organic emissions were slightly higher under low NO (BOOS) firing.
rt
8-27
-------
TABLE 8-9. SUMMARY OF POM EMISSIONS FROM HATFIELD UNIT NO. 3
MEASURED UPSTREAM OF ESP (Reference 8-8)
Substance
Anthracene/Phenanthrene
Methyl Anthracenes
Fluor ant hene
Pyrene
Chrysene/Benz(a) Anthracene
Total POM
Baseline
U9/MJ
54.3
16.3
15.6
4.55
0.09
90.9
BOOS Operation
ug/MJ
54.6
14.8
33.6
15.8
—
18.8
Percent Difference
from Baseline
+0.5
-9.3
+114.5
+247.9
—
+30.7
FGR Operation
U9/MJ
44.3
27.2
7.49
7.11
—
86.1
Percent Difference
from Baseline
-18.5
+66.9
-52.1
+56.3
—
-5.3
BOOS + FGR Operation
ug/MJ
70.1
30.0
13.3
14.3
--
127.7
Percent Difference
from Baseline
+29.1
+83.7
-15.2
+214.6
—
+40.5
CD
I
ro
CD
-------
Table 8-10 shows that the organic material concentrations in the bottom ash,
mechanical collector ash, electrostatic precipitator ash, and the flue gas
outlet (vapor phase) were higher for low NO firing. The flue gas outlet
/\
particulate organic content was slightly higher under baseline conditions.
However, that effect is overshadowed by the significantly larger (order of
magnitude) vapor phase organic emissions under low NO firing. Thus,
^
although organic emissions were low in these tests, there is a need to
conduct more quantitative organic analyses due to the high relative hazard
of certain organic compounds.
8.1.7 Source Analysis Model
To help quantify the potential change in environmental impact of a
utility boiler which switches from baseline to low NO firing, a source
n
analysis model, SAM IA (References 8-1 and 8-2), was applied to the effluent
data from the 180 MW coal-fired utility boiler tested in the NOX EA
program. EPA has been developing a series of source analysis models to
define methods of comparing emission data to environmental objectives,
termed Multimedia Environmental Goals (MEG's) (Reference 8-35). The model
selected for the level of data detail obtained from the utility boiler tests
was SAM IA, designed for rapid screening purposes. As such, it includes no
treatment of pollutant transport or transformation. Goal comparisons employ
threshold effluent stream concentration goals, termed discharge multimedia
environmental goals (DMEG's).
For the purposes of screening pollutant emissions data to identify
species requiring further study, a discharge severity (OS) is defined as
follows:
Concentration of Pollutant i in Effluent Stream
DS. =
DMEG of Pollutant i
The DMEG value, the threshold effluent concentration, is the maximum
pollutant concentration considered safe for occupational exposure. When DS
exceeds unity, more refined chemical analysis may be required to quantify
specific compounds present.
8-29
-------
TABLE 8-10. ORGANIC EMISSIONS FROM A 180 MW COAL-FIRED UTILITY BOILER
(Reference 8-1)
Organic Material in Ash Streams
Firing Mode Sample
Baseline
BOOS
Bottom Ash
Mechanical Collector
Electrostatic Precipitator
Bottom Ash
Mechanical Collector
Electrostatic Precipitator
Equivalent Organics in Ash Stream
yg/gm
1.5
<1.3
1.4
4.2
3.2
6.7
yg/J
2.2 x 10-6
<5.9 x ID'6
1.6 x 10-6
5.9 x 10-6
1.4 x ID'5
7.2 x 10-6
Organic Material in Flue Gas Outlet (ESP Outlet)
Firing Mode
Baseline
BOOS
Sample
Particulate
Vapor Phase
Particulate
Vapor Phase
Equivalent Organics in Flue Gas
yg/m3 yg/J
60 2.1 x 10-5
75 2.7 x lO-5
44 1.6 x 10-5
788 2.9 x 10-4
8-30
-------
To compare waste stream potential hazards, a weighted discharge
severity (WDS) is defined as follows:
WDS = ( ? OS.) x Stream Mass Flowrate,
where the DS.. are summed over all species analyzed. The WDS is an
indicator of output of hazardous pollutants and can be used to rank the
needs for controls for waste streams. It can also be used as a preliminary
measure of how a pollutant control, say a combustion modification NO
control, affects the overall environmental hazard of the source. An
extensive exposition of SAM IA and list of DMEG's are presented in
References 8-1, 8-2, and 8-36 and will not be repeated here.
SAM IA was applied to the analysis results from the 180 MW unit.
Table 8-11 summarizes the boiler outlet flue gas effluent concentrations
(ESP outlet for particulates and trace species) for baseline and low NO
^
firing. Two levels of NO reduction were tested. Retrofit bias firing
rt
gave a 32 percent NO reduction, and operation with the upper row of
^
nozzles on air only gave a 38 percent NO reduction. The furnace
^
efficiency either remained constant or increased slightly (due to lower
excess air) under low NO operation. There was no appreciable increase in
^
carbon-in-flyash with NO controls. It should be mentioned that these
^
tests were for short periods, so the long term operability under these low
NO conditions was not necessarily validated.
A
For the majority of elements listed in Table 8-11, the changes in
emission rates between baseline operation and low NO firing were within
^
the accuracy of the analysis and are not judged to be significant. Notable
exceptions are the Teachable nitrates and ammonium compounds. Here, it is
possible that local fuel rich conditions under low NO operation
A
suppresses reduced nitrogen compound oxidation normal to baseline operation.
