You are on page 1of 22

Prog. Energy Combust. Sci.. Vol. 6, pp. 201 222 (C) Pergamon Press Ltd., 1980.

Printed in Great Britain.

0360 1285/80/0701 0201505.00/0

F U N D A M E N T A L COAL C O M B U S T I O N MECHANISMS A N D P O L L U T A N T F O R M A T I O N IN FURNACES


J. O. L. WENDT Department of Chemical Enyineeriny, University of Arizona, Tucson, Arizona 85721, U.S.A.

1. I N T R O D U C T I O N

1.1, Motivation

Pulverized coal has been utilized as a fuel for over 50 years. Furthermore, there exist in operation throughout the world thousands of combustion units designed to raise steam, utilizing pulverized coal as the primary fuel. On the whole these units work well and efficiently. This being the case, one may question the need to undertake research programs directed towards the understanding of fundamental mechanisms during pulverized coal combustion. It is therefore appropriate to delineate why research activity in this area has grown, and what new facets of the overall problem have emerged to require a deeper understanding of the fundamental processes involved. Three new problem areas can be identified. First, a rapidly changing energy supply situation has forced operators of combustion equipment designed for liquid or gaseous fuels to consider switching to solid fuels. 1 The operator is then forced to determine: (1) whether he should burn conventional pulverized fuel after incorporating drastic design changes in his equipment, (2) whether he can use his existing unit but at reduced load, or (3) whether he should attempt to fire the solid fuel in an unconventional mode, such as with ultra-fine particles (less than 400 mesh) or mixed with oil to form a slurry. The operator will find that basic technical information to guide his decisionmaking process is presently unavailable. Second, increased utilization of solid fuel for new combustion units has led to the introduction of coal types different (usually of lower quality) from those used in the past. The designer uses his experience with conventional coal-types, and hopes that the untested coal will cause no unusual problems. Predictive tools to extrapolate experience gleaned from one coal-type to one radically different are unavailable. 2 Third, and most important, environmental considerations have required combustion modifications for pollutant emission abatement. Nitrogen oxide abatement can be accomplished economically only through prevention of their formation during the combustion process, and not through stackgas removal. Recent research 3 has confirmed that the chemically bound nitrogen in coal is the major contributor to NOx emissions from pulverized coal combustion and that formation of this byproduct can be radically modified through proper burner design and control of air and fuel contact. However, the basic
IP[(1 ~

information leading to development of pulverized coal burners to give very low fuel-nitrogen conversions to NOx (leading to less than 100ppm) is only now becoming available. Because this facet of the problem focusses on the fate of a trace species in coal, and on a tiny byproduct of reaction, a sophisticated understanding of chemical and physical principles involved is required to help development of the necessary predictive tools. These tools should enable a designer to determine the effects of changes in both hardware and solid-fuel composition on pollutant emissions. Also, they should allow the government regulatory agency to determine the maximum level of pollutant abatement achievable through combustion modifications. An understanding of mechanisms controlling the formation of these pollutants will provide insight into what coals can be burned safely, and how such combustion should proceed. Mechanisms of pollutant formation are inextricably intertwined with mechanisms of coal combustion, and so in addressing the problem of pollutant formation mechanisms a researcher will automatically address problems involving the very essence of coal combustion itself. Therefore, although this paper focusses primarily on the fate of fuel-nitrogen and other trace species during the coal combustion process, it also addresses the general problem of fundamental mechanisms in coal combustion. In this way it is hoped to shed light both on how knowledge of fundamental mechanisms can be used to solve a specific pertinent problem and on what types of mechanisms play important roles. This paper should therefore be viewed neither as a comprehensive review of what is known about coal combustion nor as a review of homogeneous pollutant formation mechanisms; such reviews are available and are discussed in a subsequent section. Rather, it is the objective of this paper to : (1) identify fundamental mechanisms that are important, (2) to delineate why they are important, and (3) to show how knowledge of these mechanisms can help solve the practical problems of fuel switching and pollutant abatement outlined above.

1.2. Previous Reviews Considerable research on the combustion of pulverized coal has been conducted since Faraday's work 4 on coal-mine explosions in 1845. A useful starting point for researchers can be found in the monograph by 201

202

J.O.L. WI~NDT that little is known about coal behavior, and so some elementary concepts are explained. First, in order to provide the proper background in which mechanisms can be discussed, three different physical pulverized coal combustion configurations are described. These are presented from the viewpoint of how they affect the environment in which coal particles burn. Detailed hardware descriptions are outside the scope of this paper; rather, the focus is on fundamental processes. Then, the salient features of combustion modifications for pollutant abatement are described. Following this, important aspects of coal characterization, regimes of combustion, volatilization, volatiles combustion and char burnout are discussed in turn. In each case, relationships between important features of the process and various combustion configurations are identified. What is known, and what is not known is delineated, and some fundamental modeling approaches are evaluated and described. 2. COMBUSTIONCONFIGURATIONS The primary purpose of any combustion configuration is to burn a fuel in such a way that stable t i m e s produce a specified heat-flux distribution with a minimum pollutant emission. These three constraints--stability, heat flux and pollutant emission--are usually conflicting, and the challenge presented to the combustion researcher is to allow a practical optimum to evolve. In both wall-fired and tangentially-fired configurations, coal is continuously pulverized, 75% by weight through a 200 mesh screen, (mean particle size 60 #m) and transported at a velocity of approximately 15 m/sec by primary air to the burners, to be described subsequently. Primary air constitutes approximately 20% of the total combustion air. Cyclone-fired configurations do not utilize pulverized coal, but rather use crushed coal (95% through a 4 mesh screen, 0.5 cm size particles). Here, gravity is used to feed the coal, which is mixed with primary air just prior to entering the burner described in a separate section below. 2.1. Wall Fired A schematic of a wall-fired swirl-stabilized flame is shown on Fig. 1. The primary-air/pulverized-fuel mixture passes through an injector, from which the coal jet fans out using a spreader. The spreader imparts radial momentum to the coal a.nd this allows fairly rapid contact with secondary air. The primary recirculation zone is a result of local pressure variations caused by swirl, and is of critical importance in maintaining a stable flame attached to the injector. In the sketch on Fig. 1, hot gases flow towards the incoming fuel, to ignite it continuously and thus maintain a Type III flame. 11 Variations in swirl momentum and burner geometry may cause this recirculation zone to move downstream and/or change

Field et al.,s and published by the British Coal Utilization Research Association. The purpose of this book was to review the background information needed for constructing mathematical models of pulverized-coal-fired combustion chambers. This reference contains a wealth of useful data and methods that had been developed through 1967. No consideration of byproducts of combustion or of pollutant emissions is given however. A list of 110 references on pulverized coal firing and published in the Journal of the Institute of Fuel is given in the brief review of Gill, 6 where the primary emphasis is on accelerated heattransfer and on ash control. Devolatilization of coal and the companion problem of hydrogasification have been exhaustively treated in a recent review by Anthony and Howard, 7 while char combustion is the focus of another review by Laurendeau a in this journal. Essenhigh's critical review 2 of combustion and flame propagation in coal systems contains a list of 160 references, although none of these is concerned with NOx or pollutant emissions. Useful information on combustion kinetics, is also available in the review by Mulcahy and Smith. 9 That review focusses on both overall burning rates and the burnout of porous charparticles. The fate of trace species is not considered. Mechanisms of pollutant formation during pulverized coal combustion have been reviewed by Malte and Rees. 1 In that timely and useful review, the fundamental aspects of homogeneous gas-phase reactions of nitrogen and sulfur compounds were discussed, as well as their relationship to coal combustion behavior. Flyash and particulate formation were also described and 102 references were cited. 1.3. Scope of Present Paper The purpose of this paper is to delineate how fundamental mechanisms in coal combustion will aid in the design of future pulverized coal combustion configurations that allow for both fuel variability and low pollutant emissions. The appropriate mechanisms are identified and the reasons why they are important delineated. The present state of knowledge of these mechanisms is summarized and critical gaps are identified. Liberal use is made of available literature reviews, to which the reader is referred for more detailed lists of references. This paper differs from these reviews in that it is restricted only to those specific mechanistic aspects that affect the practical engineer or designer in his/her attempt to meet the objectives outlined above. Special attention is given to the fate of trace species. The relationship between practical application and fundamental mechanisms is a central feature of this work. It is hoped that this paper will be of value to both the practical engineer and to the researcher: the practising engineer because knowledge of some important intricacies of coal behavior will provide the insight required for future design concepts; the researcher because challenging but useful fundamental research problems will be identified. In either case, it is assumed

Fundamental coal combustion mechanisms and pollutant formation in furnaces


I \

203

Water cooled wol Is

i COOL I \\ \

\ \ \ \ \ \\

\ \ \ \ \ \ \

(i) (~ ~

Free Shear Layer Ax,ol Region-For Flelct


Phenomeno

Neor F~eld Phenomeno

//

C//OOL

FIG. 1. Wall-fired configuration. its shape, or to be penetrated by the primary fuel jet to form a Type II flame, 11 described elsewhere. Rather than enter into a detailed discussion of burner aerodynamics let us examine possible environments surrounding coal particles, and how these environments might influence how the particles behave. Initially, at the injection point, all the particles are in contact with primary air. The particles receive heat from two sources: (1) radiation from the flame downstream, or from hot walls, and (2) convection from hot gases recirculated in the primary recirculation zone. Regardless of which mechanism dominates, the particles heat up at a rate of 105-106 K/sec, while the fuel jet still maintains some of its integrity. Particle ignition occurs at a temperature of 1300 K or so, possibly prior to significant devolatilization. 12 Gases devolatilize from the coal, then ignite and react directly with the oxygen in the primary air stream. If nitrogenous species were evolved during this initial devolatilization phase, the intimate contact with available oxygen would allow significant fuel-NO formation.: 3 If, however, the oxygen in the primary air stream were consumed by nitrogen-free volatiles, and if some of the nitrogen were evolved (say, in Region (1) on Fig. 1), where there is a paucity of available oxygen, then the possibility exists whereby fuel-N will pyrolyze to form N2. Early contact of nitrogeneous species with oxygen should be avoided in order to minimize their oxidation to NO. Increasing the quantity of primary air (per mass of coal) would lead to two different effects : first, it would increase the quantity of oxygen immediately available for reaction with those volatiles leaving the coal particle early; second, it would increase the axial momentum imparted to the fuel jet, thus possibly splitting and penetrating the recirculation zone (possibly blowing the flame off) and/or allowing the increased shear between primary fuel jet and secondary air to accelerate mixing between these streams. Wendt and Pershing 14 reported that replacement of primary oxygen by some inert CO2, without changing flow-rates, caused a marked change in flame characteristics and loss of stability. This was because lack of primary oxygen delayed particle ignition and volatile combustion until secondary air mixing had occurred. Clearly, primary air plays an important role in coal particle ignition and flame stability, and not merely in aerodynamic transport. One might expect coal particle size to be important for wall-fired units. Small particles heat up rapidly, and tend to follow the fluid streamlines. Larger particles heat up slowly and will be ballistically transported into regions (such as the recirculation zone or Region (1)) where heat transfer by convection dominates. In the recirculation zone there may be an oxygen-deficient atmosphere and the fate of volatiles and trace species evolved is uncertain. The fate of coal particles ballistically transported into the free shear-layer (Region (1)) and of volatile combustion in that neighborhood would appear to be an area worthy of further study. 2.2. Tangentially Fired Pure axial pulverized-coal (Type I) flames 11 without swirl can be stabilized in a cold-wall furnace only with difficulty. This is because, in the absence of wall radiation or hot product recirculation, the essential heat feedback mechanism to allow sufficiently high initial heating-rates is lacking. In hot-wall, Kiln furnaces, a combination of wall radiation and convection of hot burned gas entrained into the axial jet is sufficient to ignite the fuel and maintain stability. The tangentially- or corner-fired configuration, shown in Fig. 2 successfully avoids stability problems with axial flames by allowing flame-flame interaction, such that each flame in a sense supports the other. Although Fig. 2 shows a photograph of an oil flame, the essential characteristics of a pulverized coal flame

