Sunteți pe pagina 1din 24

Pergamon

PII: S03601285(97)000221

Prog. Energy Combust. Sci. Vol. 24, pp. 385408, 1998


q 1998 Elsevier Science Ltd
Printed in Great Britain. All rights reserved
03601285/98 $19.00

NO x CONTROL THROUGH REBURNING*


L. D. Smoot a , b, S. C. Hill b and H. Xu a
b

a
Chemical Engineering Department, Brigham Young University, Provo, UT 84602, U.S.A.
Advanced Combustion Engineering Research Center, Brigham Young University, Provo, UT 84602, U.S.A.

AbstractReburning is a process whereby a hydrocarbon fuel is injected immediately downstream of the


combustion zone to establish a fuel-rich zone in order to convert nitric oxide to HCN. The reburning fuel can be
gaseous (e.g., natural gas), solid (e.g., coal char or wood) or liquid (e.g., residual oil). Typically, the amount of
reburning fuel used is 1030% of the total fuel. This technology is practiced commercially with nitric oxide
reduction levels of 3565%, depending on the type and scale of the boiler or combustion, the primary and
reburning fuels and other variables. Current research and development are suggesting several advanced reburning
concepts including injection of ammonia or urea aft of the reburning fuel injection. Nitric oxide reductions of over
90% are anticipated. In this mini-review, a review of reburning technologies, measurements and mechanisms is
presented. Predictive methods for reburning are also discussed. Recent work on reburning, including development
of a global reburning reaction rate, is summarized, and results of application of a comprehensive combustion
model to reburning measurements are summarized. q 1998 Elsevier Science Ltd. All rights reserved.
Keywords: advanced reburning, combustion modeling, nitric oxide, NO x reduction, reburning, global model.
CONTENTS

Nomenclature
1. Introduction
2. Technologies
3. Measurements
4. Mechanisms and Rates
5. Predictive Methods
6. Reburning Model Applications
6.1. Reaction Parameters
6.2. Particle Size
6.3. Reburn Gas Composition
6.4. Injection Velocity
6.5. Comparisons with Data
7. Summary
Acknowledgments
References

NOMENCLATURE

385
386
386
389
392
394
398
399
400
400
403
405
405
405
405

Gas temperature

Time

Xi

Concentration of species i

Equivalence ratio

Pre-exponential factor of the global reburning rate

Exponent on oxygen concentration or temperature

d mm

Mass mean particle size

Activation energy of global reburning rate

fi

Coefficients in Eqs (26)(29)

ABBCE

ABB Combustion Engineering Co.

ki

Rate constant of reaction i

ACERC

Advanced Combustion Engineering Research


Center

Gas law constant

B&W

Babcock and Wilcox Company

ri

Rate of reaction i

CAA

Clean Air Act Amendment

CARS

Coherent anti-Stokes Raman spectroscopy

CCT

Clean Coal Technology

CFD

Computational fluid dynamics

CHEMKIN

Chemical kinetics code package developed by


Sandia National Laboratory

* This mini-review paper was presented, together with a


series of other review papers, at the Tenth Annual Technical
Conference of the Advanced Combustion Engineering Research
Center, held in Salt Lake City, Utah, in March 1997.
Professor, Chemical Engineering; Director, ACERC.
Research Associate; Head, Combustion Computations
Laboratory.
Graduate Research Assistant, Chemical Engineering.

Abbreviations and acronyms

385

386

L. D. Smoot et al.

EER

Energy and Environmental Research


Corporation

GRI

Gas Research Institute

HiPPS

High performance power system

LDA

Laser doppler anemometry

LNB

Low NO x burner

MCRD

Micronized coal reburning demonstration

MSW

Municipal solid waste

MWe

Megawatt, electric

MWt

Megawatt, thermal

NYSEG

New York State Electric and Gas Co.

PCGC-3

Pulverized coal gasification and combustion in


three dimensions

PETC

Pittsburgh Energy Technology Center

ppm

Parts per million, by volume

REI

Research Engineering International Company

SCR

Selective catalytic reduction

SNCR

Selective non-catalytic reduction


1. INTRODUCTION

Nitrogen oxides (NO x) have been recognized as acid


rain precursors that impose a significant threat to the
environment. Coal combustion is a major anthropogenic
source of NO x. In coal combustion, NO x originates
mostly from nitrogen bound in the coal matrix, namely,
fuel-NO x1. Molecular nitrogen in air typically contributes less than 10% of overall NO x emissions from coal
combustion2. Several technologies have been demonstrated to reduce NO x emissions during fossil fuel
combustion, as summarized by Boardman and Smoot3
and others. During the past decade, work on NO x
formation and control has been substantial, including
gas-phase and coal-phase NO x formation modeling and
measurement, demonstration of very low NO x burner
technologies, and prediction and measurement of NO x
control through reburning. International work in NO x
formation and control has been particularly extensive
during this decade, as noted in reviews by Miller and
Bowman4, Hayhurst and Lawrence5, Bowman6, and
Kramlich and Linak7. Among the most recent developments for reducing NO x emissions from coal systems are
the reburning technologies, wherein gaseous, liquid or
solid hydrocarbon fuels are injected downstream of the
main combustion zone with NO and produce HCN3,8,9.
Advanced reburning is an even more recent technology
whereby ammonia, urea or similar substances are
injected aft of hydrocarbon injection for reburning to
further reduce NO x species10. Reburning and advanced
reburning technologies are commercially available

retrofit technologies, capable of providing 50%85%


NO x reduction beyond other combustion modifications.
This paper provides a review of reburning work,
including work at ACERC in the area of reburning.
No published review of reburning was located.
However, a review chapter on reburning is reported to
be forthcoming in a new encyclopedia11. To meet the
requirements of the Clean Air Act Amendments of 1990
(CAA) within budget limitations and scheduling constraints, electric utilities have considered a variety of
novel approaches to reduce NO x emissions from powerplants. Categories of NO x control options include burner
replacement, combustion modifications, fuel staging,
fuel reburning, steam or water injection, selective
catalytic reduction (SCR), and selective non-catalytic
reduction (SNCR)12. The efficiencies and limitations of
various methods have been summarized by Boardman
and Smoot3. Reburning technology is not fully understood theoretically, but it is a successful and commercially available retrofit technology, capable of providing
50%70% NO x reduction beyond other combustion
modifications13. Advanced reburning technologies are
reportedly capable of greater reductions.
The reduction of NO by hydrocarbons was
observed nearly half a century ago by Patry and
Engel14, and subsequently by Drummond15. This concept, and the term reburning, were first proposed by
Wendt et al.16 who noticed that, with the injection of
CH 4 just downstream of the primary flame zone, up
to 50% of NO was reduced, as shown in Fig. 1.
Myerson17 proposed that the overall reactions for NO
reduction by methane (and other hydrocarbons) could be
written as:
2NO 2CH4 2HCN 2H2 O H2 88 kcal

(1)

6NO 2CH4 2CO 4H2 O 3N2 428 kcal

(2)

Since this earlier work, much recent work has been


completed in several countries. This work has included
commercial demonstration, detailed laboratory measurement, and demonstration of improved and advanced
reburning technologies. Also, work on understanding
the kinetics of reburning and on predicting reburning
effectiveness during turbulent fuel combustion has
been studied. What follows is a brief review of this
recent work, including work in process.

2. TECHNOLOGIES

Reburning was apparently first demonstrated as a


practical NO x reduction method in Japan where the
concept of reburning was first applied to a full-scale
boiler by Mitsubishi in the early 1980s, more than 50%
reduction of NO being achieved18. BabcockHitachi
K.K. has also applied the technology successfully to
numerous wall-fired utility boilers in Japan19. Because of
these successful examples and the high efficiency of the
reburning technology on the reduction of NO x emissions,

NO x control through reburning

387

Fig. 1. Effect of CH 4 flowrate on NO x concentration with primary flame at 10% excess air, for a small laboratory-scale burner
without tertiary air addition16.

many investigators have conducted bench-scale and


pilot-scale reburning tests.
In a fossil fuel boiler configured for reburning, natural
gas, oil or coal (1030% of total fuel)20 is injected into
the upper furnace region. The overall process occurs
within three zones of a boilerprimary zone, reburning
zone, and burnout zone, as shown in Fig. 213. NO x forms
in the primary combustion zone with excess air, and then
reacts with hydrocarbon in the reburning zone to form
HCN and then N 2. The unreacted fuel completes
combustion in the burnout zone, where additional air is
added.
Laboratory experiments by Folsom et al.21 showed
that the NO x and SO 2 emissions may be reduced up to
60% and 20%, respectively, by reburning. Furthermore,
EER22 reported that in their gas reburning tests on a wallfired boiler, approximately 68% reduction of CO 2 was
achieved from the use of natural gas in place of coal for
part of the fuel. Several power plants in the U.S. have
employed this technique, and some experimental data

are being collected23. According to GRI24 and Folsom


et al.9, three full-scale, long-firing time evaluations of
gas reburning in coal-fired power plants had goals of
60% NO x reduction on a 172 MWe wall-fired unit, 65%
NO x reduction on a cyclone boiler, and 55% NO x
reduction on a 71 MWe tangentially fired unit, respectively (see Fig. 3). A planned gas reburning test was to be

Fig. 2. Schematic of the reburning process (adapted from


Pratapas and Bluestein13).

388

L. D. Smoot et al.

Fig. 3. Results of NO x reduction of three GRI-sponsored, fullscale, long-term projects for evaluation of gas reburning in coalfired power plants24.

carried out during 1996 on a 330 MWe coal-fired


cyclone boiler25. It has also recently been reported26
that Portugal has initiated a project to demonstrate coal
as a reburning fuel in the VadoLiguire power plant, with
a goal of less than 200 mg Nm 3 NO x emissions. EER27
has designed low NO x burners, to be used jointly with the
reburning process, and is demonstrating the performance
of a second-generation gas reburning system for NO x
control, which follows the successful demonstration of a
gas reburning system for over 14 months.
Four demonstration projects in the Department of
Energys Clean Coal Technology Program28 include
reburning, as listed in Table 1. One of these uses coal as
the reburning fuel while the others use natural gas. NO x
reductions varied from 30% to over 70%, depending on
furnace type and scale, reburning fuel and simultaneous
use of low-NO x burners. In the U.S. its principal
application has been to coal-fired boilers, while in
Europe and Japan reburning applications include coal,
oil and gas-fired utility boilers as well as municipal solid
waste (MSW) incinerators.