As mentioned earlier, organic species analyses were inconclusive, though in
total organic emissions increased with low NO firing. The analysis
A
results for the other waste streams — cyclone ash, ESP ash, and bottom ash
slurry — are all presented in Reference 8-1. Table 8-12 lists the DS
values for those inorganic species or compounds where DS ^ 1. It is evident
that the gaseous pollutants, particularly SO, and NO dominate the
£ "
8-31
-------
TABLE 8-11. ANALYSIS RESULTS FOR A 180 MW TANGENTIAL COAL-FIRED UTILITY
BOILER: FLUE GAS, INORGANICS
TEST
Heat input
(% of baseline)
Emissions ^
m dry
N0v(ppm @ 3% Oy dry)
f\ C-
S02(ppm (9 3% 02 dry)
S03(ppm (3 3% 02 dry)
C0x(ppm @ 3% 02 dry)
co2(%)
02
Parti cul ate
Antimony
Arsenic
Barium
Beryllium
Bismuth
Boron
Cadmium
Chromium
Cobalt
Copper
Iron
Lead
Manganese
Mercury
Molybdenum
Nickel
Selenium
Tellurium
BASELINE
100
1.16xl06 (490)
4.18xl06 (1668)
1.45xl04 (3)
3.07xl04 (28.6)
2.72xl08 (13.9)
(5.2)
6.3xl05
3.9
95
2.25xl03
9.0
<53
2.3
1.69xl03
66
2.9xl02
4.5xl04
74
2.4xl02
1.8
1.5X102
8.4xl02
10
<4.1
BIAS (Test 1)
100.9
7.35xl05 (336)
3.5xl06 (1354)
1.32xl04 (3)
4.58xl04 (35.0)
2.82xl08 (14.4)
(4.7)
6.7xl05
<2.6
78
1.7xl03
11
<56
<2.4
4.8xl02
75
3.4xl02
3.4xl04
86
1.3xl02
3.1
<56
l.OxlO3
8.2
<4.0
BOOS (Test2)
92.4
6.54xl05 (304)
4.21xl06 (1591)
9580 (3)
3.19xl04 (21.7)
2.86xl08 (14.6)
(4.4)
4.3xl05
<2.6
81
1.5xl03
7.3
2.3xl02
8.8xl02
2.4xl03
89
3.2xl02
3.3xl04
51
1.9xl02
3.5
8.7xl02
1.5xl03
5.1
<3.7
8-32
-------
TABLE 8-11. Concluded
TEST
Thallium
Tin
Titanium
Uranium
Vanadium
Zinc
Zirconium
Chloride
Fluoride
Cyani de
Nitrate
Sulfate
Ammonium
Coal Analysis
CX
H%
0%
N%
S%
H20%
Ash*
HHV, J/g
Btu/lb
BASELINE
<2.6
<6.4
6.1xl03
<3.9
2.6xl02
4.3xl02
1.9xl02
2.7xl02
84
<1.3
<3.9
6.5x103
<5.3
63.13
4.27
7.34
1.38
2.19
2.04
19.60
26288
11302
BIAS (Test 1)
<2.7
<6.7
5.7xl03
44
2.3xl02
5.9xl02
2.6xl02
4.1xl02
3.5xl02
0.3
24
3.9xl03
7.2
63.46
4.24
7.97
1.13
1.75
2.34
19.09
26363
11334
BOOS (Test2)
<2.1
<5.1
3.6xl03
<2.1
1.6xl02
8.4xl02
6.8xl02
8.6xl02
1.2xl02
<1.3
7.7xl02
2.1xl03
1.4xl02
64
4.23
7.11
1.38
2.13
2.58
18.49
26521
11402
i
8-33
-------
TABLE 8-12. FLUE GAS DISCHARGE SEVERITY - INORGANICS: 180 MW
TANGENTIAL COAL-FIRED UTILITY BOILER
NCv
X
so2
so3
CO
CO,
L
Be
Ba
As
Ti
N (Mainly NH4)
SO,
4
Chlorides
BASELINE
129
322
15
0.77
30
4.5
4.5
48
1
0.07
6.5
0.68
BIAS
84
269
13
1.1
31
5.5
3.4
39
0.95
0.22
3.9
1
BOOS
73
324
9.6
0.80
32
3.6
3.0
41
0.60
6.1
2.1
2.1
TABLE 8-13. TOTAL WEIGHTED DISCHARGE SEVERITY (g/s) - INORGANICS:
180 MW TANGENTIAL COAL-FIRED UTILITY BOILER
Flue Gas
Cyclone Ash
ESP Ash
Bottom Ash Slurry
Total
BASELINE
4.3xl07
1.9x10''
6.1x103
5.7xlOu
4.3xl07
BIAS
3.5xl07
1.6x10^
6.1xl03
B.SxlO1*
3.5xl07
BOOS
3.7xl07
1.6X101*
S.lxlO3
4.2xlOu
3.7xl07
8-34
-------
potential toxicity of the flue gas stream. Of the trace metals, arsenic
shows the highest DS, but none of the metals show any large change under low
NO conditions. As may be expected, SO, decreased under low NO
J\ J A
operation and reduced N compounds increased.
The total weighted discharge severity for the inorganic component of
four waste streams of the boiler are compared in Table 8-13. Clearly the
flue gas stream dominates the TWOS, with the solid streams three orders of
magnitude potentially less toxic, according to the model. With low NO
^
firing, the flue gas stream TWOS is reduced, primarily due to the decrease
in NO concentration. The TWOS's for the other waste streams either
rt
decreased or were constant when going to low NO firing. As mentioned
earlier, more data are needed for waste stream organic composition before
the degree of hazard for organic compounds, relative to inorganics, can be
estimated.
From the application of SAM IA to the admittedly sparse data base of
a few short tests on a single coal-fired boiler, the results indicate that
NO controls are generally beneficial, reducing the overall adverse
^
environmental impact of waste streams. These results, along with the
general indications from other reported tests, tend to confirm that
combustion modification NO controls are environmentally sound, though
rt
work remains to confirm and correct any potential adverse environmental
impacts from incremental emissions.
8.1.8 Evaluation and Summary
Based on the previous discussions, NO control techniques and
pollutants can be classified into one of the following three groups
according to potential for increased emissions:
• High potential emissions impact, where the data clearly show that
applying the NO control results in significantly increased
/\
emissions of a specific pollutant
• Intermediate potential emissions impact, where the NO control
could conceivably cause increased pollutant emissions, but
confirming data are lacking, contradictory, or inconclusive
• Low potential emissions impact, where the data clearly show that
specific pollutant emission levels decrease or do not change when
the NO control is applied, or a similar conclusion, is
n
indicated even though data are lacking
8-35
-------
These groupings appear in Table 8-14.
As Table 8-14 illustrates, applying preferred NOX combustion
controls to boilers should have few adverse effects on incremental emissions
of CO, vapor phase hydrocarbons, or particulates. It is true that
indiscriminantly lowering excess air can have drastic effects on boiler CO
emissions, and that particulate emissions can increase with off
stoichiometric combustion and flue gas recirculation. However, with
suitable engineering during development and implementation of these
modifications, adverse incremental emissions problems can be minimized. In
contrast, residual emissions of sulfate, organics, and trace metals have
intermediate to high potential impact associated with applying almost every
combustion control. For trace metal and organic emissions, substantiating
data are largely lacking, but fundamental formation mechanisms give cause
for justifiable concern. Indeed, the 180 MW coal-fired utility boiler test
indicated a marked increase in organic emissions with off stoichiometric
combustion. In the case of sulfate emissions, fundamental formation
mechanisms suggest that these emissions should remain unchanged or decrease
with all controls except ammonia injection. Data from the recent utility
boiler test lend support to this hypothesis in the case of off
stoichiometric combustion. However, complex interactive effects are
difficult to elucidate, and sulfates are considered sufficiently hazardous
to justify expressing some concern in the present absence of conclusive
data. The potential effects of postcombustion ammonia injection on plume
sulfate formation deserve special attention.
The incremental emission evaluations of Table 8-14 are not intended
to signify any potential for adverse environmental impact. Rather, the
evaluation notes control/pollutant combinations for which emissions may
increase due to the use of NO controls. Evaluation of potential adverse
^
impact requires comparison of the source generated ambient pollutant
concentration with an upper limit threshold concentration of the pollutant
based on health or ecological effects. A preliminary attempt at such a
comparison has been made in Section 8.1.7.