204

J.O.L. WENOT

FIG. 2. Tangentially-fired configuration (courtesy of Combustion Engineering, Inc.)

in the tangentially-fired mode are similar. One important difference, however, is that in a pulverized-coal flame the luminous zone of the flame jet is usually detached from the burner by several burner diameters. The pulverized-coal/primary-air mixture, and the secondary air are introduced at the corners through straight axial pipes with no swirl. Primary and secondary air flows are roughly equal (30 m/sec), with little shear between them. Mixing between them is therefore slow. Whether initial heating-rate of the coal particle is determined by radiation or by convection is not clear. Certainly, it would appear that the coal particles at the outside of the jet ignite and burn prior to those in the inside, and so volatilization may not be complete prior to the flame reaching the doughnut portion of the combustion, shown in Fig. 2. It should, however, be pointed out that the overall aerodynamics in a tangentially-fired furnace are quite complex. Important secondary flows make the overall flow-field quite three dimensional. In considering the fate of fuel-N it is logical to assume that some volatile nitrogeneous species are evolved at the outside edge of the jet and are therefore in intimate contact with oxygen, while a larger portion of the volatile nitrogen is evolved in the fuel-rich core of the jet. This has great implications when comparing

fuel-NO emissions from tangentially-fired to those from wall-fired configurations. As in the wall-fired case, increase of primary air (transporting the coal) would have two effects: (1) the quantity of oxygen available for the reaction with volatiles is immediately increased, and (2) the shear between the primary fuel-jet and the secondary air is also increased, thus allowing more rapid mixing between the two. It should be possible to quantify these concepts through the use of mathematical models describing the interaction of two concentric axial jets, coupled with global descriptions of devolatilization and combustion. Some combustion modification may be possible through changes in burner spacing and air and fuel inlets. Progress in modeling the heat-transfer characteristics of tangentially-fired furnaces has been good. 15 Prediction of NOx formation and/or the combustion behavior of different solid fuels when fired in the tangentially-fired mode has not yet been achieved.

2.3.Cyclone Fired
The cyclone-fired combustion configuration is quite different from the other two mentioned above. In a cyclone furnace, coal combustion occurs in a pre-

Fundamental coal combustion mechanisms and pollutant formation in furnaces chamber at very high intensities, such that the walls of this precombustion chamber are coated with molten slag. Crushed coal is fed into this chamber with a vortex motion. The larger coal pieces are imbedded in the molten slag at the surface and combustion occurs with a large fraction of the fuel so entrained. The smaller particles are entrained in the gas stream and burnt in a similar mode as in wall-fired units. Mixing between secondary air and any fuel not imbedded in the slag is very rapid because of the high vortex motion imparted to all incoming streams. The release of heat in the cyclone furnace is very high. Moreover, any heattransferring refractory surface is partially insulated by the covering slag layer. This combination of high heat release and low heat absorption assures the high temperatures necessary to complete combustion and to provide the desired liquid-slag covering of the surface. The gaseous products of combustion are discharged through a water-cooled throat into a water-walled furnace, where the bulk of the heat transfer takes place.
I I I DIVERGENT INJECTOR CRAWFORD, ET AL. (1974, 1976) WALL FIRED UNITS

205

the ash is easily collected by tapping off the slag. The disadvantage is that this mode of combustion can be used primarily with coals having fairly low ash fusiontemperatures. Another problem is the resulting high level of NOx emissions probably due to the rapid contacting between suspended fuel and air. Future work is required to determine combustion modifications in the cyclone mode for NO~ control, and possibly the capture of sulfur through the use of alkali metals that tend to lower the ash fusiontemperature. The consequences of operating the cyclone prechamber at a fuel-rich condition might be investigated, although the ensuing materials problems may be insurmountable. 2.4. Pollutant Abatement Through Aerodynamic Control A major constraint in the burning of coal is the emission of air pollutants. One such pollutant, NOx, can be removed from the stack gas only with difficulty,
I AXIAL INJECTOR I ------ C R A W F O R D ,ETAL (1974,1976) -I TANGENTIALLY FIREDUNITS
O

1200 -----I000

-o o o -/

~
0 :7

600 ti

-4

t
i

/"
/ / // / / t" / /'/ ~. ~" t

O,/,, /'" t 1 / t
i _

4 l

CL

400

f /1/
Oj/~-~ t~ ~

200

1.0

II

$2

I 13

1.0

i.

I I

II
RATIO

1.2

13

STOICHIOMETRIC

FIG. 3. Comparison of uncontrolled exhaust NO~ emissions from wall-fired and tangentially-fired pulverized coal units (Ref. Crawford et al., EPA Reports EPA-650/2-74-066 (1974) and EPA-600/2-76-152c (1976)). Clearly, fundamental mechansims involved in prediction of coal combustion behavior and in the formation of pollutants are likely to be quite different in this case than in the other two configurations discussed above. Very little is known about exactly how large coal particles imbedded in a molten slag layer burn. The metal content of the molten slag is likely to exert a strong catalytic influence on the gasification rate of coal particles so imbedded.16 Very little is also known about a characteristic residencetime distribution of the coal particles and how this varies with the inlet size distribution. The advantage of the cyclone furnace is that most of and the most promising method of control is that of combustion modification. With coal the primary source of NO~ emissions arises from oxidation of chemically bound nitrogen in the fuel. a Figure 3 shows field data on exhaust NOx emissions from wall-fired and tangentially-fired utility boilers. Clearly, tangentially fired configurations have, on the whole, lower NO emissions than do wall-fired configurations. Furthermore, it has been shown 3 that this difference is due to a difference in conversion of fuel-N to NOx, rather than to a difference in thermal NO~. Figure 3 shows that from an NOx point of view, wall-fired units can be simulated in the laboratory by a burner utilizing

206

J.O.L. WENDT
Q (~ (~ Free Sheor Loyer
Axlol Region-For Field Phenomeno

Neor F=eld Phenomeno

AXIAL TERTIARYAIR

_1

[~,NG

L~ ~ O ~ T ZOnE

I
Fro. 4. NO~ control through flame aerodynamic changes : The distributed mixing concept (Ref. 17). a divergent (or spreader type) injector, while tangentially-fired units can be simulated by a pure axial injector. Therefore, a major difference in the two configurations is contained in the mixing processes occurring in the immediate neighborhood of the injector, and this region has a strong influence on the conversion of coal-N to NOx. Low NOx emissions are favored where devolatilization of nitrogeneous species from coal occurs in the absence of oxygen. Type I flames (tangentially-fired) have fuel-rich regions that are hot and that allow pyrolysis of fuel-N and little NO. However, it cannot be ruled out that even in the fuel-rich core of the jet some NO may be formed by the available primary oxygen and that this NO may be subsequently reduced. This phenomenon of formation and subsequent reduction of N O is discussed further in Section 2.5. For wall-fired furnaces the potential for NOx abatement through flame aerodynamic control is great. This has been accomplished t7 (Fig. 4) through a combustion modification in which secondary air is further subdivided into secondary and tertiary air. In this case the secondary air is insufficient for complete combustion. It is given a high degree of swirl to allow recirculation for flame stability, but the zones denoted as 1 and 3 are essentially fuel-rich. The tertiary air is introduced at some radius far from the fuel injector and mixes slowly with the combustible mixture in Region 2. NOx emission data from a pilot-scale version of this, developed under contract from the U.S. Environmental Protection Agency, are shown in Fig. 5, where the abscissa denotes the equivalence ratio resulting from primary plus secondary air. It should be noted that the overall combustion is fuel-lean. Clearly, very low NOx emissions from pulverized coal can be achieved with this type of flame aerodynamic control. Variations of this design are presently being put into commercial use. 18 Even though uncontrolled NOx emissions from tangentially fired configurations are within present standards (see Fig. 3), the potential for further NOx control through aerodynamic modification may not be high because this configuration contains a fundamental lack of flexibility in burner parameters. At present, both primary fuel jets and secondary air jets emerge with a pure axial momentum and little relative velocity between them. Any variation in this would lead to increased shear between fuel and air streams, and therefore increase mixing and thus increase the conversion of fuel-N. It is not known how much fuel-N is converted prior to combustion in the doughnut and how much is converted in the doughnut itself. Likewise, the role of flame--flame interactions on possible reprocessing of NO and other nitrogeneous

8 O IO P U L V E R C Z I O E ~ ) A {L S M H O A M T :L E ~ U D S ; M W I w G R H H I L R w s L O T L I S U W N R N L IE L A o
/"/ 600

WATER-COOLED SIMULATOR 14,64 MW

,,% \

2 o o , 0 t 0 \
0.8 1.0 12 PRIMARY

i
I4

~
16 f8

Z O N E E Q U V IA L E N R C A E T O I

FIG. 5. Reduction of NO., emissions possible through flame

aerodynamic control (Ref. 17).

Fundamental coal combustion mechanisms and pollutant formation in furnaces species to form N 2 is not clear. Therefore, it is not clear whether future research on NO~ control from tangentially-fired furnaces should focus on aerodynamics of the axial jets or on the aerodynamics of the interactions between flames in the center of the furnace. Mathematical models describing how burner hardware controls the environment in which combustion takes place, and how coal devolatilization and combustion characteristics in each of the applicable environments (shown in Fig. 4) control coal-N conversion can help determine whether advanced burner designs can be further optimized to achieve yet lower levels of fuel nitrogen conversion. Models can also be utilized for scale-up. Insight derived from models is useful to generalize specific results, such as those shown on Fig. 5, and to see how they can be extrapolated to another coal combustion configuration.
1200

207

times are not desirable in practice, because of corrosion problems. Figure 6 shows results of staged combustion of pulverized coal for "premixed" coal-air mixtures in a one dimensional flow laboratory furnace. In this case all the coal is mixed with all the air prior to ignition. The first stage stoichiometric ratio was 0.4, which is very fuel-rich. N O profiles are plotted as a function of residence time. The vertical dashed lines correspond to three staging positions, indicating three different firststage residence-times. The salient phenomena are that : even under very fuel-rich conditions (SR = 0.4) some NO is formed rapidly and then reduced; reduction is slow; additional air, added in the second stage, can cause an increase in NO because nitrogeneous species (precursors to second-stage NO) pass through the first stage, to be oxidized subsequently. Therefore, increased NOx abatement does not always result from increasingly fuel-rich first stages (because of increasing