On the basis of an assumed 10% market penetration,


many countries, including Britain, France, Germany,
Russia and Japan, have established an international
research partnership to address the question of whether
gas reburning has "scale-up" limitations for large furnace
applications23. For example, a demonstration of gas
reburning on a 600 MWe coal-fired boiler at Longannet,
U.K., is scheduled to be completed during 1997. It is
supported by companies from Europe and the U.S., and
will be the largest scale application of gas reburning to
date29. Table 2 summarizes the commercial-scale
reburning demonstrations around the world. Several
variables are associated with the reburning system and
influence its effectiveness30, including:
reburn zone temperature;
residence time;
boiler load;
reburn fuel percentage of total boiler heat input;
reburn fuel composition;
reburn zone stoichiometry;
reburn burner stoichiometry (when coal is used as the
secondary fuel);
reburn burner pulverized coal fineness (when coal is
used as the secondary fuel);
flue gas recirculation rates to the reburn system;
reburn system spin vane and impeller (characteristics
and location);
overfire air port spin vane (characteristics and
location);
economiser outlet O 2%.
Another promising concept for NO x reduction
combines aspects of reburning with concepts from
selective non-catalytic reduction (SNCR), through the
joint use of a hydrocarbon (e.g., CH 4) and ammonia or
urea. In this scheme, urea or ammonia is injected into
the flue gas stream downstream of the aft-hydrocarbon
fuel injection point to further reduce NO x. This approach
has been labeled "advanced reburning"10,13. This
technique is quite attractive since it has been reported
that up to 85% NO x reduction can be achieved and, at
the same time, the problems of carbon loss, slagging
and tube wastage may be avoided10. However, residual
ammonia (i.e., ammonia breakthrough or slip) is a
challenge. The first observation of advanced
reburning data was reported by Chen et al.31, as shown
in Fig. 4.
The use of ammonia or urea in advanced reburning

Table 1. CCT reburning demonstrationscoal-fired boilers28


Co.

Boiler

Reburn fuel

% NO reduction

Comments

B&W

60110 MWe cyclone, coal-fired


Wisconsin

Coal
2030%

3652%, w/bitum
5362%, w/sub-bitum

Complete

EER

4075 MWe, tangential


Illinois
2033 MWe, cyclone
Lakeside, Illinois
172 MWe, wall-fired
Colorado

Natural gas
1525%
Natural gas
2233%
Natural gas
520%

NO x 67%
SO 2 5280%
NO x 6066%
SO 2 3263%
NO x . 30% LNB
6073% LNB/RB

Reburn with Ca-sorbent,


complete
Reburn with Ca-sorbent,
complete
Low NO x burners and
reburning in process

EER
EER

NO x control through reburning

389

In Sweden, the combination of gas reburning with


SNCR has been used in an MSW incinerator34. In shortterm tests, 70% reduction of NO x was achieved. In the
U.S., demonstrations of advanced reburning at a ten
million Btu hr 1 boiler have recently been completed10.
Currently, the second-generation advanced reburning
technology is being developed, in which a specific watersoluble, inorganic promoter salt (e.g., a sodium salt) is
added to ammonia and injected into the reburning zone.
Initial tests showed that 95% reduction of NO could be
achieved. Like CO, the presence of Na provides more
OH radicals to initiate the NH 3 oxidation sequence
through the reaction32

Table 2. Reburning commercial-scale demonstration


United States
67 coal-fired boilers
1 oil-fired boiler
1 with coal as reburning fuel
23 likely in current use
4 in CCT program
Europe
7 coal-fired boilers
2 municipal solid waste incinerators
Several glass furnaces
Japan
Several installations
Currently inactive
Demanding NO x regulation

[(Na)HOH] M (Na) OH H M

(8)

Without an N-agent, Na has no effect on NO reduction


efficiency. De Angelo and Sjoberg35 report new projects
to investigate the advanced reburning technique and the
micronized coal reburning process at two NYSEG utility
boilers.

3. MEASUREMENTS

Fig. 4. Influence of agent injection temperature into the rich


zone on percent NO removed, from the first advanced reburning
data31.

differs from its use in SNCR. In the advanced reburning


scheme shown in Fig. 5, NH 3 injection follows
hydrocarbon fuel injection to create a slightly fuel-rich
zone, combined with or followed by more air. The main
reactions of the advanced reburning are32:
NH3 OH, O, H NH2

(3)

NH2 NO N2 H2 O

(4)

With increased CO present, a larger temperature window


and reduction of NH 3 slip can be achieved. Chen et al.33
noted that, with reburning fuel injection, CO increases,
leading to:
CO OH CO2 H

(5)

H O2 OH O

(6)

O H2 O OH OH

(7)

This chain branching sequence provides additional OH


radicals to initiate the NH 3 oxidation sequence. In the
SNCR scheme alone, NH 3 is injected only into the postcombustion zone, at lower temperatures within the fuellean region.

Some experimental results with gas-based reburning


in the U.S. are discussed below. Marion, et al.36 have
shown that greater than 50% NO x reduction was
achieved with gas reburning in ABBCEs Boiler
Simulation Facility, which allows for carefully controlled experiments at a sufficient scale to replicate the
reburning process in full-scale units. Wendt and Mereb37
have completed a number of laboratory-scale, gas-based
reburning tests which consist of numerous premixedflame cases and a few diffusion-flame cases. They also
reported the effects of various parameters on the
efficiency of reburning, including the length of the
reburning zone and stoichiometric ratio. These data
could be used to evaluate predictions of reburning,
though no information on particle size distribution or
wall temperature profile was reported. Large-scale gas
reburning tests were conducted at the Illinois Power
Companys Hennepin Station, which was the first gas
reburning project ever demonstrated commercially for a
tangentially coal-fired boiler (80 MWe) in the U.S.23.
Recently, ABBCE has been involved in the application
of reburning technologies to two coal-fired utility boilers
and one oil-fired tangential boiler in the U.S. and Italy38.
Bilbao et al.39 in Spain made some exploratory study of
the effect of the operating parameters on reburning
efficiency. Their lab-scale data can be used for
comparisons of the modeling predictions.
In addition, several investigations with coal or other
hydrocarbons as the reburning fuel have been conducted
in the past few years, due to lower fuel costs. B&W19
demonstrated coal reburning for NO x control in a
100 MWe cyclone boiler. By using Southern Indiana
bituminous coal, greater than 50% removal of NO x was
obtained. The research group at Tennessee River Valley
Authority40 has planned a project called the Micronized

390

L. D. Smoot et al.

Fig. 5. Schematic of advanced reburning process (from Folsom et al.10).

Table 3. Wisconsin Power & Light sponsored project to demonstrate that pulverized coal is an effective reburning fuel for cyclone
boiler (from PETC23)
Load (MWe)

Baseline NO x
(ppm @ 3% O 2)

Reburn NO x
(ppm @ 3% O 2)

Percent reduction

Bituminous coal
110
82
60
3738

609
531
506
600

290
265
325
400

52.4
50.1
35.8
33.3

Sub-bituminous coal
110
82
60

560
480
464

250
230
220

55.4
52.1
52.6

Coal Reburning Demonstration (MCRD) on a 175 MWe


wall-fired unit. 5060% removal of NO x was projected
through use of micronized (80% less than 325 mesh)
eastern Kentucky and West Virginia coals. Syverud
et al.41 reported that, with automobile tires as a reburning
fuel, reduction of up to 40% of NO x emissions, 50% of
CO emissions, and 25% of SO 2 emissions can be
achieved. According to PETC Review23, a project
conducted by Wisconsin Power and Light (WP&L)
showed that pulverized coal is an effective reburning
fuel for reducing NO x emissions from cyclone boilers.
As shown in Table 3, using bituminous coal, NO x
reduction was as high as 58% at full load. With subbituminous coal, the full load NO x reduction was up to
62%.
Natural gas as the reburning fuel has some advantage
over other fuel types since it contains no fuel-bound
nitrogen, sulfur, or particulate matter, and it reacts faster
than coal or oil. Wendt42 suggested that natural gas
produces more hydrocarbon species in the NO reduction
zone. However, with coal reburning, HCN was destroyed
more rapidly, which is a key intermediate to reduce NO
to N 2 in fuel-rich regions. He indicated that bound
nitrogen in the reburning fuel is not an issue, due to only

minor differences in the measured NO profiles, regardless of the nitrogen content in different reburning fuels.
On the basis of investigations into the reduction
efficiency of different reburning fuels, Kicherer et al.43
concluded that a high NO x reduction level can be
achieved under the following conditions: high volatile
matter of reburning fuel, long residence time, optimized
mixing conditions, and very fine grinding if solid
reburning fuels are used. He indicated that the homogeneous reduction mechanisms are more effective than
the heterogeneous processes. However, Chen and Ma44
observed, as shown in Fig. 6, that lignite as a reburning
fuel has a higher efficiency for NO x reduction than CH 4
or other coals in their small-scale laboratory reactor.
They noted that heterogeneous reactions contributed to
higher levels of NO x reduction than homogeneous
reactions under certain conditions. Moyeda et al.45 also
confirmed experimentally that the low-rank coals generally performed in an equivalent manner to natural gas,
while high-rank coals were generally less effective as
reburning fuels. They found that increasing the reburning
zone residence time improved the NO x reductions
achieved with all of the coals, but not for natural gas.
Payne et al.46 believe that fuel volatility influences the

NO x control through reburning

391

Fig. 6. Comparison of coal, char and methane as reburning fuels simulated for a lab-scale reactor44.