In general, the data on incremental multimedia emissions due to NO
controls are still very sparse. More data are available for flue gas
emissions than for liquid or solid effluent streams. Even so, the only data
which allow quantified conclusions are for emissions of criteria pollutants
8-36
-------
TABLE 8-14. EVALUATION OF INCREMENTAL EMISSIONS DUE TO NOX CONTROLS
APPLIED TO BOILERS
NOX Control
Low Excess Air
Off
Stoichiometric
Combustion
Flue Gas
Reclrculatlon
Reduced Air
Preheat
Reduced Load
Hater
Injection
Amnonla
Injection
Incremental Emission
CO
•n-
0
0
0
0
0
0
Vapor Phase
HC
0
0
0
0
0
0
0
Sulfate
+
0
+
+
+
+
++
Participate
0
0
•f
0
0
+
+
Organ ics
++
++
+
+
•f
+
0
Segregating
Trace Metals
+
•»•
•i-
0
0
0
+
Nonsegregating
Trace Metals
0
0
t
+
0
0
0
00
CO
Key: ++ denotes having high potential emissions impact
+ denotes having Intermediate potential emissions impact, data needed
0 denotes having low potential emissions impact
-------
with the major control applications. Data on sulfates, trace metals, and
organics (POM) are sparse, experimentally uncertain and highly dependent on
fuel properties. Incremental emissions from liquid and solid effluent
streams and during transient or nonstandard operation are almost
nonexistent. Because of this, they have generally been excluded in the
present evaluation.
Emissions of CO, HC, particulate (smoke), and SO, with or without
NO controls have been constrained in the past for operational reasons
A
rather than environmental impact. CO, HC, and smoke emissions reduce
efficiency and may present a safety hazard. 50^ leads to acid
condensation and corrosion. All of these emissions are sensitive to
combustion process modifications for NO control. With the exception of
A
SO,, incremental emissions tend to increase with NO controls,
0 A
particularly low excess air and off stoichiometric combustion. Development
experience has shown, however, that with proper engineering these emissions
can generally be constrained under low NO conditions. This is
A
particularly true for factory-instailed controls on new equipment. In this
case, the flexibility for applying NO controls with minimal adverse
A
impact is greater than for retrofit on existing equipment. In light of this
situation, incremental emissions are seen more as a constraining criteria to
be addressed during control development than as an immutable consequence of
low NOX firing. Moreover, the constraint on emissions for satisfactory
operational performance is oftentimes more stringent than the constraint for
acceptable environmental impact.
The situation for other flue gas pollutants is more uncertain. There
is concern that conventional combustion process modifications -- low excess
air, off stoichiometric combustion, flue gas recirculation -- will increase
emissions of organics and segregating trace metals from sources firing coal
or residual oil. It should be noted, however, that this conclusion is based
on sparse data or, lacking that, on fundamental speculation. Clearly, more
data are needed.
In conclusion, there is reasonable concern that NO controls will
A
increase incremental emissions of some pollutants. More data are still
needed to determine if incremental emissions have a significant
environmental impact and to suggest corrective action if needed.
8-38
-------
8.2 ENERGY IMPACT
Changes in energy consumption with application of combustion
modification NO controls is one of many potential process impacts.
/\
Although these process impacts are reviewed in the next subsection, energy
impact is of such paramount importance (since it can account for up to half
of the cost-to-control) that it warrants a separate review.
The largest potential energy impact of combustion modifications is
their effect upon boiler thermal efficiency. Another significant source of
energy impact is the change in fan power requirements caused by these
controls. Boiler control systems installed for low NO operation also
increase electricity and instrument air requirements, but the energy impact
is usually minimal. Section 6 has already discussed on a boiler-by-boiler
basis the energy impacts of applied NOX controls. As noted there, with
proper engineering and implementation, there should be no major adverse
energy impacts with preferred combustion modifications. A review of that
analysis follows.
Applying low excess air (LEA) firing not only results in a small
decrease in NO emissions but also an increase in boiler efficiency
through reduced sensible heat loss out the stack. For this reason the
technique has gained acceptance and has become more a standard operating
procedure than a specific NO control method in both old and new units.
^
The other commonly applied combustion modifications, off
stoichiometric combustion (OSC) and flue gas recirculation (FGR), often lead
to decreases in boiler efficiency when implemented on a retrofit basis. Off
stoichiometric combustion usually increases excess air requirements
resulting in decreases in efficiency of up to 0.5 percent. Unburned fuel
losses either due to OSC or FGR may cause a decrease in efficiency of up to
0.5 percent. If a substantial increase in reheat steam attemperation is
required due to OSC or FGR, cycle efficiency losses of up to 1 percent may
occur. Increased fan power requirements due to OSC or FGR will also impact
efficiency, resulting in losses of up to 0.2 to 0.3 percent. No significant
energy impact is expected with low NOX burners (LNB), either retrofit or
new installation.
Ammonia injection requires energy for the injectors, NH3, handling
equipment, and carrier gas, resulting in an energy loss of about
0.25 percent. Moreover, the impact of increased ammonia consumption on the
8-39
-------
nationwide energy situation may be significant since ammonia is synthesized
primarily from natural gas. These impacts are discussed elsewhere
(Reference 8-37).
Other combustion modification techniques, water injection and reduced
air preheat, can impose quite significant energy penalties on boiler
operation, with decreases in efficiency from 5 to 10 percent. As a
consequence, these techniques are quite unpopular, and have found little
acceptance.
In summary, the decreases in boiler efficiency (increases in energy
consumption) discussed above for the preferred NO control techniques
J\
(OSC, FGR, and LNB) represent upper estimates when applied on a retrofit
basis. These same combustion modifications are not expected to adversely
affect unit efficiency when designed in as part of a new unit. This
illustrates that with proper engineering and development, combustion
modification NO controls can be incorporated into new unit designs with
A
no significant adverse energy impacts.
8.3 PROCESS IMPACTS
Low NO operation of utility boilers has been a source of concern
A
among utility plant operators due to potential adverse effects associated
with NO control techniques. The impact of combustion modifications on
^
boiler performance, operation, and maintenance has been discussed in detail
in the process analysis of Section 5. The major concerns are summarized
briefly below.
8.3.1 Efficiency
This potential impact has just been reviewed in the previous
subsection, which concluded that preferred NO controls should not cause
/\
significant adverse energy impacts on new units. However, for retrofit
applications, decreases in boiler efficiency, though minor, are of concern
because of the rapidly rising cost of fuel.
8.3.2 Corrosion
Corrosion is potentially a major problem with off stoichiometric
combustion (OSC) on coal-fired boilers because of possible local reducing
conditions when staging. Furnaces fired with certain Eastern U.S.
bituminous coals with high sulfur contents may be especially susceptible to
corrosion attack under reducing atmospheres. Tests with corrosion coupons
show wide scatter in data but generally indicate no significant increase in
8-40
-------
corrosion due to OSC. Current EPA-sponsored long-term tests on actual tube
walls should provide more definitive conclusions. There have been no
reports of corrosion in oil-fired boilers due to OSC. No corrosion problems
are expected with either oil- or gas-fired boilers.
8.3.3 Slagging and Fouling
In coal-fired equipment operating under OSC there has been some
concern regarding slagging. Slag usually fuses at a lower temperature under
reducing conditions. It was surmised that in certain cases molten, hard to
remove slag would form near the burners fired under fuel-rich conditions.
In the many tests conducted, however, no increase in slagging has been
noted. In oil-fired equipment also no increased fouling has been reported.
In gas-fired boilers fouling is a problem when switching from oil to gas as
the ash deposited on the walls during oil-firing causes reduced furnace heat
absorption and hence, increased furnace outlet gas temperatures.
8.3.4 Derating
Loss in boiler load capacity due to limited coal pulverizer capacity
will occur in many coal-fired boilers operated with burners-out-of-service
(BOOS). Derates of 10 to 25 percent may occur. For oil-fired boilers on
OSC, higher excess air requirements may cause fan capacity limits to be
reached in some cases. Although derates due to fan capacity are not common,
reductions of up to 15 percent have been reported. With OSC and flue gas
recirculation (FGR), excessive tube and steam temperatures may lead to
derating, expecially for gas-fired boilers, and in some cases for oil-fired
boilers. Derates of as high as 50 percent have been reported with
gas-firing immediately after switching from oil firing when the problem is
most severe.