I000-

800--

5
600-

I
I

0.-

I
~ A JX I

/
I
00.0 0 2 0.4 0.6

c~
L.,,+0.B

oo

1.0 I2 TIME, sec 14 I6 1.8 20 2 2 24 26

FIG. 6. Staged combustion : NO profiles under fuel-rich and staged combustion of pulverized coal. First

stage SR = 0.4, second stage, SR = 1.2. Vertical dashed lines indicate three separate staging positions, separate experiments. 2.5. Staged Combustion In the staged combustion mode, fuel-rich and fuellean regimes of combustion are segregated through external divided air ports rather than through flame aerodynamic control. Staged combustion has been applied with some success in the field, 19 both for wallfired and tangentially-fired units. A fuel-rich zone in the furnace allows partial oxidation of the fuel and, presumably, devolatilization of nitrogeneous species in an oxygen-deficient environment. This is followed by a fuel-lean zone in which the unburnt fuel is combusted, but in which the temperatures are sufficiently low so that thermal NOx formation is insignificant. The critical questions to be answered are: what are the optimum temperatures, heat releaserates, residence times, and stoichiometric ratios in the first stage for maximum NOx abatement? Pilot scale experiments 2 have shown that very low NOx emissions from staged combustion of pulverized coal can result when first stage residence-times are long (on the order of 3 sec). However, such long fuel-rich residencequantities of second stage NO). Neither is it true that increasing residence-times in the first stage always leads to significantly lower exhaust NOx emissions. Figure 6 indicates that second stage N O can be only a weak function of first-stage residence-times, and that the rate of destruction of first-stage N O is slow. Subsequent data 21 on a similar system has shown that the behavior of the N O profiles at the staging point is a function of coal composition, and that even at the overall fuel-lean conditions at the staging point, N O can sometimes be destroyed. This destruction of N O has also been observed in a 3 M M Btu/lar pilot study, 22 although the mechanisms through which this occurs, and why this should depend on coal composition, are as yet unresolved. The fundamental phenomena of importance are: (1) the relative rates at which nitrogen is evolved compared to the rate of consumption of oxygen, since this controls the peak N O under fuel-rich conditions; (2) the subsequent homogeneous kinetics controlling destruction of fuel-N species that may act as N O precursors in the second stage; (3) the role of hetero-

208

J.O.L. WENDT percentage (to be defined below) on a dry ash-free basis, Higher Heating Value on a wet ash-free basis, and rank, for a large number of U.S. coals. Scatter is large, but several trends are apparent. In general, coals of increasing rank have increasing heating values and a higher fixed-carbon content. Likewise, increased fixedcarbon content is correlated with heating value. Figure 7 also denotes how coals are ranked into various subcategories. Note that neither fixed-carbon nor heating value alone can determine rank throughout the entire range. Table 1 shows the composition breakdown of a number of U.S. coals. The ultimate analysis gives the elemental composition on a dry but not ash-free basis. Therefore moisture is not included in the weight-%

geneous reaction reducing NO already formed, on char and/or ash surfaces; 23 (4) the mechanisms of N O destruction, when it occurs, at the staging point; and (5) the dependence of all these mechanisms on local environment, temperature and coal composition. An understanding ofthese phenomena can aid not only in determining an optimum staged combustion configuration, but also in the development of NOx control combustion modifications in general. Clearly, much still remains to be understood in order to enhance our confidence that combustion modifications for low NOx emissions can always be made to work, even though substantial progress has already been made with only an incomplete understanding of the phenomena involved.

I00

15 1

MOIST ASH FREE, HIGHER HEATING VALUE 106 d/kg 20 25 30 t I I

35 Meta Anthrocit~ ~

Anthracite

9O

Semi Anthracite LOw Volatile Bituminous

n..

so
_

B"~i~''"e

,'7

.5o
C3

40
Sub

Sub BituminousB

High Volatile C Bitum or Sub-bituminous A

High V01. B Bltum.

Hi(jh Volatde A Bituminous I60(X)

Bituminous C

6000

I 1 I0000 120(0 MOIST, ASH FREE. HIGHER HEATING VALUE Btu/Ib

14000

FIG. 7. Relationship between coal rank, fixed carbon and heating value.

3. C O A L C H A R A C T E R I S T I C S

3.1. Chemical Composition Coal is not a homogeneous fuel, but rather a substance whose constitution is not well understood. Several texts z4'25 on this subject are available. Here we will merely introduce some elementary concepts. Coal is formed as a result of vegetation (organic matter) being subjected to moisture, heat and pressure. The geological process is as follows : vegetable matter is transformed to peat, then to lignite, then to subbituminous coal, then to bituminous coal, and then to anthracite. These categories allow classification of coal by rank, depending on the geological age. Attempts have been made to correlate rank with composition. Figure 7 shows the relationship between fixed carbon

hydrogen. In Table I the Colorado, Pittsburgh and W. Kentucky coals are bituminous, while the Montana coal is sub-bituminous and the Texas lignite is of course of lower rank still. Note how sulfur content varies with rank, the lower-rank coals having (generally) lower sulfur contents. Note also how the lowerrank coals tend to have higher oxygen contents. It should also be noted that oxygen contents are determined only by difference, and tffus contain a sum of all the errors involved in the analyses for the other elements. Fuel-N contents vary over a fairly narrow range, however, and do not correlate with fuel-S. This may be important should SOx/NOx interactions be significant. The ash content is not equal to the mineral-matter content in the coal prior to the ASTM Ultimate

Fundamental coal combustion mechanisms and pollutant formation in furnaces TAaLE 1. Pulverized-fuel compositions Ultimate analysis (~o dry) C H N S O Ash Heating value (Btu/lb wet) Proximate analysis (o) Volatile matter Fixed carbon Moisture Ash 38.9 52.6 3.3 8.9 37.0 54.0 1.2 7.8 36.1 51.2 4.8 7.8 30.5 39.0 21.2 9.2 28.68 24.08 35.96 11.28 Colorado 73.1 5.1 1.16 1.1 9.7 9.8 12,400 Pittsburgh No. 8 77.2 5.2 1.19 2.6 5.9 7.9 13,700 Western Kentucky 73.0 5.0 1.4 3.1 9.3 8.2 12,450 Montana Powder River 67.2 4.4 1.1 0.9 14.0 11.7 8,900 Texas Lignite 60.44 4.61 1.21 1.75 14.38 17.61 6,677

209

Analysis test, because the ash metal constituents may be oxidized, sulfated or otherwise chemically transformed during the test. The relationship between the ash content and the mineral-matter content, on a dry basis depends on the form of the sulfur in the coal. F o r example, a typical correlation is W t - ~ ash = (Wt-~o mineral matter) - (wt-% pyrite) - 0 . 1 4 [(wt-Vo mineral matter) - (wt-~ pyrite)]. Ash consists primarily of silicon, but there are significant variations between different types of coals. For example, the lower-rank lignites tend to contain ash which is high in alkaline metals such as calcium, sodium, and magnesium. The presence of these alkaline metals leads to a low ash fusion-temperature, which is desirable for cyclone combustion but not so for other configurations, since it allows liquid metal particles to impact on convection-section heat-transfer surfaces and thus contribute to fouling. It should also be noted that the flyash constituents in coal may have a large influence on the speciation of particulate matter resulting from the combustion of that coal. F o r example, a high alkali content will allow a substantial portion of the sulfur in the coal to be chemically bound with the flyash, thus contributing to lower SO2 emissions, but to higher particulate sulfate emissions. Certain metals such as potassium, sodium, or nickel are likely to vaporize during the combustion process and subsequently condense to form a large number of very small submicron-sized particles. These particles act as nuclei for H,SO4 aerosol condensation in the exhaust plume and thus contribute to the resulting (small) aerosol size distribution. This, in turn, will lead to increased sulfuric acid plume opacity. Furthermore, gas-phase reaction between metals and sulfur cannot be ruled out, depending on the local temperature and environment. Future work will pay increased attention to the speciation of small particulate matter and to

problems of how this is affected by coal composition: ash composition and combustion characteristics. Proximate analysis is obtained as the result of an ASTM test in which first the moisture is driven off, then the weight loss at a controlled temperature (950C) is measured. This weight loss is classified as volatile matter. Fixed carbon is then determined through measurements of weight loss due to oxidation of the carbonaceous residue. The remaining mass is classified as ash. Volatile matter contains hydrocarbons (as well as other trace nitrogeneous and sulfur species), and so the fixed carbon is not equal to the elemental carbon analysis. As will be discussed later, the volatile matter content of coal is n o t equal to the amount of material which is evolved during the combustion of a particular coal, since the latter depends on the temperature of the particle during devolatilization. However, it is a standard measure of the ability of the coal to drive-off gases and therefore of the combustion characteristics in the primary fuel jet. It correlates the ability of various coals to support injector stabilized flames. Tradition has it that primary air used for transporting the coal through the injector should be sufficient to burn the Proximate ASTM Volatile Matter stoichiometrically. Coal can be viewed as being a physical mixture of several "macerar' components. Thin slices of coal can transmit and reflect light. Variations in reflective and light-transmission capabilities of these thin slices can be correlated with their maceral constitution, which, in turn, controls physical behavior of coal particles upon heating. F r o m a combustion point of view this means that in any pulverized coal mixture some particles have different properties and compositions from other particles. This adds substantially to the complexity of developing predictive techniques for coal-particle behavior. Variations occurring between different particles of the same coal may have significance in determining the behavior of different sized particles in a combustion process, and this may exert an undue

210

J.O.L. WENOT Physical processes may be quite important in determining how rapidly a particle undergoes devolatilization, what species are evolved, and what the fate of the ash is. Recent work 14'26 has shown that for particles as large as 500/zm the rate at which gases leave the surface of the particle is controlled by a physical resistance to passage of the volatiles through the particle. For nonmelting coals the physical resistance to fluid motion within the particle can be described by a form of Darcy's Law. It was shown that under certain circumstances, where outside pores are slowly being closed, pressure inside the particle can build up to very high levels, possibly causing particle rupture. Particle rupture can be of importance in determining the effect of a rapid heat-flux to the particle on the fate of both nitrogeneous and sulfur species in the coal, and can cause increased contact or mixing with oxygen. 14 Mathematical models describing these processes can be useful to estimate the probability of events such as these occurring. There is evidence 27 that a wide variety of types of char characteristics can be generated from pulverized coal combustion depending on the type of physical process undergone. These types range from broken spheres to porous balls. However, causal relationships between the type of char characteristics and the effect of particle heating-rate, temperature distribution, and environments, have not yet been determined. The review by Anthony and Howard 7 discusses physical factors at some length.
4. R E G I M E S OF COMBUSTION

influence on combustion characteristics and pollutant emissions. 3.2. Physical Behavior When a coal particle undergoes rapid heating, gases evolve from the particle surface. This may occur through several physical mechanisms. Some coal particles will first appear to become soft and almost liquid in form as gases or volatiles produced within the particle erupt from the surface through bublales which are passing through the particles. Figure 8a shows an artist's rendition (taken from high-speed cinemicrographs of an individual coal particle undergoing rapid heating) of such a coal particle, and the ensuing eruptions of bubbles at the surface. Such behavior is most common in the Eastern Bituminous and Appalachian coals. The process of bubble formation and passage through the liquid particle may control the rate at which gases evolve, and also determine the species evolved. 7 Other particles of the same coal or the majority of particles of other coals may not become liquid during heating, but rather crack and allow pores or fissures to open up, allowing volatiles to escape through them. These particles, somewhat more typical of the Western low-rank coals, do not change their size appreciably during devolatilization, even though they lose mass. Indeed, they form a porous residue. Devolatilization of a nonmelting coal is shown in Fig. 8b, in which the visible fissures on the particle surface should be compared to the bubbles erupting in the upper figure. The tendency of a coal particle to follow behavior depicted in Fig. 8a or 8b is determined by the particles' maceral constitution.