Fig. 7. Schematic diagram of a novel reburning burner47.

availability of fuel in the reburn zone and the evolution


of radical species; therefore, fuels with higher volatile
content would be expected to achieve higher NO x
reduction.
Another interesting study was conducted in a semiindustrial scale (2.5 MWt) furnace with a swirlstabilized internal fuel-staged burner, shown in Fig. 7,
with different reburning fuels including two highvolatile bituminous coals, heavy fuel oil, natural gas
and coke oven gas47. This work showed that the degree
of NO x reduction is independent of reburning fuel types,
and that up to 89% NO x reduction can be achieved. Chen
et al.48 also showed that the reburning efficiencies with
ethylene, isobutane and methane were similar. Benchscale experimental results by Spliethoff et al.49 also
showed that, with concentrations of various hydrocarbons

between 20 and 100% in reburning fuels, no effect of fuel


type on NO emission could be observed. Although the
reburning efficiencies are somewhat dependent on the
concentrations and types of hydrocarbons (i.e., fuel
types), the experiments of Greul et al.50 show that
pyrolysis gas may be one of the best reburning fuels,
because N-containing tars from pyrolysis gas can further
selectively reduce nitrogen oxides in the reduction zone.
In 1992, a joint work effort was initiated as a part of
the Phase 1 Combustion 2000 (HiPPS) program to
evaluate low-NO x burner concepts. The overall concept
was to develop a new, combined-cycle power generation
system using state-of-the-art gas turbines in conjunction
with innovative, low-NO x designs for coal combustion.
In this project, a unique ceramic heat exchanger is
integrated with the combustor to produce hot air which is

392

L. D. Smoot et al.

used to drive the turbines, and use is made of state-ofthe-art ash management technologies. Bench-scale
experimentation has been conducted to determine the
role of reburning to lower NO x levels in the coal-fired
laboratory flame, in conjunction with other low-NO x
technologies51.
A new 29 kW, refractory-lined, pulverized coal
combustion research facility was fabricated with a
16 cm diameter and a 7.3 m length, with composite
refractory walls to minimize heat losses. The burner was
mounted on the top left side of the U-shaped furnace.
The long, vertical path below the burner allowed long,
axial flames. This facility has been used extensively to
evaluate technologies to achieve NO x emissions below
0.06 lb/MM Btu using several alternative combinations
of optimized SNCR with stage, gas-stabilized axial coal
flames and reburning. Results show an almost 90% NO x
reduction from the current standards51. This work has
also explored the important parameters in the nearburner region, using a natural gas-stabilized burner. The
final step in this study was to determine the lowest levels
of NO that could be obtained from the optimized use of a
gas-stabilized, pulverized coal flame in conjunction with
air staging, natural gas reburning and SNCR. Figure 8
summarizes the results of a wide spectrum of tests
conducted to evaluate various combinations of these
schemes in an integrated mode. These data indicate that
the desired range of emissions can be achieved,
depending on which of the various combinations of
techniques are used51.
A hardwood and softwood have also been evaluated as
reburning fuels in the U-shaped furnace52. Results
showed that a reduction of 5060% NO was obtained
with approximately 10% wood heat input. There is a net
reduction in CO 2 and SO 2 since wood is a regenerable
biofuel and contains no sulfur. These results suggest that
wood is as effective as natural gas or coal as a reburning
fuel.
Measurements of the reburning process are important
to gain a fundamental understanding of this process, and
for validation of predictive techniques for NO reactions
like that described previously. Tree53 and colleagues are
working to provide detailed in situ measurements for a
laboratory-scale, coal-fired, control-profile reactor
(CPR) utilizing reburning and advanced reburning

technologies. The objective of this project is to develop


a better understanding of reburning and advanced
reburning by comparing detailed combustion measurements of temperature, gaseous species and velocity with
predicted values, focusing primarily on improving the
models used to describe the reburning and advanced
reburning reactions. PCGC-3 model predictions of
temperature, velocity and species entering the reburning
and advanced reburning zones will be refined to match
upstream measurements; then comparison of the results
of NO, NH 3 and HCN measurements within the
reburning zone with predicted values will be made.
Mapping will include measurements at a matrix of radial
and axial positions in one of the four quadrants of the
CPR. Species to be mapped include CO, CO 2, O 2, N 2,
NO x, HCN and NH 3. Conventional instrumentation
includes suction pyrometer for temperature, Pitot tube
for velocity and various gas analyzers for chemical
analysis of the gaseous species. Advanced optical
instrumentation will include laser doppler anemometry
(LDA) for velocity and coherent Raman spectroscopy
(CARS) for temperature. Reburning technologies to be
tested will include downstream natural gas injection and
joint use of natural gas and ammonia (advanced
reburning). Initial reburning results in a pulverized coal
flame show NO reductions of up to 65% with natural gas
injection closer to the flame zone54. Natural gas injection
downstream led to a lower NO reduction of 3540%.

4. MECHANISMS AND RATES

Wendt42 has recently published a review of mechanisms governing the formation and destruction of NO x.
NO reduction through the reburning-NO mechanism
usually includes the interactions of HCN and NO
species, as described in the elementary reaction steps
of reburning noted by Wendt et al.16. Under fuel-rich
conditions, the formation of HCN relies strongly on the
concentration of hydrocarbon species,
CHi NO HCN

(9)

which then decays through NCO NH N, as shown


in Eqs. (1012), and ultimately reaches N 2 via the

Fig. 8. Strategies for NO x reduction based on measured data in 29 kW U-shaped furnace51.

NO x control through reburning

reverse Zeldovich reaction, Eq. (13):


HCN O NCO H

(10)

NCO H NH CO

(11)

NH H N H2

(12)

N NO N2 O

(13)

The above mechanisms illustrate why pure hydrogen is


less effective for reburning. However, under fuel-lean
conditions, hydrocarbons react with oxygen and/or
hydroxyl radicals via Eq. (14) to form CO:
CHi O CO H

(14)

Therefore, reactants in Eqs (9), () and (14) consume CH i


competitively. A goal in reburning optimization is to
maximize the exposure of NO to CH i and minimize
CH i interaction with oxygen48.
Reliable predictions of NO x reduction with reburning
depend partly on an understanding of the controlling
sequence of elementary reactions. The structure of
existing comprehensive elementary reaction schemes
for hydrocarbon and other oxidation systems has been
painstakingly developed over many years by a number of
different groups (e.g., Miller and Bowman4), and the
mechanism sets are becoming larger and more reliable55.
However, a reliable kinetic scheme is not easy to
assemble since the collective performance of reactions in
a kinetic scheme on the basis of their independently
determined Arrhenius parameters does not necessarily
yield the closest match to a range of experimental
results56. Commonly used schemes of kinetic models
include the MB model4, Bians model57, and Warnatzs
model58. The new version 2.11 of GRIMech59 includes
nitrogen chemistry relevant to reburning. The accuracies
of these models vary from low temperature to high
temperature, and from low pressure to high pressure, for
different flames.
Many researchers studying reburning mechanisms
through elementary chemical reactions, such as Glarborg
and Hadvig60, Baulch et al.61, Thorne et al.62, Li et al.63,
Hura and Breent64, and Burch et al.65, found for various
reburning fuel types that two major kinetic pathways
control the efficiency of reburning, i.e.:
C, CH, CH2 NO HCN

(15)

HCN O, OH N2

(16)

20

Chagger et al. found experimentally that NO x is


slightly reduced by the presence of SO 2. On the basis
of a sensitivity analysis for a 163-step elementary
mechanism and experiments, Thorne et al.62 pointed
out that the largest sensitivity coefficients are for the
two reactions:
H O2 OH O

(17)

C, CH, CH2 NO HCN

(18)

393

He believed that the CH 2 is less important, but Chen and


Ma44 suggested that CH 2 is as important as CH. Wendt42
included the following reaction as an important path for
NO x destruction:
NO NHi N2 products

(19)

Comparing the predicted results from Baulchs


elementary chemistry mechanisms with the experimental
data, Etzkorn et al.66 indicated that the current reburn
mechanisms cannot describe the practical reburn process
and should be improved. Stapf and Leuckel67 concluded
that current comprehensive mechanisms, such as that of
Miller and Bowman4 or Glarborg and Hadvig60, show
strong dependence on the initial oxygen content and
predict shorter global reaction time scales than are
found in the flow reactor experiments. They suggested
that further investigations are necessary to clarify
whether the NO CH i pathways are sufficiently understood. Burch et al.68 also showed that their experimental
data did not support the mechanism of Eqs. (9)(14).
Kristensen et al.69 investigated nitrogen chemistry in
the burnout zone of reburning, and found that the minimum in NO is related to the amount of CO, the oxidation
of NH 3 or HCN, and the reduction of NO by NH 3 or
HCN.
While detailed chemical kinetics of NO x have been
studied extensively and tabulated, it is currently
impractical, because of unacceptable computer requirements, to incorporate large elementary chemical kinetic
schemes into comprehensive combustion computer
models which include turbulent fluid dynamics, heat
transfer and chemical reactions. In addition, the formation and destruction rates of NO x are of the same
magnitude as the turbulent mixing rates, and consequently the equilibrium assumption is not adequate for
calculation of NO x concentrations70,71.
Work has been reported recently to deduce a global
reaction rate for the reburning reaction C iH j NO
HCN ...72. A method for obtaining a global chemical
reaction rate expression and its associated rate constant
was documented73, and the global reaction rate was
subsequently reported74. A reaction scheme with 254
gaseous elementary reactions was used4,75 in simulations
of NO formation in premixed, laminar hydrocarboncontaining flames, together with the CHEMKIN code76.
De Soete7779 had conducted extensive experiments on
the formation of nitrogen-containing pollutant species in
premixed flames. From correlation of these experimental
results, a set of global reactions and corresponding rate
constants was determined. Methane (CH 4) and ethylene
(C 2H 4) were used as fuels, with NO, C 2N 2 or ammonia
(NH 3) being added as fuel-nitrogen sources. Though De
Soetes experimental work was performed over two
decades ago, global rates of NO formation from fuelnitrogen measured in those studies are still widely used in
predictions of fuel-NO x in hydrocarbonair combustion
systems.
Rate constants for the following global reactions were

394

L. D. Smoot et al.

obtained by De Soete79 by correlating these laboratory


data:
(20)
HCN O2 NO
HCN NO N2

(21)

73

The method used by Chen et al. for determining the


global reburning rate was identical to that used by De
Soete7779 for global fuel-NO reactions, with one exception. De Soete determined mole fractions and their
derivatives in the flow direction from test data, whereas
Chen et al.73 used calculated mole fraction profiles for
one-dimensional, premixed flames. Otherwise, De
Soetes method was applied to deduction of the reaction
rate for the NOC iH j reburning reaction 74:
X
(22)
Ci Hj NO HCN
The global fuel-NO formation pathway illustrated by
Fig. 9 includes the reburning reaction. Sensitivity
analyses72 indicated that reburning-NO steps were
important and sometimes rate determining in the formation and destruction of NO in hydrocarbon-rich regions
of flames. Apparently, without the reburning-NO steps,
simulations of NO formation will result in inaccurate
profiles of NO, HCN and NH 3 in these fuel-rich
hydrocarbon flames.
oC iH j was taken to include all hydrocarbon species
which occur locally in the flame zone. In this work, these
species included C 2H 4, C 2H 3, C 2H 2, C 2H, CH 4, CH 3,
CH 2 and CH. However, throughout the simulations used
to determine a global reburning kinetic rate72, only CH
and CH 2 were predicted in significant concentrations,
which was also observed experimentally by Wendt and
Mereb37.
A method based on the simulations discussed
previously was used to quantitatively evaluate the
global reburning-NO pathway. From reaction (22) and
NO
, in the absence
Fig. 9, the global reburning-NO rate, r22
of NH 3 reactions, thermal-NO and prompt-NO, was
expressed by material balance as:
dX
NO
r20 NO r21
(23)
r22
dt
An expression for k 22 is:
dX
k20 XHCN XOb 2 NO k21 XHCN XNO
r22
dt

k22
XHC XNO
XHC XNO
(24)