8.3.5 Steam and Tube Temperatures
Excessive steam and tube temperatures may be encountered with oil-
and gas-firing when operated with OSC and FGR. The problem with tube
temperatures are expecially severe with units switched from oil to gas
firing. Increased tube failures may occur. Unless the furnace is
completely water washed to clean heat transfer surfaces before the switch,
derates of up to 50 percent may be required to prevent excessive tube
temperatures. Excessive reheat attemperation would necessitate removal of
some reheater surface in order to avoid a reduction in cycle efficiency.
Superheater surfaces may need to be removed if superheater attemperator
8-41
-------
capacities are exceeded. Ususally removal of reheater or superheater
surface must be accompanied by adding to the economizer surface if boiler
efficiency is to be maintained. No excessive steam and tube temperatures
have been reported with coal-firing.
8.3.6 Flame Instability and Vibrations
Problems with flame instability and furnace fan or duct vibrations
often occur with FGR operation on oil- and gas-fired boilers. Changes in
burner geometry and design are usually required to correct flame instability
and associated furnace vibration problems. Fan and duct vibration problems
may be avoided by careful design. In some cases, unit startup and load
pickup response will be altered due to FGR fan preheating requirements. No
instability or vibration problems have been reported with coal firing.
8.3.7 Particulates
On coal-fired boilers, particulate emissions may increase with OSC,
although there is wide scatter in the data. Increases are usually around 20
percent on the average, although numbers as high as 50 percent and 100
percent have been reported. Increase in particulates may also increase
erosion; but this should show up on corrosion tests, and as mentioned
earlier the results of of those tests have been inconclusive. No
significant change in particle size distribution has been observed with
OSC. With low NO burners (LNB) there may be a shift towards smaller
/\
particle sizes. An increase in particulate loading or number of smaller
particles may require installation of larger or more efficient particulate
collection devices. There is very little data on oil-fired boilers, but one
test has shown no significant change in particulate loading or size
distribution with OSC (Reference 8-10). With gas-fired boilers there should
be no problems with particulates.
8.3.8 Auxiliary Equipment
Implementation of low NO techniques often impacts the operation of
/\
boiler auxiliary equipment. OSC usually increases fan power requirements.
Average increases of 10 percent are reported if excess air requirements do
not increase substantially. If excess air rises significantly due to OSC
operation, F.D. or I.D. fan capacities may be reached. OSC and FGR also
affect superheater and reheater attemperation requirements. Again, in some
cases, spray attemperation flow limits may be reached. In coal-fired
boilers, OSC and LNB can result in increased carbon loss in flyash. Carbon
8-42
-------
loss may increase by as much as 130 percent. Increased carbon in flyash may
have an adverse impact on electrostatic precipitator operation due to
changes in flyash resistivity. However, in some tests conducted to measure
flyash resistivity, no change was noted due to low NO operation.
y\
8.3.9 Other Operational Impacts
Low NO operation can impact the safety and control aspects of a
^
boiler. Hazier flames and obscure flame zones associated with OSC firing on
oil- and gas-fired boilers ususally requires new flame scanners and
detectors. OSC firing also changes minimum air requirements which requires
appropriate combustion control modifications. Boilers may also often be
more prone to smoke or emit CO emissions under OSC firing requiring greater
operator attention. In some cases the boiler may require modified startup
procedures, e.g., FGR fan preheating and temperature change limitations.
8.3.10 Maintenance
As most low NO techniques are very sensitive to boiler conditions
rt
an accelerated maintenance and overhaul schedule may be necessary. Boiler
cleaning, burner tuning, checking fuel and air distributions, checking for
signs of tube wear or incipient failure, etc., may all need to be carried
out at regular intervals to maintain low NO operating conditions and
^
prevent serious problems prior to their occurrence.
8.3.11 Concluding Remarks
This subsection has highlighted some of the potential process impacts
of combustion modification NO controls. They are meant only as a guide
A
to control developers and users to aid in avoiding potential problems. Of
course, a particular boiler/control application may have none or only a few
of these problems. With proper engineering and implementation, potential
adverse process impacts can often be eliminated or minimized.
8.4 ECONOMIC IMPACT
Costs are particularly important in regulated utility economics,
especially because all "allowable" costs of doing business are permitted to
be recovered from the consumer. Not only is a utility concerned with the
impact of a pollution control on the final cost of electricity but also on
the impact of the initial outlay of capital. The public utility sector is
characterized by the necessity for large aggregations of capital because the
enterprises typically require high initial investment costs (Reference 8-38).
8-43
-------
Section 7 analyzed costs in detail of several representative
applications of combustion modification NO controls, both retrofit and
J\
new unit application. The following dicussion summarizes that study.
8.4.1 Retrofit Control Costs
Analysis of retrofit control costs is important as there is often a
need for controlling NO from existing boilers, as part of State
/v
Implementation Plans, in response to specific aspects of the Clean Air Act
(e.g., the emissions offset program for nonattainment areas). Besides, most
of the existing data base on combustion modification NO controls is from
/\
retrofit demonstrations. Thus, retrofit analysis should provide a good
estimate for the cost-to-control, as it is expected that factory installed
controls, properly engineered, should cost less.
Table 8-15 lists the representative boiler/retrofit control
combinations costed in this study. It was assumed that the units being
retrofitted were relatively new, say 5 to 10 years old, with at least 25
years of service remaining. As Table 8-15 shows, overfire air and low NO
A
burners were selected as the retrofit control methods for coal-firing.
Burners out of service was not necessarily recommeded for coal-fired units,
but was included to demonstrate the high cost of derating a unit, as is
often the case for pulverized coal units. Burners out of service, and flue
gas recirculation through the windbox combined with overfire air were
selected as the retrofit control methods for a single wall oil-and gas-fired
unit.
Estimated costs for applying the treated NO controls, in 1977
/\
dollars, are summarized in Table 8-16. The table shows initial capital
investment, annualized capital investment with other indirect costs,
annualized direct costs, and total annualized cost to control. The table
indicates that the preferred combustion modification generally costs between
$0.50 to 0.70/kW-yr to install and operate. One major exception to this is
the use of BOOS firing on coal-fired units if derating is required due to
insufficient mill capacity. In this instance the high cost of BOOS
implementation reflects the need to purchase makeup power, and to account
for lost capacity (a 20 percent derate is typical) through a lost capital
charge.