4.1. Particle Clouds Just as is the case for fuel-oil combustion, 2a it is unlikely that pulverized coal burns as an assemblage of individual particles where each one is an entity surrounded by its own specific reaction zone. More likely is that interparticle interactions are of great importance in determining how the coal particle burns. Clearly, radiative transport from some particles to those approaching upstream can be of great importance in the early volatilization sequence. Furthermore, it is surely possible that species evolved from one particle may react with those evolved from another particle prior to oxidation. Of special interest is the possibility that volatile nitrogen from one particle can react with nitrogen in some other fuel particle thus forming molecular nitrogen rather than NO. Likewise, NO formed by oxidation of volatile nitrogen from one particle can be reduced by reaction with volatile species arising from a subsequent particle. The overall global kinetic aspects of the combustion of particle clouds can be investigated through the use of a plane-flame furnace, which simulates a plug-flow reactor. This has been used with success, in the past, to provide insight into coal particle ignition and devolatilization 12 under combustion conditions. Clearly, kinetic data obtained in this way already contains in it some, but not all, factors due to interparticle interac-

(..- ,

Melting Coal: Volotiles escape through bubbles and shape of particle is no/ maintained. Some swelling occurs.

B)

.Tj

Non-melting Coal: Volofiles escape through fissures and shape of particle is maintained. No appreciable swelling or deformation occurs.

FIG. 8. Coal particle behavior under rapid-heating inert pyrolysis conditions.

Fundamental coal combustion mechanisms and pollutant formation in furnaces tions. Whether these interactions are the same as those likely to occur in an actual furnace flame is debatable, although mechanistic information derived from such experiments is certainly required as a first step towards their understanding. 4.2. Single Particles Some useful insight can be obtained by considering combustion behavior of single particles. When a particle is first heated, the first volatiles to come off, do so rapidly. 2 These volatiles play a role in flame stability characteristics and are probably oxidized far from the particle surface, in the first regime of combustion, shown in Fig. 9. Burning occurs in the premixed mode. Available data indicates that these volatiles are oxidized by the primary (transporting) air, although this depends on the coal spreader-angle, on particle size, and on particle heating-rates.
f--,

211

Totally Detached Flame. Rapid Mixing, Rapid Volatilization Fuel- Rich Regions Regions for Pyrolysis, Crocking HCN, N, Pyrolysis Compounds Formed Flame Front / ~ ~
"

Evidence will be presented in Section 6 of this paper that indicates that this second regime of combustion is a salient feature insofar as fuel-N conversion is concerned. Available models of single droplet combustion 29 might be adapted to simulate phenomena of interest here, and indeed might be more applicable to coal than to oil. It should be pointed out that it is during this second regime of combustion, that physical behavior of coal particles becomes important, since it is here that physical processes may control the rate of devolatilization, and thus the position of the flame front. After some time, all volatilizable matter has been driven off and oxygen reacts heterogeneously with the char surface. This allows a definition of the onset of the third regime of combustion---char burnout. In this regime, solid is directly consumed by oxygen, although primary gaseous products may react further either within particle pores or in the particle neighborhood. In practice, coal particles are not spherical, and so portions of the surface might still be blanketed by volatiles, while other portions allow penetration of oxygen directly to it. Because of this, and because a wide distribution of particle sizes and types exist, it should be recognized that the three regimes of combustion shown in Fig. 9 do not necessarily proceed sequentially. To model the fundamentals of the transition from the second to the third regime of combustion requires, therefore, some sophistication. Further discussion of the char burnout regime of combustion is found in Section 7.
5. DEVOLATILIZATION

B Attachedt Diffusio~ Flame. PyrolysisOccurs in Fuel-Rich Recjion Char is Being Formed

5.1. Introduction The thermal decomposition behavior of coal is important from a practical point of view because it may control: (1) the ignition process of the coal particle ; (2) the rate at which combustion proceeds ; (3) the rate at which oxygen is consumed ; (4) the rate and form of nitrogeneous, sulfur and other trace species evolution; and (5) the overall mechanisms governing the fate of fuel-N and sulfur. It matters greatly whether the nitrogen in coal is devolatilized and subsequently oxidized through homogeneous gas-phase reactions, or if the nitrogen (or other trace species) remains in the carbonaceous residue. The fate of volatile nitrogen will be determined, in part, by the local temperature and oxygen environment in which it is found immediately upon leaving the particle surface. The latter is determined not only by the overall mass evolution-rate, but also by the speciation of the mass evolved and by the manner in which evolution takes place. Thus, the fundamental devolatilization behavior of coal has a large bearing on the type of combustion modification that is most effective for NO= control concurrent with flame stability. Furthermore, practical experience in burning a bituminous coal and a coal char in an injector-stabilized pulverized-fuel flame will easily convince the sceptic of the importance of devolatilization in controlling flame shape and flame attachment.

C Char Burn-Out Surface Combustion Occurs when Volotiles hove Been Driven Out CO Formed as Result of Surface Reaction of Carbon and 0 2 This CO May Burn Externally to the Solid Particle in o Diffusion Flame, or Combustion Occurs Within the Particle (Internal Burning)

FIG. 9. Regimes of combustion that influence coal combustion behavior and fate of trace species. As volatilization proceeds, the rate decreases, and there must come a point when the rate of devolatilization is exactly balanced by the rate at which oxygen (if available) can consume the hydrocarbons in an attached diffusion flame. Under these conditions particle behavior is of interest, because if trace species such as fuel-N are evolved late, they must pass through a hot fuel-rich zone prior to meeting oxygen, even though the overall mixture may be fuel-lean. The "fuel" being oxidized in the flame front is neither that being formed from inside the parent coal nor that leaving the particle surface. Pyrolysis in the fuel-rich zone will vastly change the nature of the nitrogeneous and hydrocarbon species subsequently to undergo oxidation.

212

J.O.L. WENDT
-g 6O

The critical questions to be addressed can be listed : (1) What fraction of the mass of a coal particle is devolatilized? (2) What fraction of coal nitrogen is devolatilized? (3) What are the major species evolved? (4) What are the nitrogeneous species evolved? (5) What are the absolute and relative rates of hydrocarbon, CO and nitrogen evolution? (6) What is the minimum a priori information about a specific coal required to predict answers to the five questions listed above? Answers to the first and second questions will yield information on the relative importance of homogeneous versus heterogeneous combustion mechanisms, on the local stoichiometry/temperature history and on temperature and coal composition effects on fuel-NO. Answers to the third and fourth questions will provide further information on the above and, in addition, will determine the importance of subsequent secondary pyrolysis, both to form soot, and in the chemical processing of fuel-N to form N2. The fifth question is of critical importance in determining the rate at which oxygen can be consumed prior to nitrogen evolution, since this establishes the early NO formation, both in turbulent diffusion flames, and in fuel-rich or stagedcombustion configurations. The last question clearly relates to the engineering usefulness and accuracy of models that describe these phenomena, and that can be utilized to predict the influence of coal composition, temperature, and local environment on fuel-NO emissions. 5.2. Mass Devolatilized Although the ASTM proximate volatile matter of coals provides some basis for comparison of various coals, it is not equal to the amount of volatile matter that can be lost during combustion. That quantity is a function of pyrolysis temperature, and probably of particle heating-rate (although the latter point is somewhat in dispute). As the pyrolysis temperature increases, so does the weight lost during devolatilization, where the functionality depends on coal composition. The upper portion of Fig. 10 shows the asymptotic weight loss during devolatilization of both a Pittsburgh Coal and a coal char (the latter having been produced through the FMC-COED coal-conversion process), as a function of temperature. Note how the mass devolatilized increases as the pyrolysis temperature increases, within temperature ranges of practical combustion interest. These data are important because they delineate the minimum temperatures to which a coal or char particle must be heated before one might expect a homogeneous (volatile) ignition mechanism to occur. This temperature can be quite low for the coal, but much higher for the char. Therefore the particle heat-up time to volatile ignition will be longer for char than for the coal, thus causing char flames to become more easily detached from the injector. This

3
4o

20
I00 ~ - - - - - A - r - - - - - - ~ I I 1 I

~ 88o P ,, b ~ , g ,
~6o
~40

Q0 ~ '
A/

20

OJ/
[ ,4~

O/

/-,--- FMC Char

800

I000

1200

1400

1600 ~'K

1800

2000

Pyrolysis Temperoture

FIG. i0. Devolatilization yields for coal and a coal char. phenomenon has been observed in practice, a It should be kept in mind, however, that heterogeneous ignition mechanisms may also play a role, although the experimental data shows that, if so, heterogeneous ignition for char occurs more slowly than that for coal. Howard and Essenhigh 12 claimed that ignition occurred prior to significant devolatilization, and that at that point the gaseous mixture (if, indeed, it could be considered premixed) was too lean to support ignition. One might still argue that ignition of a microdiffusion-flame fueled by volatiles was not impossible under these conditions. Therefore, prevailing views on the relative importance of homogeneous versus heterogeneous ignition mechanisms have not yet reached a consensus) 7'a It is likely that many factors involving local heat transfer and other physical phenomena in addition to devolatilization also play a role in determining the ignition zone and position of the flame front, and this serves merely to add to the prevailing confusion and apparent discrepancies in experimental findings. In this regard, the following quotation from Anthony & Howard T is particularly apposite: "consideration of the results from different equipment requires a good understanding or model of the controlling phenomena. Even to scale up geometrically similar reactors without a clear appreciation of the relative influences of heat and mass transfer and chemical kinetics is risky". Some other more general comments are also appropriate. Different coals show markedly different functional relationships between asymptotic mass lost and pyrolysis temperature. 26 Furthermore, it is erroneous to assume that changes in proximate analysis are equal to mass of volatiles lost, although it is claimed 3 ~ that the two are linearly related.

Fundamental coal combustion mechanisms and pollutant formation in furnaces


10

213

80

70

60 oO o9

S
Z

50

4c
& LIG # 9
50

[] SUBBIT #14 0 BIT # 4 0 BIT # 2

~o~m/

///

,o~
n,
IO00

1200

~400

*600

t800

PYROLYSIS TEMPERATURE (K)

FIG. 11. Influence of coal composition and pyrolysis temperature on nitrogen yields.