Fig. 9. Global NO x formationdepletion mechanism74.

where b is the function of the oxygen concentration


given by De Soete79, and X HC is the sum of all hydrocarbon radical and hydrocarbon species concentrations.
In Eq. (24), k 22 is a function of k 20, and hence comparison
of k 22 with different k 20 expressions by De Soete79 was
also reported. Three correlations of the global reburningNO rate expression were obtained for CH 4 and C 2H 4
flames from this method74. Two k 22 values were
those obtained from values of k 20 derived from simulations73 with different b values, and the third k 22 value
was obtained using De Soetes experimental work79 for
k 22.
The correlated rate constants from b 1.5 and
b f (XO2 ) showed only a small difference, which means
that the global reburning rate constants were not very
dependent on the oxygen concentration in these
flames. The k 22 rate based on the simulated k 20 was
higher than k 22 based on the experimental k 20 (from De
Soete79).
Three more k 22 correlations by Eq. (24) with NOseeded flames were also obtained. When NO seeding was
used, the reburning-NO mechanism was dominant, and
the formation of NO through the fuel-N pathway was
thought to be less important. The global activation
energies from these expressions were within 6 6% of
the average value, and the pre-exponential factors were
within a factor of two and a half of the average value.
The global reburning rate with NO seeding exhibited
lower values than those without NO seeding.
The k 22 rate expression recommended by Chen et al.73
for a global NO x kinetic model which is a composite of
the several expressions is:


18800
k22 2:7 3 106 exp
(25)
RT
The quantitative evaluation of this new global rate
expression and its parameters by comparison with the
data from turbulent, particle-laden diffusion flames
with reburning is discussed subsequently.

5. PREDICTIVE METHODS

While several predictive models for formation of NO


through modeling of thermal and fuel processes have
been developed (e.g., Boardman and Smoot3), until
recently there has been little published work on
destruction of NO through reburning processes.
Several approaches to predict reburning-NO are
being considered and models are now becoming
available.
Mereb and Wendt8 simplified the detailed kinetic
mechanism of Glarborg et al.80 for natural gas reburning,
and then developed an engineering model with the
assumption of partial equilibrium of some intermediate
species:
d[NO]
[NO][NH3 ] f1 [NO][CH4 ] f2
dt

(26)

NO x control through reburning

395

Fig. 10. Comparison between measured and predicted nitrogenous species for three laboratory-scale gas reburning experiments,
using the model developed by Mereb and Wendt8.

d[HCN]
[HCN]( f3 f4 ) [CH4 ] [NO] f2
dt

(27)

[N2 ] f5 [NH3 ] f6
d[N2 ]
[NO][NH3 ] f7 [CH4 ][N2 ] f5
dt

(28)

d[NH3 ]
d[NO] d[HCN]
d[N2 ]

2
dt
dt
dt
dt

(29)

Glarborg et al. 81 developed a reduced mechanism for


nitrogen chemistry in methane combustion, based on
Miller and Bowmans4 full mechanism, by using
Peters82 systematic reduction strategy. This contained
a four-step mechanism for methane oxidation plus two
rate-of-production terms for HCN and NO. The model
generally provides a good description of formation and
destruction of nitrogen oxides compared with the full
mechanism, as shown in Fig. 11.

where f i function ([OH], [H 2O], temperature), and the


CH 4 concentration is calculated from partial equilibria:
CHi OH CHi 1 H2 O

(30)

If initial estimates for OH and H 2O are available for a


given temperature and initial measured values of CH 4,
NO, HCN, NH 3, H 2, H 2O and N 2 for this 1-D approach,
the profiles of NO, HCN, N 2 and NH 3 can be computed.
As shown in Fig. 10, this model provides reasonable
predictions for HCN but high values for NO and low
values for NH 3 in the reburning zone over a wide
range, provided CH 4 is the reburning fuel. Recently,
Wendt42 extended the model to predict N 2O concentrations. However, this simple kinetic model cannot be integrated into a comprehensive computer code since no
turbulence effects are included, and some measured
values are required.

Fig. 11. Comparison of results for methane combustion with NO


added in a stirred reactor as a function of fuel/air equivalence
ratio (circle: experimental NO data; square: experimental HCN
data; solid line: full mechanism; dashed line: skeletal
mechanism; dotted line: reduced mechanism)81.

396

L. D. Smoot et al.

Starting with full mechanisms of a hydrocarbon flame,


several other researchers have calculated the NO x
emissions during reburning. Li et al.63 used a PC-based
computer model, including 2-D furnace combustion and
heat transfer, and 1-D chemical kinetics with detailed
gaseous hydrocarbon and nitrogen reaction mechanisms
(i.e., 43-species, 201-step NO x mechanisms) to evaluate
reburning/cofiring performance without particles or
turbulence interactions. After simulating four fullscale, coal-fired gas-reburning boilers, Payne and
Moyeda83 indicated that the coupling of a detailed
chemistry model with a simplified mixing model, based
upon the results of thermal and physical models of the
full-scale system, can reasonably predict the general
performance trends observed in the full-scale system, but
that more accurate models that couple the complex
interactions between turbulence and chemistry which
occur in the turbulent mixing process are needed. Hura
and Breent64 investigated the effects of variation in
temperature, reburn zone stoichiometry, initial NO
levels, flue gas/natural gas mixing rate, and heat loss
on reburning effectiveness by using Boyles 193-step
kinetic model and a modified CHEMKIN code with the
equilibrium assumption for major species including N 2,
CO 2, H 2O, O 2, NO, CO and OH. Ballester et al.84
utilized 0.5 MWt experimental furnace measurements
and then extrapolated to a 220 MWe tangentially coalfired boiler and a 350 MWe wall-fired boiler, using CFD
modeling with detailed NO x chemistry. The modeling
results were in reasonable agreement with reburning
field experience. A CFD model combined with seven
chemical species and four reactions in a reduced NO
model for methane combustion has been used for an
MSW incinerator by Rasmussen85. The isothermal
modeling reportedly provided a reasonable prediction
of observed results. Tyson86 suggested that these
calculational processes, assuming finite rate chemistry
with instantaneous mixing, predict a limit of 8090%
NO x reduction (with reburning), but such high NO
reduction values cannot be achieved due to mixing
limitations.
None of these computational processes considered the
turbulencechemistry interactions or multidimensional
effects on the reburning process. Chen72 also suggested
that the reduced models derived with these various
techniques are still too complicated to predict combustion systems if the models are to account for turbulence
and multidimensional effects. One solution of the
problem of incorporating effects of turbulence is to
find a reduced mechanism by using various approaches
including global modeling, detailed mechanism reduction, response modeling, chemical lumping, or statistical
lumping87.
Babcock and Wilcox88 have recently incorporated the
global reburning reaction of Chen et al.74 into their NO x
formation submodel, which itself is a component of a
more general comprehensive combustion model. They
have recently applied this predictive method to a largescale utility boiler fitted with a cyclone combustor. They
first adjusted parameters in the NO x model to match

model predictions with observed results for a full-load


furnace without reburning. Then, predicted NO x concentrations with reburning were reported to be reliable at
full load. However, predictions at lower load (e.g., 75%)
with reburning under-predicted the observed reduction in
NO x concentration. They reported this same experience
in a few other cases with and without reburning.
Fiveland88 emphasized their need to adjust the predictions for full load without NO x control. A commercial
software code, SOAPP, developed by Sargen & Lundy
and EPRI, is available for the analysis of natural gas
reburning in utility boilers fired by coal or oil89.
Brouwer et al.90 developed a reduced mechanism (6
species, 7 reactions) for the SNCR process in practical
systems, with the influence of CO. The reagents included
ammonia, cyanuric acid and urea. This simplified SNCR
mechanism was derived from Miller and Bowmans
mechanism4 through sensitivity analysis and curve fitting.
For example, using NH 3 as the reagent, the global model
was expressed as follows, where for the reaction
NH3 NO N2 H2 O H

(31)

A 4.24 3 10 8 cm mol s K, b 5.30, and E


83 600 K, and for the reaction
NH3 O2 NO H2 O H

(32)

A 3.50 3 10 5 cm mol s K, b 7.65, and E


125 300 K. Hence, the rate constants (k AT be E/T)
are defined. This reduced mechanism was integrated
into their comprehensive CFD code, JASPER. Figure
12 compares predictions from JASPER with the reduced
model to independent, premixed ammonia data of
Lyon91 from a turbulent flow reactor under isothermal
conditions, and to calculations from CHEMKIN76 with
the full mechanism.

Fig. 12. Comparison of the reduced SNCR chemistry in


JASPER with the full chemistry set and the data of Lyon for
homogeneous conditions (dotted line: full chemistry; solid line:
JASPER calculations with reduced chemistry; triangle: data of
Lyon91)90.

NO x control through reburning

In order to account for the radical chemistry effects of


CO, these authors used an empirical effective temperature rather than the predicted temperature in the rate
expression, i.e.,
Teff T S(CO)

(33)

where S(CO) 17.5 3 ln ([CO]) 68.0 has been


deduced from curve fitting to match CHEMKIN
predictions with the full mechanism.
Brouwer et al.90 also made experimental SNCR
measurements in a 29 kW, refractory-lined furnace, in
which a typical product stream flue gas was flowing with
500 ppm NO, 750 ppm NH 3; various levels of CO and
N 2 were then injected with the NH 3 reagent downstream.
In this work, three important parameters on SNCR
effectiveness were investigated: local CO concentration,
temperature quench rate, and ammonia injection
momentum. As shown in Fig. 13, JASPER with the
reduced SNCR submodel with the turbulence gave good
predictions for this turbulent case.
This work may be the first global model for the SNCR
process with the effect of CO. This approach is being
applied to advanced reburning technology by the authors
in ongoing work.
In order to predict the advanced reburning process,
Brouwer92 has recently extended the above work90 to
treat advanced reburning through use of the set of global
reactions. The predicted results showed agreement with
measurements, but this work has not been published.
At this center, development of an advanced reburning
model is also in process93. Using Peters82 method, a
7-step reduced mechanism has been derived from a
312-step, 50-species full mechanism, which correctly
predicts observed trends, including the effects of
temperatures, the ratio of NH 3 to NO, and concentrations
of CO, O 2 and H 2O on NO reduction94. Incorporation of
this 7-step mechanism into a comprehensive, threedimensional combustion code (PCGC-3) and its
evaluation is in process93.
All of the predictive approaches noted above are for
gas as a reburning fuel. So far, there is no model reported
to predict NO x emissions with coal or other solid
hydrocarbon compound as the reburning fuel where
heterogeneous kinetics have been reported. In fact, many
researchers have noticed that the heterogeneous reaction
of NO with char is important in the reduction of NO from
coal combustion processes9597. As Guo and Hecker98
summarized, the investigations of the reaction of NO
with char involve the kinetics and mechanism95,96, the
effects of char surface area99,100, the effects of char ash
and its composition, the catalytic effects of metals97,101,
and the effects of feed gas composition95,102. The
reaction of NO with char has generally been reported
to be first order with respect to NO partial pressure. Guo
and Heckers experiments showed an increase in the
apparent activation energy with increasing temperature,
and a decrease as the CaO content increases. A sharp
shift in the activation energy has been observed, as
shown in Fig. 14. This shift to higher activation energy