8-44
-------
TABLE 8-15. BOILER/RETROFIT CONTROL COMBINATIONS COSTED
Boiler Fuel
Tangential/Coal
Opposed Wall /Coal
Opposed Wall /Coal
Opposed Wall /Coal
Single Wall/Oil, Gas
Single Wall/Oil, Gas
MCRa
(MW)
225
540
540
540
90
90
NOX
Control
OFA
OFA
LNB
BOOS
BOOS
OFA & FGR
aMaximum continuous rating in MW of electrical output
8-45
-------
TABLE 8-16. SUMMARY OF RETROFIT CONTROL COSTS (1977 DOLLARS)'
Boiler/Fuel Type
Tangent i al /Coal -F i red
OFA
Opposed Wall /Coal -Fired
OFA
LNB
BOOSC
Single Wall/Oil- and Gas-Fired
BOOS
FGR/OFA
Initial
Investment
(SAW)
0.90
0.62
2.03
0.08
0.30
5.71
Annualized Indirect
Operating Cost
($/kW-yr)
0.21
0.16
0.34
5.34
0.05
1.14
Annualized Direct
Operating Cost
($/kW-yr)b
0.32
0.52
0.06
24.78
0.44
1.91
Total to Cost
Control
($/kW-yr)b
0.53
0.69
0.40
30.12
0.49
3.05
00
I
aBased on assumptions given in Section 7 and cost input parameters listed in Appendix E.
bBased on 7000 n operating year. Typical costs only.
cAssumes twenty percent derate required.
-------
8.4.2 Control Costs for New Units
Estimating the incremental cost of NO controls for NSPS boilers is
A
in some respects an even more difficult task than costing retrofits.
Certain modifications on new units, through effective in reducing NO
A
emissions, were originally incorporated due to operational considerations
rather than from a control viewpoint. For example, the furnace of a typical
unit designed to meet 1971 NSPS has been enlarged to reduce slagging
potential. But this also reduces NO due to the lowered release rate.
Thus, since the design change would have been implemented even without the
anticipated NO reduction, the cost of that design modification should not
A
be attributed to NO control.
A
Babcock & Mil cox and Foster Wheeler have estimated the cost of
preferred NO controls for new coal-fired boilers using low NO burners
A A
and overfire air. Both manufacturers indicate incremental costs in the
$1.75 to $2.80/kW range, or $0.28 to 0.43/kW-yr annualized, for a typical
NSPS boiler. These costs are discussed in detail in Section 7.
8.4.3 Cost Effectiveness of Controls
Combustion modifications represent cost-effective, demonstrated means
of NO control for utility boilers, reducing NO emissions 20 to 60
/> ' X
percent at relatively low cost, usually less than 1 percent of the cost of
electricity. Furthermore, the initial capital cost is usually less than 1
percent of the cost of the boiler. Table 8-17 summarizes projected control
requirements for alternative NO emission levels. Control requirements
A
are recommended to achieve a given NO emission level. These control
A
levels combined with the cost of control column, complete the
cost-effectiveness picture.
Compared to the $0.30 to $0.50/kW-yr cost of preferred combustion
modification controls for new coal-fired boilers, alternative NO control
A\
techniques, ammonia (NH,) injection and selective catalytic reduction
(SCR), neither of which represent demonstrated technology, are projected to
cost significantly more: $2.50 to $3.40/kW-yr for NH3 injection
(Reference 8-37), and $15 to $25/kW-yr for SCR (Reference 8-39). However,
these latter two techniques have the potential for achieving lower NO
emission levels from coal-firing, 129 ng/J (0.3 lb/106 Btu) for NH3
injection, and 43 ng/J (0.1 lb/106 Btu) for SCR. These control levels
assume that combustion modifications are already applied.
8-47
-------
TABLE 8-17. PROJECTED CONTROL REQUIREMENTS FOR ALTERNATE
NOX EMISSION LEVELS
Fuel/N0x Emission Level:
ng/J (lb/106 Btu)
Coal
301 (0.7)
258 (0.6)
215 (0.5)
172 (0.4)
Oil
129 (0.3)
86 (0.2)
Gas
129 (0.3)
86 (0.2)
43 (0.1)
Recommended Control
Requirement3
OFAC
OFAC
LNB
OF A + LNB
BOOS
FGR + OFA
BOOS
FGR + OFA
FGR + OFA
Cost to Control :
$/kW-yrb
Retrofit New Boiler
0.50 to 0.70
0.50 to 0.70
0.40 to 0.50
0.95 to 1.20
0.50 to 0.60
3.00
0.50 to 0.60
3.00
3.00
0.10 to 0.20
0.10 to 0.20
0.30 to 0.40
0.40 to 0.50
N/A d
d
N/A
aLEA considered standard operating practice.
bTypical installation only; could be significantly higher. 1977 dollars.
cAs manufacturers acquire more experience with LNB, they are now
recommending LNB over OFA.
dN/A - Not applicable, no new oil- or gas-fired boilers being sold.
8-48
-------
Advanced combustion modification concepts under development, such as
the EPA advanced low NO burner (Reference 8-40) and EPRI primary
A
combustion furnace (Reference 8-41), are targeted to achieve NO emissions
6
levels below 86 ng/J (0.2 lb/10 Btu) on a commercial basis in the
1980's. Projected cost for the EPRI furnace is $5/kW or $0.80/kW-yr
(Reference 8-42). The EPA advanced burner costs should fall in the same
range between conventional combustion modification costs and the EPRI
furnace costs (Reference 8-43). Thus developing advanced combustion
modifications should eventually prove much more cost-effective than the
developing postcombustion techniques; however, the latter techniques are
currently closer to commercialization.
8.4.4 Concluding Remarks
Use of combustion modification NO controls should have no major
adverse economic impact on the boiler manufacturing or utility industry.
The four major boiler manufacturers all offer competitive designs of the
preferred techniques for new boilers: off stoichiometric combustion and/or
low NO burners (Reference 8-44). And the relatively low cost of
/\
combustion modifications, combined with accumulating favorable experience
with their application, should aid in their acceptance by the utility sector
(Reference 8-42).
8.5 EFFECTIVENESS OF N0v CONTROLS
/\
The effectiveness of combustion modifications NO controls has been
/\
examined in detail in Sections 4 through 6, with a detailed cost analysis in
Section 7. This subsection highlights the major controls.
8.5.1 Coal-Fired Boilers
The most commonly applied low NO technique for coal-fired boilers
A
is off stoichiometric combustion (OSC) through overfire air (OFA).
Application of burners out of service (BOOS), an alternate staging
technique, is limited because it is often accompanied by a 10 to 25 percent
load reduction. Average NO reductions of 30 to 50 percent (controlled
emissions of 215 to 301 ng/J, 0.5 to 0.7 lb/106 Btu) can be expected with
either technique. Flue gas recirculation (FGR) has been tested, but was
found to be a relatively ineffective control, giving only about 15 percent
N0¥ reduction. More recently, new low NOV burners (LNB) have been
A X
installed on some units and have been found to be at least as effective as
OFA. The combination of OFA with LNB has resulted in 40 to 60 percent NO
8-49
-------
reductions (controlled emissions of 172 to 215 ng/J, 0.4 to 0.5 lb/106
Btu).
There has been a steady improvement in combustion modification
control technology over recent years. Figure 8-4 conceptually reviews the
past, current, and projected development of major controls. As shown,
current demonstrated technology is capable of 40 to 60 percent NO
/\
reductions, readily meeting the current New Source Performance Standard
(NSPS) of 258 ng/J (0.6 lb/106 Btu) for bituminous coal and 215 ng/J
(0.5 lb/10 Btu) for subbituminous coal. Current R&D programs, such as the
EPA advanced low NO burner and the EPRI primary combustion furnace,
/\
should result in combustion modification techniques capable of meeting
projected future NOX emission control levels (1980's) of 86 ng/J
(0.2 lb/106 Btu) to 129 ng/J (0.3 lb/106 Btu).