5.3. Nitrogen Devolatilized


Mechanisms governing the fate of coal-N during combustion will be gas-phase for volatile nitrogen and heterogeneous surface for char nitrogen. Pershing and Wendt a2 showed that in a typical (Type II) pulverized coal flame: (1) approximately 50% of the coal-N was devolatilized; (2) conversion of volatile nitrogen to nitrogen oxides was high, but could be controlled through near-field burner aerodynamic changes; and that (3) conversion of char-N to nitrogen oxides was low but could not be controlled using near-field aerodynamics. Therefore, whether and how the nitrogen is devolatilized or remains in the char is of prime importance. The lower portion of Fig. 10 shows the weight-% N lost from a Pittsburgh Coal and from a coal char. In each case essentially all the nitrogen can be devolatilized at pyrolysis temperatures in excess of 1800 K. This indicates that char-NO may not exist if pyrolysis temperatures are sufficiently high. At low temperatures the char releases very little nitrogen, while the coal releases a substantial fraction. It is important to note that as pyrolysis temperature increases the nitrogen content of the volatiles must also increase, since the slope of the nitrogen loss-curve is much greater than that of the total-mass loss-curve. Since conversion of nitrogeneous compounds to N O decreases with increasing nitrogen content in gaseous fuels, this may explain why fuel-NO from pulverized coal flames has as weak a temperature dependence as is observed experimentally. 3'32 It should be noted, however, that this weak temperature dependence was found only for turbulent diffusion-mixed pulverized-

coal flames. Recent data 33 indicates that under overall "premixed" flat-flame conditions, fuel-NO from pulverized coal does have significant temperature dependence. The fundamental reasons for this apparent discrepancy still remain to be resolved, but it is of practical importance that they be resolved, since flame temperature, and/or heat loss is a variable that can be controlled (through, e.g. flue-gas recirculation). Further data, 34 shown on Fig. 11, indicate that the weight-% N loss can vary greatly with coal composition. A purely empirical correlation for twenty coals and two coal chars has the form 0.4861 x VM + 3545 exp L(T) =

( 0460
---~--]

}/_20460", ~ 1 + 3545 exp \ RT ]

(1)

where

L(T)
= mass fraction of nitrogen lost at temperature T VM = proximate volatile matter T = absolute temperature, K. The above correlation, shown by the shaded symbols in Fig- 11, is best for 800 K < T < 1800 K ; the r 2 fit for the 22 fuels is 0.919, and the correlation tends to overpredict nitrogen loss when T < 1000 K. Although the empirical model above is based on the asymptotic yield predicted from two competitive reactions, of which one yields a nondevolatilizable residue, it should be emphasized that this is probably not what really happens, since Fig. 10 clearly shows that char pro-

214

J. O. L. WENDT TABLE2. Comparison of sum of major gas fractions collected with fractionalcoalweightloss at 1000Cand 1400C pyrolysis temperatures (Ref. 26) Coal mass-fraction represented by gas collected at designated temperatures Species CO CH4 CO 2
C2H 6

Arkwright coal T = 1000C 1400C 0.030 0.035 0.003 0.011 0.015 0.000 0.094 0.428 0.220 0.057 0.057 0.008 0.002 0.028 0.008 0.225 0.459 0.490

Wyodak coal 1000C 1400C 0.107 0.027 0.094 0.009 0.007 0.000 0.244 0.540 0.452 0.202 0.036 0.067 0.001 0.023 0.004 0.333 0.608 0.548

Illinois No. 6 coal 1000C 1400C 0.060 0.024 0.026 0.005 0.005 0.000 0.120 0.461 0.260 0.123 0.020 0.024 0.001 0.013 0.002 0.183 0.579 0.316

CzH4 C2H 2 Sum of gas detected Wt.-fraction lost Sum/wt.-fraction lost

duced is definitely quite devolatilizable at high temperatures. This facet will be discussed in more detail in a following subsection. However, the fact that the volatile-N yield does depend on coal composition indicates that the NOx emissions from different coals of similar nitrogen contents will vary. 33

5.4. Major Species Evolved Data on major species evolved at temperatures below 1300 K, a at higher temperatures, 26 and under vacuum conditions, a~ point to the following conclusions: (1) Oxygen in coal is evolved as CO, CO2 and HzO, and this occurs at temperatures up to 1300 K ; (2) For low-rank coals containing appreciable oxygen, these are the main major-constituents evolved at lower pyrolysis temperatures. However, as temperatures increase, heavy molecular weight "tar" hydrocarbons dominate the composition of the additional volatiles; (3) For bituminous coals containing little oxygen, methane, other permanent hydrocarbon gases, and indeed the heavy "tars" themselves, dominate the composition of volatile major-species, with the "tars" the major fraction over the entire temperature range. This information is of practical importance for two reasons : first, the heating value of the early volatiles will have a direct bearing on the heat release-rate and subsequent coal particle heat-up rate in the early part of the combustion process ; and second, the capacity of the early volatiles to consume primary oxygen (or the oxygen available for ignition and initial oxidation) will influence how much oxygen remains, or the richness of the environment in which nitrogeneous species are evolved. The latter point is quite important, since the nitrogen is evolved later at the higher temperatures. According to the above arguments, a low-rank coal will both heat up more slowly and consume less oxygen than will a bituminous coal, in the early stages of combustion. Slower heating might possibly indicate that low-rank coals might produce less volatile-NO, and more char-NO, than high-rank coals burned under equivalent conditions.

One would expect that light permanent gases constitute the major portion of early volatiles, while the heavier components comprise the late volatiles, provided that no secondary pyrolysis has occurred. This is because light gases probably result from the breaking of weaker bonds. Table 2 shows a comparison of light species evolved as a function of pyrolysis temperature. Light gaseous species considered are CO, CH4, CO2, C2H6, C2H4 and C2H2. F o r each of the three coals the fraction of these early volatiles to the total volatile-loss increases with pyrolysis temperature. This is important, since it further demonstrates the importance of rapid heating for easy ignition of the primary fuel jet and for resulting flame stability. Table 2 also shows that the species evolved are a strong function of coal composition. The Wyodak Coal is a sub-bituminous coal with a high oxygen content, and its devolatilization results in high CO. This confirms that CO can be correlated with oxygen content and that the remaining oxygen appears as CO2 and H20. It is also interesting to note that acetylene is evolved only at the higher temperature, and this is significant, since acetylene is a known precursor to soot. Provided they are high, the heating rates themselves do not appear to influence significantly the species evolved for a given ultimate pyrolysis temperature. The late volatiles (40-80~o of the total) evolve essentially as heavy organics that can condense at temperatures in excess of 1000 K. These heavy ends may well break down through pyrolysis prior to oxidation, but in so doing will almost certainly form soot. Ubhayakar et al. 36 estimated that in a temperature range between 1900-2300 K, almost one third of the mass of volatiles was transformed to soot. The high luminosity associated with pulverized coal flames is in part attributable to the burnout of soot formed from volatiles rather than to char burnout. One would expect reactions involving late volatiles to be of primary importance in region (i) in Fig. 1. Very little is known about kinetic mechanisms involving pyrolysis and oxidation of heavy hydrocarbons typical of late volatiles.

Fundamental coal combustion mechanisms and pollutant formation in furnaces 5.5. Nitrogen Species Evolved Recent measurements 26 on the speciation of nitrogen evolved show that most of the nitrogen evolved does so with the tar. Data taken under atmospheric conditions 34 indicates that up to 9 0 ~ of all nitrogen evolved under high-temperature conditions consist of high-molecular compounds with boiling points in excess of 900K. However, other data, taken under vacuum pyrolysis conditions, shows appreciable quantities of HCN at temperatures in excess of 1500 K. 37 There is general agreement, however, that the first nitrogen to evolve, does so not in the form of permanent nitrogeneous gases, but in heavy "tar" species containing nitrogen. This implies that it belongs to the "late volatiles". Speciation of nitrogen compounds is important because it may influence the kinetics of fuel-N pyrolysis relative to oxidation. For example, one may pose the question: How long and at which temperature must a nitrogeneous species remain in the absence of oxygen in order to be pyrolyzed to Nz? Speciation of primary-N volatiles may also influence the subsequent speciation of secondary pyrolysis compounds. In a previous section it was noted that the high molecular weight "tar" will likely form soot, which subsequently burns out. It remains to be seen whether this soot contains fuel-N, and, if so, w h a t t h e resulting conversion of this nitrogen to NOx may be. 5.6. Relative Rates of Volatile and Volatile-N Evolution Time-resolved weight-loss and nitrogen weight-loss measurements are difficult to perform precisely, because of difficulty in quenching the process in millisecond intervals. Therefore, there is a paucity of absolute-rate measurements. However, certain conclusions may be inferred from the data on asymptotic yields presented in previous sections. One might expect that light permanent gases observed at low temperatures come off early (comprising early volatiles), while the heavy high-boiling-point fractions (tars) evolve late in the devolatilization sequence (late volatiles). Therefore, based on the product yield data, one can infer that nitrogeneous species are evolved after the early volatiles (CO, COz, H20, CH4, other hydrocarbons) have come off and consumed their portion of primary oxygen. This hypothesis has been confirmed by Pohl and Sarofim, 3s who showed that coal lost nitrogen only after 15~o or more of the mass was lost. Experiments on a swirling turbulent diffusion flame, ~4 in which the primary oxygen was replaced by an inert CO2, showed that this did not influence NO greatly. This implies, again, that fuel-N is not evolved in time to react with primary oxygen. Clearly, for wall-fired configurations, control of local environment in region (1) (Fig. 1), and possibly region (2), will determine the ultimate fate of volatile coal-N. How these regions can be maintained oxygen deficient is shown on Fig. 4.

215

Research effort is presently underway to quantify how nitrogen and sulfur evolution-rates compare to hydrocarbon devolatilization rates, and how they depend on coal type, and temperature history. It is thought that a fundamental key to controlling fuel-NO emissions from pulverized coal is this competition between hydrocarbons and volatile-N for oxygen, where this competition is controlled by the relative rates of appearance of each species. Additional effort is focussing on the relationship between devolatilization rates and coal constitution and chemistry. These efforts are required in order to develop a predictive model which correlates the relative combustion behavior of various coals. The latter will aid the combustion engineer in predicting the behavior of his furnace as he switches from one solid fuel to another. 5.7. Devolatilization Models The foregoing demonstrates the need to develop mathematical models that predict and generalize the hydrocarbon and nitrogen devolatilization behavior of a wide range of coals. Soloman and Colket 35 present a strong argument that the species evolved, and indeed the rates of evolution, can be predicted from detailed knowledge of the chemistry (functional groups) of coal constitution. With this knowledge, some very simple and elegant generalizations can be made. However, the basic information on coal constitution required is less easy to obtain and so there is room for even more empirical generalizations that require only the conventional ultimate and proximate analysis. Other workers (see Ref. 7) have formulated phenomenological kinetic models, that either neglect or globally average over reactions occurring inside the particle. Use of single Arrhenius rate expression 3~ may be useful for interpolation of data from small particles, but is questionable for extrapolation to difference size ranges or conditions. Therefore, rates reported in the literature may be quite invalid for cyclone-furnace combustion conditions. It has been shown, v however, that a set of independent first-order reactions can be successfully approximated by a single first-order reaction having lower frequency-factor and activation energy than any of the multiple reactions. Kobayashi et al. 39 claim that at least two parallel reactions are essential to predict the dependence of ultimate yield on temperature, and moreover, were able to fit their resolved data for two coals (see Fig. 12) with the following reaction model :
1/7

volatile 1 + residue 1 xl 1 - xa (2)

~ k / Coal 2 .k 2

"-~ volatile 2 + residue 2


X2 1 -- X 2

Some features of this two component hypothesis (or some variation thereof) are of use in relating devolatilization parameters to pulverized coat behavior

216
801 70[-~ 60 ~'[ -r +~
,oH
I I ,

J.O.L. WENDT energies,f (E), such thatf(E) dE represents the fraction of nitrogen released through reactions with activation energies lying between E and E + dE. If I(E) = ~ 1 exp [ ( E ~ j o ) 2] --(3)

LIGNITE
-

/ ,*---'w---"--

o 2100 K * 1940 0 1740


o ,ooo
~

J
0

then the nitrogen loss as a function of time is given by


,5o
TIME (ms)