397

with increasing temperature is opposite to that expected


if a reaction is changing from chemical rate control to
mass transfer control, and suggests different mechanisms
or rate-determining steps at high and low temperatures.
Questions concerning N 2 formation, the surface complexes, the nature of active surface sites, and the effects
of active minerals in char are still not well understood.
Also, most studies of heterogeneous reactions have been
conducted at relatively low temperatures (,1000 K)
compared to those of reburning injection locations
(,1500 K). Only a few investigations have been done
at high temperatures. Teng et al.96 tabulated the

Fig. 13. Measurements and predictions of the effect of reagent


jet/product stream momentum ratio on SNCR with and without
CO at high and low quench rates. Symbols are data (square: low
quench with 0 ppm CO; circle: low quench with 750 ppm CO;
triangle: high quench with 0 ppm CO; diamond: high quench
with 750 ppm CO), and lines are results from mixing and
reduced chemistry model (solid line: low quench with 0 ppm
CO; dashed line: low quench with 750 ppm CO; solid line: high
quench with 0 ppm CO; dotted line: high quench with 750 ppm
CO)90.

Fig. 14. Arrhenius plots of the reaction constants for NO reduction for three char types (from Guo and Hecker98), where NDL
is North Dakota Beulah Zap lignite char, NDW is NDL washed
with HCl, and NCa is NDW reloaded with calcium oxide.

398

L. D. Smoot et al.

activation energies for various types of NOcarbon


reactions, but the kinetic data at both low and high
temperatures are not yet sufficiently accurate. Consequently, reliable heterogeneous models for the reaction
of NO with char or coal at high temperatures have not
been reported.

6. REBURNING MODEL APPLICATIONS

Chen72 used PCGC-2 to initially evaluate the NO x


submodel with the global reburning reaction. Two
laboratory-scale reburning cases from Mereb and
Wendt8 were simulated. The predicted NO profiles
were consistent with the measured results.
NO x predictions including the reburning reaction
(Eq. (22)) have also been made with a comprehensive
combustion model, PCGC-3, which incorporated the
NO x submodel of Fig. 993. PCGC-3 is a generalized,
comprehensive, three-dimensional, steady-state combustion flow code for modeling turbulent combustion
in practical systems103. It is the third generation of
PCGC-3 (pulverized coal gasification and combustion in
3 dimensions), and incorporates advanced technology for
describing NO x reburning, coal devolatilization, char
oxidation, oil-droplet combustion, and multiple coal offgas progress variables. It can be used to characterize nonreactive and reactive flows with entrained particles or
droplets, and is based on the general momentum, energy
and mass conservation equations for turbulent systems.
Details of this CFD/combustion predictive method are
documented elsewhere104,105.

The NO x pollutant submodel incorporated in PCGC-3


uses a post-processing approach to describe the formation and reduction of NO in the flow field. Reactions of
NO involve thermal, fuel and reburning processes,
according to Fig. 9. Formation of thermal NO is
determined from the extended Zeldovich reactions,
with two resulting rate expressions, one for fuel-rich
and one for fuel-lean conditions. Reactions of fuel
nitrogen involve two additional species, HCN and NH 3,
and rate constants based on the work of several
researchers can be selected for the fuel NO mechanism
reaction rate constants. This NO x pollutant submodel
includes reduction of NO by the reburning process
according to reaction (22).
Two other laboratory-scale, coal-fired diffusion-flame
cases with gas reburning from Mereb and Wendt8 have
subsequently been simulated with PCGC-3 to further
evaluate the reburning model93. The input conditions of
the two cases are listed in Table 4. The diameter and
length of the cylindrical drop-tube reactor were 0.15 m
and 2.12 m, respectively. The location of the reburning
injector is 0.53 m. Case 1 has a tertiary stream located at
0.99 m for the addition of final combustion air. There
was no tertiary air injection in Case 2. In Case 1 the
reburning stream was methane, and in Case 2 the
reburning stream consisted of methane and N 2. The ratio
of reburning fuel to total fuel was about 1113% for both
cases. Coal particle size distributions and wall temperature profiles were not reported for either case. However,
Wendt106 indicated that the wall temperatures were
thought to be within 50 K of the centerline gas
temperatures and that the particle size distribution was

Table 4. Input conditions for two cases with gas reburning laboratory furnace, coal-fired8
Variables

Case 1

Case 2

Primary zone stoichiometry


Reburning zone stoichiometry
Tertiary zone stoichiometry

1.1
0.9
1.06

1.1
0.88

Primary stream
Temperature (K)
Air flow rate (kg s 1)
Type of coal
Est. mass mean particle size (mm)
Mass coal/mass air

590
5.5 3 10 4
Utah bituminous
80
0.494

590
5.5 3 10 4
Utah bituminous
80
0.691

Secondary stream
Temperature (K)
Air flow rate (kg s 1)

590
2.079 3 10 3

590
3.2079 3 10 3

Reburning stream
Temperature (K)
CH 4 flow rate (kg s 1)
N 2 flow rate (kg s 1)
Mass CH 4 /(coal CH 4)
Injection location (m)

300
3.088 3 10 5
0.00
0.125
0.533

300
4.88 3 10 5
2.216 3 10 4
0.114
0.533

Tertiary stream
Temperature (K)
Air flow rate (kg s 1)
Injection location (m)

590
4.673 3 10 4
0.99

Est. wall temperature (K)

950 a

1400 a

No measured wall temperatures were available. Value adjusted to match temperature profile data.

NO x control through reburning

a standard unclassified grind of about 70% through 200


mesh. Both of these parameters are shown later to have a
significant impact on the NO profiles. The mass mean
particle diameter used herein was estimated to be about
80 mm for the baseline calculations. Effects of activation
energy (E), pre-exponential factor (A), particle diameter,
particle size distribution, reburning gas injection velocity
and reburning gas composition were considered93. Base
values of E 1.89 3 10 4 and A 2.72 3 10 6 were used
in these computations, essentially the same as those
recommended by Chen72 and Chen et al.73.
6.1. Reaction Parameters
Figure 15 shows the effects of the activation energy
(E) for the reburning reaction on the predicted NO
concentrations for Case 1. All other parameters remained
constant for the simulations. The average NO concentration is shown as a function of axial location in the
furnace for the cases with no reburning and three
different values of activation energy. The value of
activation energy, E 1.89 3 10 4, results in NO

399

concentrations similar to those achieved with no


reburning, in which methane was still injected but the
reburning reaction was inactive. Reducing the value of
the activation energy resulted in more reduction of NO
by the reburning reaction. The value of E 0.75 3 10 4
for the activation energy subsequently shows good
agreement of the measured and predicted values.
Figure 16 shows the effects of changes in the preexponential factor (A) for the reburning reaction on the
predicted NO concentrations for Case 1 with the value of
E 0.75 3 10 4. All other parameters remained constant
in the simulations. The average NO concentration is
shown as a function of axial location in the furnace for
cases with no reburning for three different values of preexponential factors. The predicted values are also
compared with the experimental values. Changes in the
value of the A had a significant effect on the predicted
NO concentrations, and the base value, A 2.72 3 10 6,
resulted in the best agreement between measured and
predicted values.
Figure 17 shows profiles of average NO concentration
for simulations with no reburning and with reburning for

Fig. 15. Effect of activation energy (E) on predicted NO concentrations for Case 1.

Fig. 16. Effect of pre-exponential factor (A) on predicted NO concentrations for Case 1.

400

L. D. Smoot et al.

Fig. 17. Effect of activation energy (E) on predicted NO concentrations for Case 2.
Table 5. Selected particle size distributions for model simulations
P1
P2
P3
P4
P5

Size (mm)
Mass fraction
Size (mm)
Mass fraction
Size (mm)
Mass fraction
Size (mm)
Mass fraction
Size (mm)
Mass fraction

1.7
0.0024
11.8
0.1
25.0
0.15
51.4
0.479
80.0
0.0024

6.5
0.0126
35.3
0.3
55.0
0.3
73.5
0.169
80.0
0.0126

two activation energies compared with the measured


values for Case 293. As before, the trial without
reburning was performed by still injecting the same
amount of methane but turning off the reburning
reactions. E 1.89 3 10 4 is the base value. E 0.75
3 10 4 is the value of the activation energy which showed
the best agreement with measured values of NO
concentration for Case 1. As before, the E base value
shows little difference with the trial without reburning,
and the predicted NO concentrations obtained with E
0.75 3 10 4 show a significantly better agreement with
the measured values. Consequently, the values of E
0.75 3 10 4 and A 2.72 3 10 6 were used in the balance
of the simulations.
6.2. Particle Size
Wall temperatures and particle size distribution were
not given by Mereb and Wendt8. The wall temperatures
were estimated from the experimental average gas
temperature, but the particle size and distribution were
not reported. Therefore, the influence of particle size and
distribution on the NO x emissions was investigated. A
specific particle size distribution was selected as baseline
trial, namely, P1 (see Table 5). The effects of mass mean
particle diameter from 64 to 96 mm for the same
distribution shape for Case 2 are shown in Fig. 18. Effects
on predicted NO x profiles were not very significant.
Computations were also made for five different

40.0
0.51
75.0
0.3
89.0
0.3
86.4
0.094
80.0
0.51

105.0
0.366
132.0
0.2
120.0
0.2
122.5
0.194
80.0
0.366

195.0
0.109
195.0
0.1
185.0
0.05
175.1
0.064
80.0
0.109

d mm 80.2
80.2
80.2
80.1
80.0

particle distributions at a constant mean particle


diameter of 80 mm, as shown in Table 5. From P1 to
P5, the particle distributions become narrower and
narrower. These five sets of particle distributions were
used to simulate Case 1, and results are shown in Fig. 19.
In Fig. 19(a), the location of ignition varied with
changing particle distribution. With the P1 distribution
containing a large fraction of small particles earlier
ignition was observed, whereas with the P5 distribution
later ignition was observed. From Fig. 19(b), using the
broad P1 distribution, the lowest predicted NO level was
obtained, whereas with the use of the P5 distribution the
highest NO value was predicted. Smaller particles cause
earlier ignition and much lower NO concentrations. The
same observations resulted for Case 2, as shown in Fig.
20. For the same mass mean particle diameter, the
highest NO was observed with the uniform particle size.
Predicted NO levels were far more sensitive to the
presence of small particles than to changes in mean
particle diameter, even though only very small mass
percentages of very small particles are present.
6.3. Reburn Gas Composition
Figure 21 shows the effects of changes in the mass
percentage of CH 4 in the reburning fuel on the predicted
NO concentrations. In the N 2 100% trial, no CH 4 is
injected, and the reduction in the NO concentration
results from dilution and reduction in the gas

NO x control through reburning

401

Fig. 18. Effect of mean particle size on NO x profiles for Case 2.