8.5.2 Oil-Fired Boilers
The most commonly used low NO techniques for oil-fired boilers are
J\
off stoichiometric combustion and flue gas recirculation (FGR), both
employed with low excess air firing. Other techniques which have been
tested are water injection (WI) and reduced air preheat (RAP). However,
these latter two techniques have found little application due to attendant
efficiency losses.
Off stoichiometric combustion has been applied through the use of
overfire air ports (OFA) and by removing burners from service (BOOS).
Typical NO reductions using OFA are 20 to 30 percent (controlled
emissions of 150 to 172 ng/J, 0.35 to 0.4 lb/106 Btu), while BOOS has been
slightly more effective giving 20 to 40 percent reductions (controlled
levels of 129 to 172 ng/J, 0.3 to 0.4 lb/106 Btu). Flue gas recirculation
also typically gives 20 to 30 percent NO reductions, but requires more
«
hardware modifications. The combination of BOOS or OFA with FGR has been
most effective, resulting in 30 to 60 percent reductions (controlled
emissions of 86 to 172 ng/J, 0.2 to 0.4 lb/106 Btu). With FGR, OFA is
preferred over BOOS because flame stability is expected to be more of a
problem with the combination of FGR + BOOS.
There has been some R&D effort by EPA and private industry on low
NOX emission burners for oil-firing. One manufacturer has reported the
successful retrofit of an oil-fired low NO burner, producing NO
emissions of below 129 ng/J (0.3 lb/106 Btu) (Reference 8-45). The
8-50
-------
NO Emission Level
Percent Reduction 1970
1975
1980
1985
X •
430 ng/J (1.0 Ib/lO^ Btu) 0 IZZTBaseline
777"? -* — Enlarged
345 (0.8) 20 -- Lu. ^ V V V \ -»-Bias
?fin in &\ /in .. v\\\ « /////
tlb lU.j; jU --
170 (0.4) 60 —
00
i, 85 (0.2) 80 --
•-»
furnace, low excess air
ed burner firing
•«-Overfire air or low NO burners
,\\\\\
t\\N
-------
combination of overfire air and low emissions burners may potentially
achieve emissions below 86 ng/J (0.2 lb/10 Btu).
8.5.3 Gas-Fired Boilers
The most commonly applied NOX control techniques for gas-fired
boilers, as with oil-fired boilers, are staged combustion through the use of
OFA or BOOS with FGR; however, flame stability may be of greater concern
when FGR is combined with BOOS. Typical NO reduction under either OFA,
/\
BOOS, or FGR are 30 to 60 percent (controlled emissions of 86 to 150 ng/J,
0.2 to 0.35 lb/10 Btu). The combination of staged combustion and FGR is
capable of 50 to 80 percent reductions (controlled levels of 43 to 108 ng/J,
0.1 to 0.25 lb/106 Btu).
There are no major efforts toward developing a low NO burner or
A
other new combustion modification techniques for gas-firing because NO
A
emissions under current control techniques are already relatively low, and
no new gas-fired utility boilers are being sold currently.
8.6 CONCLUSIONS AND RECOMMENDATIONS
Combustion modification NO controls are cost-effective techniques,
A
causing no apparent major adverse environmental impacts. It is recommended
that data acquisition from long term NO control applications continue, in
A
order to eliminate potential areas of concern and optimize boiler
performance.
8.6.1 Conclusions
Modifying the combustion process conditions is currently the most
cost-effective and best demonstrated method of effecting 20 to 60 percent
reductions in NO emissions from utility boilers. Table 8-18 summarizes
A
the capabilities of combustion modification NO controls. The methods in
A
the best available control technology (BACT) and advanced technology
categories are listed in preference of application. They were selected
based on an assessment of their effectiveness (Sections 4 through 6),
operational (Section 6), energy (Section 6), cost (Section 7), and
environmental (Section 8) impact, and commercial availability or R&D status
(Section 4).
In the BACT category, low NOV burners (LNB) or off stoichiometric
A
combustion (OSC) through overfire air addition (OFA) are the preferred
techniques for retrofit application to coal-fired units. The actual choice
would be determined on a site-specific basis, depending on the fuel/furnace
8-52
-------
TABLE 8-18. COMBUSTION MODIFICATION NOX CONTROLS: BEST AVAILABLE
CONTROL TECHNOLOGY (BACT) AND ADVANCED TECHNOLOGY
BACT
Advanced
Technology
Fuel
Coal
Oil
Gas
Coal
Oil
Control Technique
Overfire aira
Low NOX burners
Low NOX burners plus
overfire air
Burners out of service or
overfire air
Flue gas recirculation plus
overfire air
Burners out of service or
overfire air
Flue gas recirculation plus
overfire air
Ammonia injection (1983)^
(combined with BACT
combustion modifications)
Advanced low NOX burners
(1985)
Advanced burner/furnace
concepts (1985)
Ammonia injection (1983)
(combined with BACT
combustion modifications)
NOX Control Level,
ng/J (lb/106 Btu)
258 (0.6)
215 (0.5)
172 (0.4)
129 (0.3)
86 (0.2)
129 (0.3)
43 (0.1)
129 (0.3)
86 (0.2)
60 (0.15)
43 (0.1)
aAs manufacturers acquire more experience with LNB, they are now
recommending LNB over OFA.
Estimated date of commercial availability of demonstrated technology.
8-53
-------
design, etc. The use of low NOX burners, or low NOX burners in
combination with OFA is favored. For new units, off stoichiometric
combustion through OFA or removing burners from service, flue gas
recirculation, or the combination of F6R with OSC is recommended for
retrofit application to oil- and gas-fired boilers. No sales of new oil- or
gas-fired units are projected.
While BACT can achieve 172 ng/J (0.4 lb/10 Btu) for coal-firing,
86 ng/J (0.2 lb/106 Btu) for oil-firing, and 43 ng/0 (0.1 lb/106 Btu)
for gas-firing, Table 8-18 indicates that advanced techniques have the
potential of reducing NO to 86 ng/J (0.2 lb/10 Btu) for coal and 43
fi
ng/J (0.1 lb/10 Btu) for oil. However, ammonia injection, advanced low
NO burners, and advanced burner/furnace concepts are several years away.
A
Ammonia injection is considered a near-term intermediate control option
between BACT and the more distant advanced concepts, intermediate from the
point of view of control effectiveness and availability. However, ammonia
injection has many potential operational and environmental hazards that need
to be assessed, as discussed in Section 4, as well as much higher projected
costs than either BACT or the more promising advanced concepts, as noted in
Section 7.
The use of conventional combustion modifications (BACT) has potential
for adverse effects on boiler efficiency, load capacity, water wall tube
corrosion, slagging, fouling, carbon loss, steam temperature, flame
stability, and vibration. However, recent field experience has shown that
adverse effects can be minimized to acceptable levels with proper care in
design for retrofit application, and largely eliminated in new unit designs.
Another area of concern with combustion modification NO controls
/\
is a possible increase in incremental emissions of other pollutants to the
environment. Recent test data with BACT techniques seem to indicate that
low NOX firing has negligible effects on emissions of most pollutants
other than N0x. Based on the comprehensive environmental assessment test
run on a 180 MW boiler, low NO firing does indeed lower the overall
A
potential environmental impact of the source. However, there are areas of
continued concern, such as possible increased organic emissions. More
extensive field testing will be required to identify and better quantify
these emissions, and compare these results with developing information in
the health effects area.