%
80,/

50
,

,oo
~

%0

RESIDENCE 7nL ,. v ./,.-,~., ~ 60 P:Io


5 Hd'%

ctN = 1 - a,//~------~ exp - k o exp ( - E / R T ) dt 1 f;oo E f


- ~ dE. (4)

r/k'.-llll
^

BITUMINOUS COAL ~ oelOOK + 1940


+ ,,4o
^

ol51o

~1260

,or

t,

+
,W'o ,,

o,ooo

sol-Ill

50 IO0 150 RESIDENCE TIME (ms)

200

FIG. 12. Time-resolved devolatilization profiles as functions of pyrolysis temperature (Ref. 39). in furnaces. The rate coefficient kl has an activation energy of 25kcal/mole compared to that of k2, 40 kcal/mole. We can consider volatile 1 to be early volatiles; i.e. those that come off rapidly, promote ignition, and react with the primary oxygen, probably in a premixed mode (regime 1). The other component, late volatiles, arise slowly, probably after the particle is in contact with secondary air (or in region (1), see Fig. 1, for the wall-fired case). These late volatiles may then burn in a diffusion flame in the neighborhood of each particle (regime 2). Early volatiles are desirable for rapid ignition and flame stability. Combustion of partially devolatilized coal or coal char is difficult because of a lack of early volatiles. Changes in either coal composition or heating rate will cause changes in the relative quantities of early and late volatiles, and accompanying changes in flame stability characteristics. Increasing temperature will increase the quantity of early volatiles, and therefore enhance flame stability. Other features of the scheme shown in eq. (2) are in direct conflict with experimental facts. For example, the simple competitive reaction model ofeq. (I) implies that the residues formed are essentially inert, i.e. they do no longer devolatilize when heated to a higher temperature. The top half of Fig. 10 shows that this concept is not really valid as far as the FMC char is concerned, since there the residue does lose appreciable mass when heated. Examination of the lower portion of Fig. 10 indicates that as far as nitrogen evolution is concerned, the residue-N can certainly not be considered an inert product. Rather, it behaves just like the parent coal, with its low-temperature-N absent. Another conceptual approach, therefore, is to consider nitrogen in coal as being bound with a distribution of bond energies. This distribution may be linked to the temperature-dependence of asymptotic nitrogen yields, and implies a distribution of activation

Clearly, the latter cannot be valid as t ---, oo, at which point ct, -. 1, but it may be valid for short times typical in a combustor. This model does not view char-N as an inert product, but as a partially reacted reactant. Ultimately, some combination of the models described in eqs. (1) and (4) may be closer to reality for 0 < t < oo for both nitrogen-loss and weight-loss predictions. The latter requires, in addition, rate models for CO, CO2 and H 2 0 evolution. However, the distributed activation energy approach may have an engineering usefulness if the three parameters required can be predicted for both nitrogen and hydrocarbon evolution. It would then be possible to model the relative devolatilization rates as functions of temperature, and this could be utilized in predicting combustor performance. Clearly, there are altogether many loose ends still to be tied up and much further work required to be done. The purpose here is merely to provide a very limited overview on possible useful approaches available, and the serious deficiencies of each.

6. F A T E O F V O L A T I L E

SPECIES

Having considered how various species are devolatilized, the next step is to investigate how they react in given environments. 6.1. Premixed Mode Although premixed combustion of volatiles in the micro-mixing sense occurs only rarely in practical pulverized coal combustors, study of kinetics in premixed environments can provide some useful insights. Investigations of premixed homogeneous kinetic mechanisms help determine kinetic limitations to the fate of fuel-N and of other trace species. Results from these type of investigations have direct application in the fate of early volatiles, since they may well burn in a premixed mode far from the particle.'Premixed kinetics may also be controlling in cyclone combustors where local shear-rates may be sufficiently strong to prevent attached diffusion flames around particles in the disperse phase. It is appropriate to cite examples where research of premixed systems has provided insight applicable to pulverized coal. Several experimental studies =3"*'4t have shown that under premixed conditions the oxidation of fuel-N is a strong

Fundamental coal combustion mechanisms and pollutant formation in furnaces function of local stoichiometric ratio. Very low conversions have been measured at substoichiometric conditions, while under fuel-lean conditions conversions range from 70 to 98~o, depending on temperature and fuel-N species and concentration. The former point reinforces the value of staged combustion as a means to lower pollutant emissions. The latter point indicates that N O , emissions from cyclone combustors, where mixing is rapid, are high. The fact that close to quantitative conversion occurs in premixed fuel-lean systems can be used to obtain insight into whether nitrogeneous species devolatilizing from coal particles burn in a premixed or diffusion flame mode (regime 1 or 2).

217

3400 /

Weslern Kentucky Cool, SR = I 2, "Prem~xed"

100% Convers,on of Totol Cool Nitrogen

2800 l

100% Conversion of Volatile Nitrogen 0 2OOO Z c,

120C

/
1 02

~"D

/:

~ctuol

the comparison that indicated that, at best, only the early volatiles containing little volatile-N burned in the premixed mode. Another example of insight obtained through a premixed kinetic limit case study is that showing that, under fuel-rich conditions, fueI-S can have a first-order influence on fueI-N.42 Those data indicated that the conversion of volatile fueI-N to NO under fuel-rich conditions was higher in the presence of sulfur than in its absence. This has important implications on the role of coal composition and sulfur content on NO x emissions. Not only is it important to determine relative volatilization rates of hydrocarbons and nitrogeneous species, but also relative rates for nitrogeneous and sulfurous species. Should inorganic or pyritic sulfur be evolved sufficiently early to be available to interact with nitrogeneous volatiles, then it would appear that inorganic sulfur removal would have a beneficial side effect on NOx. NO can be reduced homogeneously ~3 by hydrocarbons 44 and by other nitrogeneous species. 45 This has great implication in pulverized coal combustion, since it indicates that volatiles released downstream from some particles can reduce NO formed upstream from other particles. This leads to new possibilities in arranging fuel and air streams in such a way that this phenomenon can be exploited. It is appropriate to point out that although premixed flat flames have received much attention, very little kinetic data is available on pyrolysis and oxidation of heavy hydrocarbons such as are evolved as late volatiles. Much more work requires to be done in soot formation from these tars. However, the oxidation kinetics of early volatiles (light hydrocarbons) are readily available. It is outside the scope of this paper to provide a detailed review of homogeneous gas-phase reaction mechanisms. For the latter, the reader is referred to Ref. I0. 6.2. Diffusion Mode

40C

I I 04 06 TIME, s

I 08

I 10

FIG. 13. NO profile from "premixed", fuel-lean combustion of pulverized coal in a one-dimensional reactor. Figure 13 shows time-resolved profiles of N O from a "premixed" but heterogeneous mixture of pulverized coal and air, at an overall fuel-lean stoichiometric ratio of 1.2. The yield is compared to the yield to be expected from 100% conversion of volatile-N, where the extent of nitrogen volatilization was 7 0 ~ (for a measured temperature of 1430K, see Fig. 11). Clearly, the N O formed is less than what would be expected if volatileN burned in a premixed mode. This indicates that in "premixed" heterogeneous coal-air systems fuel-N oxidation kinetics occur in a diffusion flame environment, probably around individual particles or in segregated blobs. In this case, premixed data provided

In a diffusion flame, kinetic phenomena exist that may have no analogy in a premixed flame environment. For example, in a diffusion flame, regions exist in which pyrolysis reactions occur in the absence of oxygen, and these environments are difficult to replicate with a self-sustaining premixed flame. In the sections above, evidence was presented that it is likely that late volatiles containing nitrogen react in a diffusion-flame mode, even when the particles are "premixed" with air. Clearly, the degree to which this phenomenon is important depends on the relative velocity between particle and surrounding fluid, i.e. on the inertia of the particle. Therefore, the mode of combustion of late volatiles should be influenced by primary coal jet velocity and the aerodynamics of the surrounding flow-field. It is known that conversion of volatile fuel-N depends strongly on the mode of combustion. 46 Conversions to N O in the premixed mode are always higher than conversions in the diffusion mode. Some

218

J.O.L. WENDT
7. C H A R B U R N O U T A N D FATE O F TRACE SPECIES

analysis is available to describe NO~ formation around evaporating droplets, 29 but analysis of interacting diffusion-flame regions is sorely lacking. As far as fuel-N conversion is concerned, a key parameter is the time a fuel-N species exists in a hot fuel-rich environment, since this affects the extent of pyrolysis prior to oxidation. Analysis of kinetic mechanisms in a diffusion-flame environment is sorely needed, and more effort in this direction is required.

7.1. Global R a t e s Field 5 has reviewed data relating to the combustion of pulverized coal char and found that most results can be correlated by q = 8710 exp ( - 3 5 7 0 0 / R T ) P o 2 (6)

F~ame S h e e ~

A. Flame Attached to Part~~

Locationof FlameSheet of FiniteThickness Bulk Fuel ! Bulk


Oxidizer

Fl ame Front Initially After Time~t C. TurbulentDiffusionFlame


with Stretch

Zonewhere DiffusiveFluxes are Signi flcont B Microscopic Viewof RegionSpanning the FlameFront

FIG. 14. Models of diffusion flame influences on volatile nitrogen oxidation (Ref. 47).

One approach to this problem is to consider a limit case of a flame-sheet formed in the second regime of combustion, but distorted and stretched by local turbulence. This is shown schematically in Fig. 14, where the local flame thickness is controlled by the rate of molecular diffusion to the reaction zone. The flamesheet must have a finite thickness so that competitive reactions involving hydrocarbon oxidation and fuel-N pyrolysis and oxidation are considered. Sternling and Wendt used a simple two-film model 47 to simulate the reaction-zone thickness, and they indicated that increasing local reaction combustion intensity, Ct, given by
C t = a/(102) 4/3 = bFofflR

where q = char consumed per unit time per unit external surface area, g/cm 2 sec, T = temperature in K, Po2 = partial pressure of oxygen in atm. Hamor et al. 49 correlate their data for 630K < T < 2200 K on pulverized brown coal char by q = 9.3 exp [ - - 1 6 2 0 0 / R T ] P ~ . (7)

Note the difference in activation energies by a factor of two, which may be explained through internal burning effects described in Section 7.2. The latter authors note that : (1) above 900 K this rate was subject to rate control by the coupled processes of pure diffusion and chemical reaction of pore walls ; (2) at 1800 K eq. (7) is a factor of four greater than that for anthracite, and 50% higher than that for bituminous coal ; (3) at 770K the reactivity of the char is one-six orders of magnitude higher than the reactivities of other carbons, a fact borne out by comparison of eqs (6) and (7). Since the global char burnout-rate occurs primarily in region 2 (see Fig. 1) and since it actually determines

(5)

where 1o2 is the diffusion length of oxygen film, l R is reaction-zone thickness, a, b are some constants, Fo2 is flux of oxygen to reaction zone, led to increasing fuel-N conversion, a trend which is observed in practice. This approach may be extended to include the influence of local aerodynamics through flame stretching. 48