Fig. 19. (a) Effect of particle distributions on gas temperature for Case 1. (b) Effect of particle distributions on NO profiles for Case 1.

temperature. Increasing the mass of CH 4 to 18% results


in significant NO reduction compared to the case without
CH 4. A further increase of the mass of CH 4 to 50% does
not show significantly lower NO concentrations,

especially at the exit of the reactor. This is consistent


with experimental observations by Chagger et al.20, who
reported an upper limit in hydrocarbon gas percentage
for the reburning reaction.

402

L. D. Smoot et al.

Fig. 20. Effect of particle distributions on NO profiles for Case 2.

Fig. 21. Effect of changes in the CH 4 mass percentage in the reburning fuel on predicted NO concentrations for Case 2.

Fig. 22. Effect of changes in the reburning injection velocity on predicted NO concentrations for Case 1.

NO x control through reburning

6.4. Injection Velocity


Figure 22 shows the effects of changes in the velocity
used for injection of the reburning fuel on the predicted

403

NO concentrations. The average NO concentration is


shown as a function of axial location in the furnace for
three different injection velocities compared with the
measured values. This figure shows that as the injection

Fig. 23. (a) Comparison of measured and predicted gas temperatures as a function of axial location in the reactor for Case 1. (b)
Comparison of measured and predicted CO 2 and O 2 concentrations as a function of axial location for Case 1. (c) Comparison of
measured and predicted average NO concentrations as a function of axial location for Case 1.

404

L. D. Smoot et al.

velocity is increased, NO reduction from reburning also


increases. This probably results because, at higher
injection velocities, the CH 4 is able to penetrate further
into the reactor before it is consumed, which results in a

larger region where the reburning reaction has an effect.


This is further supported by a reduction of the predicted
NO values just upstream of the reburning injection due to
greater spreading of the jet.

Fig. 24. (a) Comparison of measured and predicted temperatures as a function of axial location in the reactor for Case 2. (b)
Comparison of measured and predicted CO 2 and O 2 concentrations as a function of axial location for Case 2. (c) Comparison of
measured and predicted average NO concentrations as a function of axial location for Case 2.

NO x control through reburning

6.5. Comparisons with Data


In the following simulation, the P1 distribution (Table
5) was used with a mass mean particle diameter of
80 mm. A and E values for these calculations were 2.72 3
10 6 and 0.75 3 10 4, respectively. Figure 23 shows
comparisons between measured and predicted average
temperature, CO 2, O 2, and NO profiles for Case 1. Figure
23(a) shows the comparison of average measured and
predicted gas temperatures as a function of axial location
in the reactor. The rapid increase and subsequent
decrease of temperature at approximately 1 m is where
the tertiary air is injected. The predicted temperature
profile was adjusted using a constant wall temperature to
match the measured data, with a wall temperature of
950 K giving best results. Figure 23(b) shows the
comparisons of average measured and predicted CO 2
and O 2 concentrations as a function of axial location in
the reactor. The measured and predicted concentrations
show similar effluent magnitudes but somewhat different
trends, compared with the experimental data. This
suggests that the particle reactions occurred more
slowly in the simulations than was observed experimentally. Figure 23(c) shows the comparisons of average
measured and predicted NO concentrations as a function
of axial location in the reactor. The predicted profile
shows good agreement with the measured values and
indicates a rapid decay of NO just downstream of the
reburning gas injection location.
Figure 24 shows comparisons of measured and
predicted average values of temperature, CO 2, O 2 and
NO concentrations for Case 2, for a wall temperature of
1450 K. Figure 24(a) shows good comparison of
averaged measured and predicted gas temperatures as a
function of axial location in the reactor. Figure 24(b)
shows comparisons of average measured and predicted
CO 2 and O 2 concentrations as a function of axial location
in the reactor. As in Case 1, these comparisons show
similar trends, but the predicted CO 2 formation and O 2
consumption are slower than those observed experimentally. Figure 24(c) shows the comparisons of average
measured and predicted NO concentrations as a function
of axial location in the reactor. This simulation was
performed with E 0.75 3 10 4 and the pre-exponential
value reported by Chen72, and shows good agreement
between measured and predicted values of NO concentration. This newer work93 casts doubt on the usefulness
of the activation energy reported by Chen72 for generalized reburning computations.

405

programs. Typically, 1030% of the total fuel is used as


the reburn fuel.
Reburning with natural gas, without other NO x control
technologies, can typically provide 3565% reduction in
effluent NO x emissions. Simultaneous reductions in SO x
with addition of calcium-based sorbents, and in CO 2,
have been noted. Recent work suggests that joint
application of other NO x control technologies such as
low-NO x burners, staged combustion, urea or ammonia
injection (advanced reburning) may be able to reduce
emissions by up to 85%10. Even newer concepts show
promise of even greater reductions of up to 95% without
SCR32.
Reburning technology has most commonly been
applied to coal-fired boilers, but has also been demonstrated for oil- and gas-fired incineration systems and
waste fuels. In support of rapidly developing reburning
technologies, substantial work on laboratory measurements, kinetic mechanisms and comprehensive modeling
has been reported. Published and planned profile
measurements in laboratory reactors provide a sound
basis for model evaluation. Idealized predictions with
fundamental and reduced sets of kinetic reactions
describe some of the observed features of reburning.
Some of the first reburning predictions with comprehensive combustion codes, making use of reduced kinetic
schemes or newly reported global rate constants for
reburning reactions, are being reported. Refinement,
improvement and evaluation of these new predictive
tools are required.
Some test results show low-rank coal chars to be
particularly effective reburning fuels, but much remains
to be done to clarify the differences in reburning fuels
and associated causes. Further research and demonstration for highly efficient combined NO x reduction
technologies and new advanced reburning technologies
are also needed, some of which was ongoing at this
writing.
AcknowledgementsSome of this research and preparation of
this review paper were sponsored by the Advanced Combustion
Engineering Research Center. Funds for this center are received
from the National Science Foundation (Engineering Education
and Centers Division), the State of Utah (Centers of Excellence), 37 industrial participants, and five federal agencies. The
authors express their gratitude to all of the above, the University
of Utah, Dr Dale Tree, Mechanical Engineering at BYU, Dr
William Bartok, Consultant, and Mr Jerry Cole, EER. The work
of Mrs Eva Black in preparing this manuscript is greatly
appreciated.

REFERENCES
7. SUMMARY

Reburning, the process whereby gaseous, liquid or


solid hydrocarbon is injected downstream of the
combustion zone to reduce NO to HCN, is a commercialized technology. First tested in about 1950 and
named in 1973, it has been commercially demonstrated
in very large-scale utility boilers. It has also been
demonstrated in four of the Clean Coal Technology

1. Smoot, L. D., Boardman, R. D., Brewster, B. S., Hill, S. C.


and Foli, A. K., Development and application of an acid
rain precursor model for practical furnaces. Energy &
Fuels, 1993, 7, 786795.
2. Wendt, J. O. L., Fundamental coal combustion mechanisms and pollutant formation in furnaces. Prog. Energy
Combust. Sci., 1980, 6, 201222.
3. Boardman, R. D. and Smoot, L. D.,. Pollutant formation
and control, in Fundamentals of Coal Combustion, ed. L.
D. Smoot. Elsevier, The Netherlands, Chapter 6 (1993).

406

L. D. Smoot et al.

4. Miller, J. A. and Bowman, C. T., Mechanism and


modeling of nitrogen chemistry in combustion. Prog.
Energy Combust. Sci., 1989, 15, 287338.
5. Hayhurst, A. N. and Lawrence, A. D., Emissions of nitrous
oxide from combustion sources. Prog. Energy Combust.
Sci., 1992, 18, 529552.
6. Bowman, C. T., Control of combustion-generated nitrogen
oxide emissions: Technology driven by regulation, in
Twenty-Fourth Symposium (International) on Combustion. The Combustion Institute, Pittsburgh, PA, pp.
859878 (1993).
7. Kramlich, J. C. and Linak, W. P., Nitrous oxide behavior
in the atmosphere and in combustion and industrial
systems. Prog. Energy Combust. Sci., 1994, 20, 149202.
8. Mereb, B. J. and Wendt, J. O. L., Reburning mechanisms
in a pulverized coal combustor, in Twenty-Third
Symposium (International) on Combustion. The
Combustion Institute, Pittsburgh, PA, pp. 12731279
(1990).
9. Folsom, B. A., Sommer, T. M., Ritz, H., Pratapas, J.,
Bautista, P. and Facchiano, T., Three gas reburning field
evaluations. Final results and long term performance, in
EPRI/EPA 1995 Joint Symposium on Stationary
Combustion NO x Control, Kansas City, MO, May 19
(1995).
10. Folsom, B. A., Payne, R., Moyeda, D., Vladimir
Zamansky and Golden, J., Advanced reburning with new
process enhancements, in EPRI/EPA 1995 Joint
Symposium on Stationary Combustion NO x Control,
Kansas City, MO, May 19 (1995).
11. Pratapas, J., Bartok, W. and Folsom, B. A., Reburning, in
Encyclopedia of Environmental Analysis and Remediation, John Wiley & Sons, New York (in preparation)
(1997).
12. Makansi, J., Special report: Reducing NO x emissions from
todays powerplants. Power, 1993, May, 1128.
13. Pratapas, J. and Bluestein, J., Natural gas reburn: cost
effective NO x control. Power Eng., 1994, 98, 4750.
14. Patry, M. and Engel, G., Formation of HCN by the action
of nitric oxide on methane at atmospheric pressure, 1.
General conditions of formation. Compt. Rend., 1950, 231,
13021304.
15. Drummond, L. J., Shock induced reactions of methane
with nitrous and nitric oxides. Bull. Chem. Soc. Japan,
1969, 42, 285.
16. Wendt, J. O. L., Sternling, C. V. and Matovich, M. A.,
Reduction of sulfur trioxide and nitrogen oxides by
secondary fuel injection, in Fourteenth Symposium
(International) on Combustion. The Combustion Institute,
Pittsburgh, PA, pp. 897904 (1973).
17. Myerson, A. L., The reduction of nitric oxide in simulated
combustion effluents by hydrocarbonoxygen mixtures, in
Fifteenth Symposium (International) on Combustion. The
Combustion Institute, Pittsburgh, PA, pp. 10851092
(1974).
18. Takahashi, Y., Sakai, M., Kunimoto, T., Ohme, S.,
Haneda, H., Kawamura, T. and Kaneko, S., in Proceedings
of the 1982 Joint Symposium on Stationary NO x Control.
EPRI Report No. CS-3182, 1, July (1983).
19. Babcock & Wilcox Company, Demonstration of coal
reburning for cyclone boiler NO x control. Comprehensive
Report to Congress Clean Coal Technology Program,
DOE/FE-0157, February (1990).
20. Chagger, H. K., Goddard, P. R., Murdoch, P. and
Williams, A., Effect of SO 2 on the reduction of NO x by
reburning with methane. Fuel, 1991, 70, 11371142.
21. Folsom, B. A., Sommer, T. M. and Payne, R., Demonstration of combined NO x and SO 2 emission control
technologies involving gas reburning, in AFREJFRC
International Conference on Environmental Control of
Combustion Processes, Honolulu, HI, October 710
(1991).
22. Energy and Environmental Research Corporation,

23.