8-54
-------
Finally, conventional combustion modifications are indeed
cost-effective means of control for NO , raising the cost of electricity
less than one percent in most cases. Furthermore, the initial capital
investment required should also only be of the order of 1 percent or less of
the installed cost of a boiler. With the exception of NH^ injection,
advanced techniques such as advanced low NO burners and advanced
/\
burner/furnace concepts have projected costs in the same range as
conventional combustion modifications. Therefore, preferred current and
projected combustion modification techniques are not expected to have any
significant adverse economic impact.
8.6.2 Recommendations
Preferred conventional combustion modifications are indeed
recommended for reducing NO emissions from utility boilers, with minimal
/\
adverse environmental, operational, and cost impacts. However, long-term
testing and monitoring of field applications/demonstrations should be
continued. Although the issue of possible increased corrosion with off
stoichiometric combustion has been largely resolved in short-term tests,
long-term corrosion testing, as under current EPA programs, should be
completed to definitively establish that low NO firing does not have any
^
adverse effects. Boiler efficiency should be closely monitored during field
applications to give guidance to control developers on minimizing or
eliminating efficiency losses. The current data base indicates that
efficiency losses of up to 0.5 percent are possible. The exact number is of
significance; for example, a 0.25 percent loss in efficiency can translate
to one-third of the annualized cost to control.
Finally, the data gaps on the effect of NO controls on incremental
rt
emissions are just now beginning to be addressed. Field testing on
representative utility boiler/control applications should continue, with
special emphasis on incremental emissions such as trace metals and organics.
Research and development efforts on new combustion modification
technology, such as advanced staged combustion, low NO burners and
X
burner/furnace concepts, should continue since they have the potential of
further NOX reduction capabilities with minimal adverse impacts.
8-55
-------
REFERENCES FOR SECTION 8
8-1. Schalit, L. M., and K. J. Wolfe, "SAM IA: A Rapid Screening
Method for Environmental Assessment of Fossil Energy Process
Effluents," EPA-600/7-78-051, NTIS-PB 277 088/AS, February 1978.
8-2. Hangebrauck, R. P., et a!., "Nomenclature for Environmental
Assessment Projects: Part 1 — Terminology for Environmental
Impact Analysis," EPA IERL-RTP, Research Triangle Park, NC,
August 1979.
8-3. Higginbotham, E. B., and P. M. Goldberg, "Field Testing of a
Tangential Coal-Fired Utility Boiler — Effects of Combustion
Modification NOX Controls on Multimedia Emissions," Acurex Draft
Report No. 79-337, EPA Contract No. 68-02-2160, Acurex
Corporation, Mountain View, CA, April 1979.
8-4. Crawford, A. R., et al., "The Effect of Combustion Modification on
Pollutants and Equipment Performance of Power Generation
Equipment," in Proceedings of theStationary Source Combustion
Symposium. Volume III. EPA-600/2-76-152C. NTIS-PB 257 146/AS.
June 1976.
8-5. Crawford, A. R., et al., "Field Testing: Application of
Combustion Modifications to Control NOX Emissions for Utility
Boilers," EPA-650/2-74-066, NTIS-PB 237 344/AS, June 1974.
8-6. Hollinden, G. A., et al., "Evaluation of the Effects of Combustion
Modifications in Controlling NOX Emissions at TVA's Widow's
Creek Steam Plant," in The Proceedings of the N0y Control
Technology Seminar, EPRI SR-39, February 1976.
8-7. Mason, H. B., et al., "Preliminary Environmental Assessment of
Combustion Modification Techniques: Volume II, Technical Results,"
EPA-600/7-77-119b, NTIS-PB 276 681/AS, October 1977.
8-8. Thompson, R. E., et al., "Effectiveness of Gas Recirculation and
Staged Combustion in Reducing NOX on a 560 MW Coal-Fired
Boiler," EPRI FP-257, Electric Power Research Institute, Palo
Alto, CA, September 1976.
8-9 Kamo, R., et al., "The Effect of Air-Fuel Mixing on Recirculation
in Combustion," Paper CP-62-12, API Research Conference on
Distillate Fuel Consumption, June 1962.
8-10. Crawford, A. R., et al., "Field Testing: Application of
Combustion Modification to Power Generating Combustion Sources,"
in Proceedings of the Second Stationary Source Combustion
Symposium. Volume II, EPA-600/7-77-073b. NTIS-PB 271 756.
July 1977.
8-56
-------
8-11. Klein, D. H., et al., "Pathways of Thirty-Seven Trace Elements
Through Coal-Fired Power Plant," Environmental Science and
Technology. Volume 9, No. 10, pp 973-979, October 1975.
8-12. Ray, S. S., and F. G. Parker, "Characterization of Ash from
Coal-Fired Power Plants," EPA/7-77-010, NTIS-PB 265 374,
January 1977.
8-13. Davison, R. L., et al., "Trace Elements in Fly Ash," Environmental
Science and Technology, Volume 8, No. 13, pp. 1107-11131
December 1974.
8-14. Kaakinen, J. W., et al., "Trace Element Behavior in Coal-Fired
Power Plant," Environmental Science and Technology, Volume 9,
No. 9, pp. 862-869, September 1975.
8-15. Cato, G. A., and R. A. Venezia, "Trace Metal and Organic Emissions
of Industrial Boilers," Paper 76-27.8, 69th Annual APCA Meeting,
June 1976.
8-16. "Coal-Fired Power Plant Trace Element Study, Volume I, A Three
Station Comparison," Report for EPA Region VIII, Radian Corp.,
Austin, TX, September 1975.
8-17. Gladney, E. S., et al., "Composition and Size Distributions of
Atmospheric Particulate Matter in Boston Area," Environmental
Science and Technology, Volume 8, No. 6 p. 551, June 1974.
8-18. Ensor, D. S., et al., "Elemental Analysis of Fly Ash from
Combustion of a Low Sulfur Coal," Paper 75-33.7, 68th Annual APCA
Meeting, June 1975.
8-19. Lee, R. E., Jr., et al., "National Air Surveillance Cascade
Impactor Network II: Size Distribution Measurements of Trace
Metal Components," Environmental Science and Technology. Volume 6,
No. 12, pp. 1025-1030, November 1972.
8-20. Bolton, N. E., et al., "Trace Element Measurements at the
Coal-Fired Allen Steam Plant," Progress Report, February 1973
through July 1973, ORNL-NSF-EP-62, 1974.
8-21. Billings, C. E., et al., "Mercury Balance on a Large Pulverized
Coal-Fired Furnace," J. APCA. Volume 23, No. 9 pp. 773-777,
September 1973.
8-22. Cato, G. A., "Field Testing: Trace Element and Organic Emissions
from Industrial Boilers," EPA-600/2-76-086b, NTIS-PB 261 263/AS,
October 1976.
8-23. Vitez, B., "Trace Elements in Flue Gases and Air Quality
Criteria," Power Engineering, Volume 80, No. 1, pp. 56-60,
January 197(T
8-57
-------
8-24. Hegg, D. A., et al., "Reactions of Nitrogen Oxides, Ozone, and
Sulfur in Power Plant Plumes," EPRI EA-270, September 1976.