Fundamental coal combustion mechanisms and pollutant formation in furnaces the overall size of the radiant zone in a boiler, it is surprising that it has not been more precisely determined. It has, however, been conclusively shown that: (1) the process is not controlled by diffusion of oxygen to the surface (Nusselt hypothesis) ;50 (2) combustion does not proceed according to a shrinking core model; and (3) internal burning, pore diffusion and reaction are important. 5x The latter conclusion bears further pursuit, because it indicates a strong coupling between the devolatilization process forming pores and the subsequent burnout of the carbonaceous residue. One might expect very rapid devolatilization to form large pores and slow devolatilization to form a fine pore structure. Comprehensive surveys of detailed mechanisms of char burnout are available in this journal 8 and in the review by Mulcahy and Smith. 9 Here we shall focus on some special aspects of the problem and their significance. 7.2. Internal Burning Internal burning denotes a physicochemical process in which oxygen or other reactants diffuse into pores, there to react with the solid surface. Products formed may react further either with the surface or homogeneously in the pore volume. The phenomenological process has long been studied by researchers in catalytic reaction engineering. The important general conclusions to be drawn are as follows. The external surface area plays a very small role in the process since the internal area exceeds that by many orders of magnitude. Therefore, the external diameter of the char particle may not necessarily change in a well-defined manner as combustion proceeds. Particle size does change, but this process occurs through internal breakage and other physical processes, to be described in the following section. Pore diffusion and its effect on overall reaction rates can be described through use of the effectiveness factor, e, defined as follows :
Rc = ekAsPo~ (for first-order reaction),

219

from 1 (reaction controlled} to 0 (diffusion controlled). For non-isothermal reactions ,: can achieve values greater than 1 because of coupling between heat transfer and its effect on Arrhenius reaction-rate expressions. The reaction engineering literature is replete with analytical, asymptotic and numerical predictions for e, and the reader should consult the appropriate texts by Szekely et al., 53 Petersen, s4 Levenspiel $5 and Carberry. 56 For consecutive reactions such as :
C q- 1 0 2 --# C O C O -~- 1~0 2 ~ C O 2

the yield of an intermediate, such as CO in the above scheme, is influenced by pore diffusion. Pores of small radii will decrease the yield of intermediate product, for specified reactant concentrations at the pore mouth. The physical explanation of this result is that the ratio of the concentration of intermediate product to that of reactant within the pore is always greater than that at the pore mouth, because a positive concentration gradient of the intermediate is necessary to transport it from the pore. Hence, as pore radius decreases the ratio of the rate at which an intermediate will be consumed in the pore to that at which it is formed will increase (assuming each rate increases with concentration of reactant), leading to a lower selectivity for the intermediate. 54 The nature of the primary reaction product from the c h a r - 0 2 reaction has received substantial attention. Prevailing opinion is that oxygen forms CO at the surface s7 (although even this step involves the other more fundamental steps of adsorption, desorption, etc). The influence of pore characteristics on selectivity of CO should affect the fraction of carbon leaving the particle as CO. Moreover, it is not clear whether any CO leaving the particle is consumed in a diffusionzone surrounding the particle or in regions far from the particle. This depends on the local scale of turbulence and particle size.

k = intrinsic rate coefficient, As = true internal surface area, Po: = partial pressure of oxygen outside the particle. At high temperatures, when penetration of oxygen into the particle pores is slight, the effectiveness factor, e, is proportional to the inverse of the square root of the rate coefficient, k, thus
e oC llk 112

7.3. Fly-Ash Formation and Char Burnout Pulverized coal particles are of the order of 60 gm in size. These particles may swell or stay the same size during devolatilization. Yet the ash collected from pulverized coat combustion is of the order of 0.5-5/~m. Clearly, particle size does change drastically during char burnout. Following Flagan and Friedlander, 58 the following phenomenological process can be described. This process is consistent with the views on the importance of physical processes during devolatilization and on internal burning outlined above. Ash occurs in the form of crystallites of mineral matter in coal, and during char burnout some of these crystallites are separated from each other. However, as the particle temperature increases the ash may become molten, forming globules connected together by a char matrix. As the char is consumed, the particle breaks up,

Therefore, under these conditions the effective activation energy of reaction is one half of the true intrinsic activation energy. The discrepancy between the activation energies in eqs (6) and (7) may be accounted for by this fact. The best estimate for the intrinsic reactivity of porous chars has been obtained by Smith s2 and has an activation energy of 42876 cal/mol, which is not too far from Field's estimate. For isothermal, integral order reactions, ~ varies

220

J.O.L. WENDT
t,, TOTAL NO 12OO O FUEL NO (AR/O 2 /CO2)

continues burning, and ultimately leaves behind flyash particles. Coals with a high alkali metal content, such as lignite, will form silicon metal oxide complexes with a low ash fusion-temperature. This means that the ash globules will remain liquid as they are transported to the convection section of a furnace. Should these liquid globules impact on heat-transfer surfaces, they will fuse with the metal and cause fouling. This process is not self-correcting, since as heat-transfer surfaces foul, gas temperature increase and the ash will stay molten longer, thus causing more fouling downstream. It is not known whether combustion modifications in controlling the environment in which char burns out and in which ash globules are formed, can play a role in controlling ash fusion-temperatures and/or ash particle size. The origin of submicron particles, other than soot, has only recently been investigated: s Some metal oxide species will vaporize, possibly react with SO2, and recondense to form trace submicron particulate emissions. Both particle size (0.1-0.7 #m) and particle speciation are of interest. Specific concern has centered on not only H2SO4 aerosol but also metal sulfates in the atmosphere. Sulfates can also be formed on soot particles, s9 The formation of these trace species in a combustor may well be affected by the timetemperature history and environment suffered by coal particles as they are consumed. Moreover, it cannot be ruled out that combustion modification for, say NOx control, will not have significant influence (possibly adverse) on emission of these trace species. 7.4. NO Formation and Char Burnout Volatile-N compounds may be successfully hindered from being oxidized to NO, by aerodynamic changes promoted by a burner. This is because volatilization occurs rapidly, and the environment in which it occurs can be influenced by near-field changes in environment (regions 1 and 3 in Fig. 1). The fraction of nitrogen remaining in char can be high (Fig. 11), and its conversion to NO is not affected by near-field burner aerodynamic changes. 3 Experiments, in which a coal char containing negligible volatiles, was burned under overall excess air and "premixed" conditions gave conversions of char-N to NO of 40~/o,21 which is quite low. This indicates the importance of diffusion at or inside the particle, which is consistent with our ideas on internal burning. Wendt and Schulze 6 considered the problem ofdiffusion and reaction ofchar-N species inside a pore, and concluded the pore size characteristics could have a significant influence on the selectivity ofthe particle towards NO. This is because NO can be viewed as an intermediate product which can be both made and destroyed in the char. Again, there is a coupling between devolatilization processes and the subsequent char burnout as far as NO is concerned. That hot char can also reduce NO was shown 61 for fluidized-bed combustion. Subsequently, studies in

O CALCULATED-NO ADDITION O CALCULATED-NH3 ADDITION

Z ~

ICO0

c~ 8 0 0

5
0 600

VOLATILE NO

40O

200

~
I I
I I0

.J:z--.---co-u
CHAR NO i

0 I0

41

i
120 RATIO

I
1.30

STOICHIOMETRIC

FIG. 15. Sources of NOr emissions in turbulent-diffusion pulverized-coal flames. Wall-fired simulation: .Western Kentucky Coal (Ref. 32). which char was burned in a premixed but fuel-rich mode indicated 21 that char-NO is made and subsequently destroyed by char. The global reduction reaction was of fractional order in NO (indicating diffusion coupled with surface reaction) and with an activation energy of 47kcal/mole. That some other reducing species such as H2 or CO was also required is likely but not proven. Heterogeneous reduction of N O by char may prove to be a valuable NOx control technique and might be achieved through combustion modifications. Heterogeneous reduction of N O implies yet another coupling between devolatilization and char burnout, since volatile-NO formed early may well be subsequently reduced by char. Such coupling may be quite important in the development of combustion modifications to lower both volatile-NO and char-NO. Some insight into the relative quantities of volatileNO and char-NO in practical swirling flames was obtained through special experimentation in which volatile-N was deliberately added to a pulverized coal jet, and the ensuing increase in NO emissions attributed to the increase in volatile-N) 2 Figure 15 shows results for a wall-fired simulation, and indicates the breakdown in NO emissions to be expected. Note that, of fuel-NO, volatile-NO dominates under this unmodified combustion condition. However, volatileNO can be most easily controlled through burner aerodynamics controlling the local environment in which volatilization takes place. Therefore, char-NO, although small at present, may loom to be of greater and greater significance as abatement requirements tighten. Since the fraction of nitrogen remaining in char can be altered by pyrolysis temperature, and since volatile-N conversion is high but controllable, while that ofchar-N is low but uncontrollable, an interesting

Fundamental coal combustion mechanisms and pollutant formation in furnaces optimization problem can be posed : Is it desirable to have a very hot devolatilization zone and attempt to control the volatiles, or should one aim to put more nitrogen in the char with the resulting low conversion ? F u n d a m e n t a l research on rates of devolatilization, formation and destruction of N O can help provide the answer.
8. CONCLUSIONS

22l

Coal combustion in furnaces involves many complex aerodynamic and physicochemical phenomena. An understanding of these p h e n o m e n a can influence directly how burner hardware must be developed to allow for variations in coal composition, particle size and characteristics, without deleterious pollutant emissions. T h o u g h combustor configurations vary widely, certain fundamental mechanisms are pertinent to each one and therefore can aid in relating data from one configuration to those of another, Even in the absence of aerodynamic complications coal combustion can be described in terms of neither kinetics nor diffusion alone; both are important, especially with respect to the formation of N O x from coal-N. A proper understanding of species devolatilization can provide insight into coal combustion behavior for a wide variety of coals and combustors. H o m o g e n e o u s reactions involving formation and reduction of trace species are important, and further study is required to determine mechanisms in a diffusion influenced environment, especially in regions where oxygen is absent. Heterogeneous reactions involving pore p h e n o m e n a require significantly more experimental data for their elucidation, as do flyash formation mechanisms. Some simplification or unifying principle is highly desirable. Yet, experience has shown that oversimplified concepts such as those of simple diffusion control a n d / o r kinetic control are at vast variance with the data and cannot be used for extrapolation of data to untested conditions. The challenge to the researcher is to provide the designer with just sufficient concepts and/or theory that he/she can, with confidence, provide coal combustion equipment that is safe, effective and environmentally sound.
REFERENCES

1. Proc. Fuel Switching Forum, 6-7 June (1977) Pittsburgh, Pa. Sponsored by U.S. Department of Energy, Report CONF-77-658, available from National Technical Information Service (NTIS), U.S. Department of Commerce, 5285 Port Royal Road, Springfield, Va 22161. 2. ESSENH1GH, R. n., Combustion and flame propagation in coal-systems: a review, Sixteenth Syrup. (Int.) on Combustion, p. 353, The Combustion Institute, Pittsburgh, Pa (1977). 3. PERSHING,D. W. and WENDT, J. O. L., Pulverized coal combustion : The influence of flame temperature and coal composition on thermal and fuel NO~, Sixteenth Symp. (Int.) on Combustion, p. 389, The Combustion Institute, Pittsburgh, Pa (1977). 4. FARADAY,M. and LYELL,C., Phil. Ma~t. 26, 16 (1845). 5. FIELD, M. A., G1LL, D. W., MORGAN, B. B. and