24.
25.

26.

27.

28.
29.

30.

31.

32.

33.
34.

35.

36.

37.

Evaluation of gas reburning and low-NO x burners on a


wall-fired boiler. Comprehensive Report to Congress
Clean Coal Technology Program, Irvine, CA, September
(1990).
PETC Review, Update on NO x control technologies.
Office of Fossil Energy, U.S. Department of
Energy, Pittsburgh Energy Technology Center, P.O. Box
10940, Pittsburgh, PA 15236-0940, Spring, pp. 2829
(1995).
Power Generation Tech Update, a publication of GRI on
the use of natural gas in utility electric power generation,
2(1) (1994).
Pratapas, J., Gas reburning GRI program overview, in
ProceedingsInternational Gas Reburn Technology
Workshop, ed. W. Bartok, Malmo, Sweden, February 7
9, pp. D328 (1995).
Azevedo, J. L. T., Costa, M. and Carvalho, M. G., Center
of heat and mass transfer in radiating and combusting
systems, in Smoot, L. D., International activities in coal
combustion research centers, Prog. Energy Combust. Sci.
1998, 24, 409501.
Beshai, R. Z., Light, M. E., Sanyal, A., Folsom, B. A. and
Payne, R., First and second-generation gas reburning
technology for NO x control. ASME Paper 94-JPGCFACT-3, New York, pp. 110 (1994).
(CCT) Clean Coal Technology Program Demonstration
Program, Program Update 1994, DOE/FE0330. U.S.
Department of Energy, Washington, DC, April (1995).
Lanigan, E. P., Golland, E. S. and Rhine, J. M., The
demonstration of gas reburning at Longannet: leading the
world in low-NO x technology, in Proceedings
International Gas Reburn Technology Workshop, ed. W.
Bartok, Malmo, Sweden, February 79, pp. D121138
(1995).
Farzan, H., Maringo, G. J., Riggs, J. D., Yagiela, A. S. and
Newell, R. J., Reburning with Powder River Basin coal to
achieve SO 2 and NO x compliance, in Proceedings of the
Power-Gen Sixth International Conference, Dallas, TX,
pp. 175187 (1993).
Chen, S. L., Cole, J. A., Heap, M. P., Kreamlich, J. C.,
McCarthy, J. M. and Pershing, D. W., Advanced NO
reduction processes using NH and CN compounds in
conjunction with staged air addition, in Twenty-Second
Symposium (International) on Combustion. The
Combustion Institute, Pittsburgh, PA, pp. 11351145
(1988).
Zamansky, V. M., Ho, L., Maly, P. M. and Seeker, W. R.,
Reburning promoted by nitrogen- and sodium-containing
compounds, in Twenty-Sixth Symposium (International)
on Combustion. The Combustion Institute, Pittsburgh, PA,
pp. 20752082 (1997).
Chen, S. L., Lyon, R. K. and Seeker, W. R., Advanced
non-catalytic post combustion NO x control. Environ.
Prog., 1991, 10, 182185.
Karll, B., Gas reburning and gas injection in combustion
with SNCR in a waste incineration plant, in ProceedingsInternational Gas Reburn Technology Workshop,
ed. W. Bartok, Malmo, Sweden, February 79, pp. D109
119 (1995).
De Angelo, J. G. and Sjoberg, C. E., The effect of coal
quality on meeting the ozone season NO x cap at New York
State Electric and Gas, in Tenth Annual Technical
Conference at Advanced Combustion Engineering
Research Center, Brigham Young University, Provo,
UT, March 68 (1996) [Prog. Energy Combust. Sci., in
press (1997)].
Marion, J. L., Laflesh, R. C., Towle, D. P., Benanti, A.,
Tarli, R., Demichele, G., Galli, G., Piantanida, A. and
Mainini, G., NO x emissions control for utility coal, oil and
gas fired boilers. Presented at the Joint ASME/IEEE
Power Generation Conference, Boston, MA, October 21
25 (1990).
Wendt, J. O. L. and Mereb, J. B., Nitrogen oxide

NO x control through reburning

38.

39.

40.

41.
42.

43.
44.

45.

46.

47.
48.

49.

50.

51.

52.

53.

54.

abatement by distributed fuel addition. DOE Final Report,


DE-AC22-87PC79850, University of Arizona, Tucson,
AZ, September (1991).
Borio, R. W. and Thornock, D. E., ABBs activities in the
reburn technology area, in ProceedingsInternational
Gas Reburn Technology Workshop, ed. W. Bartok, Malm
o, Sweden, February 79, pp. D187218 (1995).
Bilbao, R., Alzueta, M. U. and Millera, A., Simplified
kinetic model of the chemistry in the reburning zone using
natural gas. Indust. Eng. Chem. Res., 1995, 34, 4540
4548.
Tennessee River Valley Authority, Micronized coal
reburning demonstration for NO x control on a 175-MWe
wall-fired unit. Comprehensive Report to Congress Clean
Coal Technology Program, DOE/FE-0256P, June (1992).
Syverud, T., Thomassen, A. and Gautestad, T., Utilization
of chipped car tyres for reducing NO x emissions in a
precalciner kiln. World Cement, 1994, 25, 3943.
Wendt, J. O. L., Mechanisms governing the formation and
destruction of NO and other nitrogenous species in low
NO coal combustion systems. Combust. Sci. Tech., 1995,
108, 323344.
Kicherer, A., Spliethoff, H., Maier, H. and Hein, K. R. G.,
The effect of different reburning fuels on NO-reduction.
Fuel, 1994, 73, 14431446.
Chen, W. Y. and Ma, L., Importance of heterogeneous
mechanisms during reburning of nitrogen oxide, in Third
Symposium (International) on Coal Combustion, Beijing,
China, pp. 594601 (1995).
Moyeda, D. K., Li, B., Maly, P. and Payne, R.,
Experimental/modeling studies of the use of coal-based
reburning fuels for NO x control, in Pittsburgh Coal
Conference, Pittsburgh, PA, pp. 11191124 (1995).
Payne, R., Moyeda, D. K., Maly, P., Glavicic, T. and
Weber, B., The use of pulverized coal and coalwater
slurry in reburning NO x control, in Proceedings of the
EPRI/EPA 1995 Joint Symposium on Stationary NO x
Control, Kansas City, MO (1995).
Smart, J. P. and Morgan, D. J., Effectiveness of multi-fuel
reburning in an internally fuel-staged burner for NO x
reduction. Fuel, 1994, 73, 14371442.
Chen, S. L., Kramlich, J. C., Seeker, W. R. and Pershing,
D. W., Optimization of reburning for advanced NO x
control on coal-fired boilers. JAPCA, 1989, 39, 1375
1379.
Spliethoff, H., Greul, U., Rudiger, H. and Hein, K. R. G.,
Basic effects on NO x emissions in air staging and
reburning at a bench-scale test facility. Fuel, 1996, 75,
560564.
Greul, U., Spliethoff, H., Magel, H.-C., Schnell, U.,
Rudiger, H., Hein, K. R. G., Li, C.-Z. and Nelson, P. F.,
Impact of temperature and fuel-nitrogen content on fuelstaged combustion with coal pyrolysis gas, in TwentySixth Symposium (International) on Combustion. The Combustion Institute, Pittsburgh, PA pp. 22312239 (1996).
Brouwer, J., Heap, M., Owens, W. and Pershing, D. W.,
Advanced low NO x concepts for high efficiency power
generation systems, in Tenth Annual ACERC Technology
Conference, Salt Lake City, UT, March (1996).
Brouwer, J., Heap, M. P., Bales, F. E., Inkley, D. S.,
Lighty, J. S. and Pershing, D. W., The use of wood as a
reburning fuel in combustion systems, in Proceedings of
BioEnergy Conference, Reno, NV, October (1994).
Tree, D., Temperature, velocity and species profile
measurements for reburning in a pulverized, entrained
flow coal combustor. Contract DE-FG22-95PC95223,
Pittsburgh Energy Technology Center, USDOE,
Advanced Combustion Engineering Research Center,
Brigham Young University, Provo, UT (1995).
Nazeer, W., Detailed natural gas reburning measurements
in a pulverized coal reactor. MS thesis, Mechanical
Engineering Department, Brigham Young University,
Provo, UT (1997).