8-25. Richards, J. and R. Gerstle, "Stationary Source Control Aspects of
Ambient Sulfates: A Data Base Assessment," PEDCo Final Report,
EPA Contract No. 68-02-1321, Task 34, PEDCo Environmental,
Cincinnati, OH, February 1976.
8-26. Bennett, R. L., and K. T. Knapp, "Chemical Characterization of
Parti cu late Emissions from Oil Fired Power Plants," presented at
the 4th National Conference on Energy and the Environment,
Cincinnati, OH, October 1976.
8-27. Homolya, J. B., et al., "A Characterization of the Gaseous Sulfur
Emissions from Coal- and Coal-Fired Boilers," presented at the 4th
National Conference on Energy and the Environment, Cincinnati, OH,
October 1976.
8-28. Crawford, A. R., et al., "Control of Utility Boiler and Gas
Turbine Pollutant Emissions by Combustion Modification --
Phase I," EPA-600/7-78-036a, NTIS-PB 280 078/AS, March 1978.
8-29. "Position Paper on Regulation of Atmospheric Sulfates,"
EPA-450/2-75-007, NTIS-PB 245 760, NTIS-PB 245 760, September 1975.
8-30. Off en, G. R., et al., "Control of Particulate Matter from Oil
Burners and Boilers," EPA-450/2-76-005, NTIS-PB 258 495,
April 1976.
8-31. Remeysen, J., "Operations of Large Boilers at Very Low Excess-Air
Levels," Paper 1 in Current Development in Fuel Utilization, the
Institute of Fuel, 1964.
8-32. Niepenberg, H., "Combustion Control of Oil -Firing Systems Operated
at Low Excess Air Levels," Paper 5 in Third Liquid Fuels
Conference; Applications of Liquid Fuels. The Institute of Fuel,
-_-
8-33. Jackson, P. J., "Generating Stations Efficiencies," Paper 8 in
Third Liquid Fuels Conference; Applications of Liquid Fuels, The
Insititute of Fue, 1966.
8-34. "Chemistry and Metallurgy," Volume 5 in Modern Power Station
Practice. Central Electricity Generating Board, Pergamon Press,
New York, 1971.
8-35. Waterland, L. R., and L. B. Anderson, "Source Analysis Models for
Environmental Assessment," presented at Fourth Symposium on
Environmental Aspects of Fuel Conversion Technology, Hollywood,
FL, April 17, 1979.
8-58
-------
8-36. del and, J. 6. and G. L. Kingsburg, "Multimedia Environmental
Goals for Environmental Assessment," EPA-600/7-77-136a and b,
NTIS-PB 276 919 and 920, November, 1977.
8-37. Castaldini, C., et al., "Technical Assessment of Thermal DeNOx
Process," EPA-600/7-79-117, May 1979.
8-38. Kaufman, et al., "The Electricity Utility Sector: Concepts,
Practices and Problems," Congressional Research Service, Committee
Print No. 95-14, U.S. Government Printing Office, May 1977.
8-39. Maxwell, D., Tennessee Valley Authority, Muscle Shoals, Alabama,
Personal Communication, February 1979.
8-40. Martin, G. B., "Field Evaluation of Low NOX Coal Burners on
Industrial and Utility Boilers," in Proceedings of the Third
Stationary Source Symposium; Volume 'I. EPA-600/7-79-050a.
February 1979.
8-41. Johnson, S. A., et al., "The Primary Combustion Furnace System —
An Advanced Low-N0x Concept for Pulverized Coal Combustion,"
presented at Second EPRI NOX Control Technology Seminar, Denver,
Colorado, November 9, 1978.
8-42. Teixeira, D., "NOX Control Technology," EPRI Journal, Volume 3,
No. 9, pp. 37, November 1978.
8-43. Martin, G. B., EPA/IERL-RTP, Research Triangle Park, NC, Personal
Communication, August 1979.
8-44. Goodwin, D. R., "Electric Utility Steam Generating Units.
Background Information for Proposed NOX Emission Standards,"
EPA-450/2-78-005a, July 1978.
8-45. Barsin, J. A., "Pulverized Coal-Firing NOX Control," in
Proceedings; Second NOX Control Technology Seminar. Electric
Power Research Institute, EPRI FP-1109-SR, Palo Alto, CA,
July 1979.
8-59
-------
TECHNICAL REPORT DATA
(Please read Instructions on the reverse before completing)
1 REPORT NO.
EPA-600/7-80-075a
2.
3. RECIPIENT'S ACCESSION NO.
4. TITLE ANOSU8TITLE
Environmental Assessment of Utility Boiler
Combustion Modification NOx Controls:
Volume 1. Technical Results
5. REPORT DATE
April 1980
6. PERFORMING ORGANIZATION CODE
7. AUTHORIS)
K.J.Lim, L.R. Water land, C. Castaldini, Z.Chiba,
and E. B. Higginbotham
8. PERFORMING ORGANIZATION REPORT NO
TR-78-105
10. PROGRAM ELEMENT NO.
EHE624A
VI. CONTRACT/GRANT NO.
68-02-2160
0. PERFORMING OROANIZATION NAME AND ADDRESS
Acurex/Energy and Environmental Division
485 Clyde Avenue
Mountain View, California 94042
12. SPONSORING AGENCY NAME AND ADDRESS
EPA, Office of Research and Development
Industrial Environmental Research Laboratory
Research Triangle Park, NC 27711
13. TYPE OF REPORT AND I
Final: 3/77-5/78
ID PERIOD COVERED
14. SPONSORING AGENCY CODE
EPA/600/13
,5. SUPPLEMENTARY NOTESIERL_RTP project officer is Joshua S. Bowen, Mail Drop 65, 919/
541-2470.
ye. ABSTRACT ^^ report gives results of an evaluation of combustion modification tech-
niques for coal-, oil-, and gas-fired utility boilers, with respect to NOx control re-
duction effectiveness, operational impact, thermal efficiency impact, capital and
annualized operating costs, and effect on emissions of pollutants other than NOx.
For gas- and oil-fired boilers, 30 to 60% NOx reductions are achievable with the
combined use of staged combustion, flue gas re circulation, and low excess air at an
annualized cost of #0. 50 to #3. OOAW-yr. For retrofit control of existing coal-fired
boilers, low NOx burners and/or staged combustion yields a 30 to 60% NOx reduction
at an annualized cost of $0.40 to #1.20AW-yr. For new sources, modified furnace
design with low NOx burners and/or overfire air can achieve emission levels of 260
to 170 ng/J (40 to 60% reduction). Detailed emission tests on a 200 MW coal-fired
boiler showed that changes in trace specie emissions due to combustion modifications
were small compared to the benefit of reduced NOx emissions.
KEY WORDS AND DOCUMENT ANALYSIS
DESCRIPTORS
b. IDENTIFIERS/OPEN ENDED TERMS
c. COSATI Field/Croup
Air Pollution
Assessments
pmbustion Control
ttrogen Oxides
Hers
Utilities
Cost Effectiveness
Fossil Fuels
Dust
Aerosols
Trace Elements
Organic Compounds
<•. oiftYAiftuYiON ATATIMINT
RELEASE TO PUBLIC
Air Pollution Control
Stationary Sources
Utility Boilers
Combustion Modification
Particulate
Environmental Assess-
m
13B
14B
21B
07B
ISA
14A
21D
11G
07D
06A
07C
CLAM
-------