HAWKSLEY, P. G. W., Combustion of pulverized coal, BCURA Leatherhead, Cherey and Sons Ltd., Surrey, England (1967). 6. GILL,D. W., Pulverized coal firing--A review of J. inst. F. Literature, Ener.qy World, p. 4 (1975). 7. ANTHONY, D. B. and H(IWARD,J. B., Coal devolatilization and hydrogasification (journal review), A IChE J., 22, 625 (1976). 8. LAURENDEAU, N. M., Heterogeneous kinetics of coal char gasification and combustion, Proq. Enerqy Comb. Sci. 4, 221 (1978). 9. MULCAHY, M. F. R. and SMITH, I. W., Kinetics of combustion of pulverized fuel; A review of theory and experiment, Rev. pure appl. Chem, 19 (1969). 10. MALTE,P. C. and REES,D. P., Mechanisms and kinetics of pollutant formation during reaction of pulverized coal, in Pulverized Coal Combustion and Gasification (Eds SMOOT, L. D. and PRATT,D. T.), Chapter 11, Plenum Press, New York (1979). 11. BEER, J. M. and CHIGn~R, N. A., Combustion Aerodynamics (Ed. BEER,J. M.), p. 125ff, Halstead Press Division, John Wiley, New York (1972). 12. HOWARD,J. B. and ESSENHIGH,R. H., Pyrolysis and coal particles in pulverized fuel flames, Ind. En.qnq Chem. Process Desiyn and Development, 6, 1 (1967). 13. DESOETE,G. G., Overall reaction rates of NO and N 2 formation from fuel nitrogen, Fifteenth Symp. (Int.) on Combustion, p. 1093, The Combustion Institute, Pittsburgh, Pa (1975). 14. WENDT,J. O. L. and PERSHING,D. W., Physical mechanisms governing the oxidation of volatile fuel nitrogen in pulverized coal flames, Comb. Sci. Tech. 16, 111 (1977). 15. BUETERS, K. A., COGOLI, J. G. and HABELT, W. W., Performance prediction of tangentially fired utility furnaces by computer model, Fifteenth Symp. (Int.) on Combustion, p. 1245, The Combustion Institute Pittsburgh, Pa (1975). 16. WALKER, P. L. JR., SHELEF, M. and ANDERSON,R. A., Catalysis of carbon gasification, Chemistry and Physics of Carbon, Vol. 4, p. 287, Marcel Dekker, New York (1968). 17. GERSI-IMAN, R., HEAP, M. P. and TYSON,T. J., Design and scale-up of low emission burners for industrial and utility boilers, Proc. Second Stationary Source Combustion Syrup,, Volume V, p. 65, EPA Publication EPA-6001777-073e (1977). 18. MICHELFELDER, S., Applications and field experience with low NOT burners, Paper presented at EPA Conf. on Coal Combustion Technology and Emission Control, California Institute of Technology, Pasadena, Ca, 5-7 February (1979). (Proceedings to be published). 19. LACrtAPELLE, D. G., Application of staged combustion of coal fired utility boilers, Paper presented at the EPRI NO~ Control Technology Seminar, San Francisco, Ca (1976). 20. BROWN, R. A., MASON, H. B. and NEUBAUER, P., Investigation of staging parameters for NO~ control in both wall fired and tangentially coal fired boilers, Proc. Second Stationary Source Combustion Syrup., Vol. lII, p. 141, EPA Publication EPA-600/7-77-073c (1977). 21. WENOT,J. O. L., PERSHING.D. W., LEE,J. W. and GLASS,J. W. Pulverized coal combustion! NOT formation mechanisms under fuel rich and staged combustion conditions, Seventeenth Symp. (Int.) on Combustion, p. 77, The Combustion Institute, Pittsburgh, Pa (1979). 22. JOHNSON, S. A. and CIOFF1, P. L., Development of an advanced combustion system to minimize NOT emissions from coal fired boilers, Presented at the 1978 Joint Power Generation Conf., Dallas, Texas, September (1978). (See also comment on p. 85, Seventeenth Symp. (Int.) on Combustion, Pittsburgh, Pa (1979)t. 23. DESOETE,G. and FAYARD,J., Reduction of nitric oxide by soot in combustion products, Paper presented at the Fifth Members Conf. of the I.F.R.F., 8-10 May, Leeuwennorst Congress Center, Th~ ? ~ herlands ( 19781.

222

J.O.L. WENDT cyanide from atmospheric and fuel nitrogen in rich atmospheric pressure flames, Comb. Flame, 27, 187 (1976). HAYNES,B. S., The oxidation of hydrogen cyanide in fuel rich flames, Comb. Flame, 28, 81 (1977). WENOT, J. O. L., MORCOMB, J. T. and CORLEY, T. L., Influence of fuel sulfur on fuel nitrogen oxidation mechanisms, 17th Symp. (Int.) on Combustion, p. 671, The Combustion Institute, Pittsburgh, Pa (1979). WENDT,J. O. L., STERNL1NG,C. V. and MATOVICH,M. A., Reduction of sulfur trioxide and nitrogen oxides by secondary fuel injection, Fourteenth Symp. (Int.) on Combustion, p. 881, The Combustion Institute, Pittsburgh, Pa (1973). MYERSON,A. L., The reduction of nitric oxide in simulated combustion effluents by hydrocarbon--oxygen mixtures, Fifteenth Symp. (Int.) on Combustion, The Combustion Institute, Pittsburgh, Pa (1975). LYON,R. K. and LONGWELL,J. P., Selective, non-catalytic reduction of NOx by NH3, Proc. of the NOx Control Technology Seminar, February EPRI Report SR-39, Palo Alto, Ca 94304 (1976). WENDT,J. O. L. and STERNLING,C. V., Effect of ammonia in gaseous fuels on NO emissions, d. Air Pollut. Control Assoc. 24 (1974). STERNLING,C. V. and WENt)T,J. O. L., On the oxidation of fuel nitrogen in a diffusion flame, AIChE J. 20 (1974). CARRIER,G. F., FENDELL,F. E. and MARaLE, F. E., The effect of strain rate on diffusion flames, SIAM J. appl. Math. 28, 463 (1975). HAMOR,R. J., SMITh, I. W. and TYLER,R. J., Kinetics of combustion of a pulverized brown coal char between 630 and 2200 K, Comb. Flame, 21, 153-162 (1973). BEER,J. M. and ESSENHIGH,R. H., Reaction rate control in dust flames, Nature, 187, 1106 (1960). SMITH, 1. W. and TYLOR, R. J., Internal burning of pulverized semi-anthracite: The relation between particle structure and reactivity, Fuel, 51, 312 (1972). SMITI~, Ian W., The intrinsic reactivity of carbons to oxygen, Fuel, 57 (1978). SZEKELY,J., EVANS,J. W. and SOHN, H. Y., Gas Solid Reactions, Academic Press, New York (1976). PETERSEN,E. E., Chemical Reaction Analysis, Prentice Hall, Englewood Cliffs, New Jersey (1965). LEVENSPIEL,O., Chemical Reaction Engineering, Second Edition, John Wiley & Sons, Inc., New York (1972). CARBERRY, J. J., Chemical and Catalytic Reaction Enyineering, McGraw Hill, New York (1976). AYLING,A. B. and SMITH,I. W., Measured temperatures of burning pulverized-fuel particles, and the nature of the primary reaction product, Comb. Flame, 18 (1972). FLAGAN,R. C., Submicron particles from coal combustion, Seventeenth Symp. (Int.) on Combustion, p. 97, The Combustion Institute, Pittsburgh, Pa 11979). NOVAKOV,T., CHANG,S. G. and HARKER,A. B., Sulfates as pollution particulates : Catalytic formation on carbon (soot) particles, Science, 186, 259-261 (1974). WENDT,J. O. L. and SCHULZE,O. E., On the fate of fuel nitrogen during char coal combustion, AIChE J. 22 (1976). PEREIRA,F. J., BEER,J. M., Gmas, B. and HEDLEY,A. B., NOx emissions from fluidized-bed coal combustors, Fifteenth Symp. Ont.) on Combustion, The Combustion Institute, Pittsburgh, Pa (1975).

24. VAN KREVELEN,D. W. and SCHUGER,A., Coal Science, Elsevier Publishing Company, Amsterdam (1957). 25. LOWRY, H. H. (Ed.) Chemistry of Coal Utilization: Supplementary Volume, John Wiley (1966). 26. BLAIR, D. W., WENDT, J. O. L. and BARTOK, W., Evolution of nitrogen and other species during controlled pyrolysis of coal, Sixteenth Symp. (Int.) on Combustion, p. 475; The Combustion Institute, Pittsburgh, Pa (1977). 27. SMOOT, L. D., HORTON, M. D. and WtLUAmS, G. A., Propagation of laminar pulverized coal-air flames, Sixteenth Syrup. (Int.) on Combustion, p. 375, The Combustion Institute, Pittsburgh, Pa (1977). 28. STYLES,A. C. and CHIGIER,N. A., Combustion of air blast ~tomized spray flames, Sixteenth Syrup. (Int.) on Combustion, p. 619, The Combustion Institute, Pittsburgh, Pa (1977). 29. BRACCO, F. V., Nitric oxide formation in droplet diffusion flames, Fourteenth Syrup. (Int.) on Combustion, p. 831, The Combustion Institute, Pittsburgh, Pa (1977). 30. StrOBERG, E. M., PETERS, W. A. and HOWARD, J. B., Product compositions and formation kinetics in rapid pyrolysis of pulverized coal, Seventeenth Syrup. (Int.) on Combustion, p. 117, The Combustion Institute, Pittsburgh, Pa (1979). 31. BAOZOICH, S. and HAWKSLEY,P. G. W., Kinetics of thermal decomposition of pulverized coal particles, Ind. Engng Chem. Process Design and Development, 9, 4 (1970). 32. PERSHING,D. W. and WENDT, J. O. L., Relative contributions of volatile nitrogen and char nitrogen to NOx emissions from pulverized coal flames, 1 & E.C. Process Design and Development, 18, 60 (1979). 33. PeRSmNG, D. W., Parametric studies on particle size, fuel-air mixing and fuel properties, Paper presented at EPA Conf. on Coal Combustion Technology and Emission Control, California Institute of Technology, Pasadena, Ca, 5-7 February (1979). (Proceedings to be published). 34. BLAIR, D. W., Evolution of coal nitrogen, Paper presented at EPA Conf. on Coal Combustion Technology and Emission Control, California Institute of Technology, Pasadena, Ca, 5-7 February (1979). (Proceedings to be published). 35. SOLOMAN, P. R. and COLKET,M. B., Coal devolatilization, Seventeenth Symp. (Int.) on Combustion, p. 131, The Combustion Institute, Pittsburgh, Pa (1979). 36. UBHAYAKAR, S. K., STICKLER,D. B., VoN ROSENBERG,C. W. and GANNON,R. E., Rapid devolatilization of pulverized coal in hot combustion gases, Sixteenth Syrup. (Int.) on Combustion, p. 427, The Combustion Institute, Pittsburgh, Pa (1977). 37. SOLOMAN, P. R. and COLKET,M. B., The evolution of fuel nitrogen in coal devolatilization, Fuel, 57, 749 (1978). (Also, Paper presented at DOE Contractors Meeting, Morgantown, W. Va., 22-23 October (1979). 38. POHL, J. H. and SAROFIM,A. F., Devolatilization and oxidation of coal nitrogen, Sixteenth Symp. (Int.) on Combustion, p. 491, The Combustion Institute, Pittsburgh, Pa (1977). 39. KOBAYASHI, H., HOWARD,J. B. and SAROFIM, A. F., Coal devolatilization at high temperatures, Sixteenth Symp. (Int.) on Combustion, p. 411, The Combustion Institute, Pittsburgh, Pa (1977). 40. MORLEY,C., The formation and destruction of hydrogen

41. 42.

43.

44.

45.

46. 47. 48. 49. 50. 51. 52. 53. 54. 55. 56. 57. 58. 59. 60. 61.

(Manuscript received 20 December 1979)

You might also like