407

55. Griffiths, J. F., Reduced kinetic models and their


application to practical combustion system. Prog.
Energy Combust. Sci., 1995, 21, 25107.
56. Freklach, M., Wang, H. and Rabinowitz, N. J., Optimization and analysis of large chemical kinetic mechanisms
using the solution mapping methodcombustion of
methane. Prog. Energy Combust. Sci., 1992, 18, 4773.
57. Bian, J., Vandooren, J. and Van Tiggelen, P. J.,
Experimental study of the formation of nitrous and nitric
oxides in H 2/O 2/Ar flames seeded with NO and/or NH 3, in
Twenty-Third Symposium (International) on Combustion.
The Combustion Institute, Pittsburgh, PA, pp. 379386
(1990).
58. Warnatz, J., Resolution of gas phase and surface
combustion chemistry into elementary reactions, in
Twenty-Fourth Symposium (International) on Combustion. The Combustion Institute, Pittsburgh, PA, pp.
553579 (1992).
59. Bowman, C. T., Hanson, R. K., Davidson, D. F., Gardner,
W. C., Lissianski, V., Smith, G. P., Golden, D. M.,
Frenklach, M. and Goldenberg, M., http://www.me.berkeley.edu/gri_ mech/ (1996).
60. Glarborg, P. and Hadvig, S., Development and test of a
kinetic model for natural gas combustion. Nordic Gas
Technology Centre, Denmark, ISBN 87-89309-44-8
(1991).
61. Baulch, D. C., Cobos, C. J., Cox, R. A., Esser, C., Frank,
P., Just, Th., Kerr, J. A., Pilling, M. J., Troe, J., Walker, R.
W. and Warnatz, J., Evaluated kinetic data for combustion
modeling. J. Phys. Chem. Ref. Data, 1992, 21, 411737.
62. Thorne, L. R., Branch, M. C., Chandler, D. W., Kee, R. J.
and Miller, J. A., Hydrocarbon/nitric oxide interactions in
low-pressure flames, in Twenty-First Symposium (International) on Combustion. The Combustion Institute,
Pittsburgh, PA, pp. 965977 (1986).
63. Li, B. W., Wu, K. T., Moyeda, D. K. and Payne, R., Use of
computer models for reburning/cofiring boiler performance evaluations: combustion modeling, cofiring and
NO x control. ASME Fact, 1993, 17, 8794.
64. Hura, H. S. and Breent, B. P., Chemical kinetic simulation
of nitric oxide reduction during natural gas reburning in
pulverized coal fired boilers: combustion modeling,
cofiring and NO x control. ASME Fact, 1993, 17, 5169.
65. Burch, T. E., Tillman, F. R., Chen, W. Y., Lester, T. W.,
Conway, R. B. and Sterling, A. M., Partitioning of
nitrogenous species in the fuel-rich stage of reburning.
Energy and Fuels, 1991, 5, 231237.
66. Etzkorn, T., Muris, S., Wolfrum, J., Dembny, C.,
Bockhorn, H., Nelson, P. F., Atta-Shahin, A. and Warnatz,
J., Destruction and formation of NO in low pressure
stoichiometric CH 4/O 2 flames, in Twenty-Fourth
Symposium (International) on Combustion. The
Combustion Institute, Pittsburgh, PA, pp. 925932
(1992).
67. Stapf, D. and Leuckel, W., Flow reactor studies and testing
of comprehensive mechanisms for NO x reburning, in
Twenty-Sixth Symposium (International) on Combustion.
The Combustion Institute, Pittsburgh, PA, pp. 20832090
(1996).
68. Burch, T. E., Chen, W. Y., Lester, T. W. and Sterling, A.
M., Interaction of fuel nitrogen with nitric oxide during
reburning with coal. Combust. Flame, 1994, 98, 391401.
69. Kristensen, P. G., Glarborg, P. and Dam-Johansen, K.,
Nitrogen chemistry during burnout in fuel-staged combustion. Combust. Flame, 1996, 107, 211222.
70. Bilger, R. W., Turbulent flow with nonpremixed reactants,
in Topics in Applied Physics, Volume 44: Turbulent
Reacting Flows, ed. P. A. Libby and F. A. Williams.
Springer-Verlag, Berlin and Heidelberg (1980).
71. Drake, M. C. and Blint, R. J., Relative importance of nitric
oxide formation mechanisms in laminar opposed-flow
diffusion flames. Combust. Flame, 1991, 83, 185203.
72. Chen, W., Modeling of nitrogen pollutants in coal

408

73.

74.
75.
76.

77.

78.

79.

80.
81.

82.

83.
84.

85.

86.

87.
88.

L. D. Smoot et al.
combustion. PhD dissertation, Chemical Engineering
Department, Brigham Young University, Provo, UT (1994).
Chen, W., Smoot, L. D., Fletcher, T. H. and Boardman, R.
D., Part 1. A computational method for determining global
fuel-NO rate expressions. Energy & Fuels, 1996, 10,
10361045.
Chen, W., Smoot, L. D., Hill, S. C. and Fletcher, T. H., A
global rate expression for nitric oxide reburning. Part 2.
Energy & Fuels, 1996, 10, 10461052.
Zabarnick, S., A comparison of CH 4/NO/O 2 and CH 4/
N 2O flames by LIF diagnostics and chemical kinetic
modeling. Combust. Sci. Tech., 1992, 83, 115134.
Kee, R. J., Rupley, F. M. and Miller, J. A. ChemkinII: A
FORTRAN chemical kinetics package for the analysis of
gas-phase chemical kinetics. SANDIA Report, SAND898009.UC-401, Livermore, CA (1991).
De Soete, G. C., "Nitric oxides" formation and decomposition in the combustion of hydrocarbon flames (La
formation et la decomposition doxyde nitrique dans les
produits de combustion de flammes dhydrocarbures).
Rev. Pet. Inst. Fr., 1972, XXVII, 372.
De Soete, G. C., Mechanisms of nitric oxides from
ammonia and amines in hydrocarbon flames (Le
mechanisme de formation doxyde azotique a partir
dammoniac et damines dans les flammes dhydrocarbures). Rev. Pet. Inst. Fr., 1973, XXVIII, 171.
De Soete, G. C., Overall reaction rates of NO and N 2
formation from fuel nitrogen, in Fifteenth Symposium
(International) on Combustion. The Combustion Institute,
Pittsburgh, PA, pp. 10931102 (1975).
Glarborg, P., Miller, J. A. and Kee, R. J., Kinetic modeling
and sensitivity analysis of nitrogen oxide. Combust.
Flame, 1986, 65, 177202.
Glarborg, P., Lilleheie, N. I., Byggstyl, S., Magnussen, B.
F., Kilpinen, P. and Hupa, M., A reduced mechanism for
nitrogen chemistry in methane combustion, in TwentyFourth Symposium (International) on Combustion. The
Combustion Institute, Pittsburgh, PA, pp. 889898
(1992).
Peters, N., Reduced mechanisms, in Reduced Kinetic
Mechanisms and Asymptotic Approximations for
MethaneAir Flames, ed. M. D. Smooke. SpringerVerlag, Berlin, Chapter 3, pp. 4867 (1991).
Payne, R. and Moyeda, D. K., Scale up and modelling of
gas reburning. Combustion modeling, scaling and air
toxins. ASME Fact, 1994, 18, 115122.
Ballester, J., Fueyo, N. and Dopazo, C., Natural gas
reburning in coal-fired boilers: experiments and computations, in ProceedingsInternational Gas Reburn
Technology Workshop, ed. W. Bartok, Malmo, Sweden,
February 79, pp. D187218 (1995).
Rasmussen, N. B. K., Modelling of flow and combustion
in a MSW incineration plant, in ProceedingsInternational Gas Reburn Technology Workshop, ed. W.
Bartok, Malmo, Sweden, February 79, pp. D175185
(1995).
Tyson, T., Gas reburning design process considerations, in
ProceedingsInternational Gas Reburn Technology
Workshop, ed. W. Bartok, Malmo, Sweden, February 7
9, pp. D121138, D173 (1995).
Freklach, M., in Numerical Approaches to Combustion
Modeling, ed. E. S. Oran and J. P. Boris. AIAA, New
York, Chapter 5 (1991).
Fiveland, W., Personal communication. Babcock and
Wilcox, Alliance, OH (1996).

89. Power Generation Tech Update, a publication of GRI on


the use of natural gas in utility electric power generation,
November (1996).
90. Brouwer, J., Heap, M. P., Pershing, D. W. and Smith, P. J.,
A model for prediction of selective noncatalytic reduction
of nitrogen oxides by ammonia, urea and cyanuric acid
with mixing limitations in the presence of CO, in TwentySixth Symposium (International) on Combustion. The
Combustion Institute, Pittsburgh, PA, pp. 21172124
(1996).
91. Lyon, R. K., Kinetics and mechanism of thermal de NO x: a
review, 1987, 32, 433443.
92. Brouwer, J., Personal communication. Reaction Engineering International, Salt Lake City, UT (1997).
93. Xu, H. J., Investigation of NO x control through reburning.
PhD dissertation prospectus, Department of Chemical
Engineering, Brigham Young University, Provo, UT, in
process (1997).
94. Xu, H., Smoot, L. D. and Hill, S., A reduced kinetic model
for advanced reburning NO x process, in 1997 Fall Meeting
of Western States Section. The Combustion Institute,
Diamond Bar, CA 91765 (1997).
95. Furusawa, T., Tsunoda, M., Tsujimura, M. and Adschiri,
T., Nitric oxide reduction by char and carbon monoxide.
Fuel, 1985, 64, 13061309.
96. Teng, H., Suuberg, E. M. and Calo, J. M., Studies on the
reduction of nitric oxide by carbon: the NOcarbon
gasification reaction. Energy and Fuels, 1992, 6, 398406.
97. Illan-Gomez, M. J., Linares-Solano, A., Radovic, L. and
Salinas-Martinez de Lecea, C., NO reduced by activated
carbons. 2. Catalytic effect of potassium. Energy and
Fuels, 1995, 9, 97103.
98. Guo, F. and Hecker, W. C., Effects of CaO and burnout
on kinetics of NO reduction by the Beulah Zap char,
in Twenty-Sixth Symposium (International) on Combustion. The Combustion Institute, Pittsburgh, PA, pp.
22512257 (1996).
99. Yamashita, H., Yamada, H. and Tomita, A., Reaction of
nitric oxide with metal-loaded carbon in the presence of
oxygen. Applied Catalysis, 1991, 78, L1L6.
100. Shimizu, T., Sazawa, Y., Adschiri, T. and Furusawa, T.,
Conversion of char-bound nitrogen to nitric oxide during
combustion. Fuel, 1992, 71, 361365.
101. Yamashita, H. and Tomita, A., Influence of char surface
chemistry on the reduction of nitric oxide with chars.
Energy and Fuels, 1993, 7, 8589.
102. Levy, J., Chan, L. K., Sarofim, A. F. and Beer, J. M., NO/
char reactions at pulverized coal flame conditions, in
Eighteenth Symposium (International) on Combustion.
The Combustion Institute, Pittsburgh, PA, pp. 111120
(1981).
103. Hill, S. C. and Smoot, L. D., A comprehensive threedimensional model for simulation of combustion systems:
PCGC-3. Energy and Fuels, 1993, 7, 874883.
104. Brewster, B. S., Eaton, A. M. and Boardman, R. D., Users
manual 93-PCGC-2: pulverized coal gasification and
combustion model (2-dimensional) with a generalized
coal reactions submodel (FGDVC). Final Report, Vol. 2,
DOE DE-AC21-86MC23075, Brigham Young University,
Provo, UT (1993).
105. Hill, S. C. and Smoot, L. D., Modeling of NO x formation
in coal combustion. Prog. Energy Combust. Sci., in
preparation (1997).
106. Wendt, J. O. L., Personal communication. University of
Arizona, Tucson, AZ (1996).

S-ar putea să vă placă și